content
stringlengths 1
15.9M
|
---|
\section*{Acknowledgment}
The authors would like to thank H. Fukuyama for valuable comments.
This work was supported by Grant-in-Aid for Scientific Research
(Nos. 20110002, 20110004, 21740270 and 22540329) from the Ministry of Education, Culture, Sports, Science
and Technology
and by Nara Women's University Intramural Grant for Project Research.
|
\section{Introduction}
Recently, an interesting new variational principle was proposed by Goenner as an alternative to the so called Palatini variational method \cite{Goenner:2010tr}. In the Palatini method, the metric and the connection are treated as independent variables, according to {\it the metric-affine variational principle}, in contrast to the usual metric variation where it is assumed a priori that the connection is the Levi-Civita connection computed from the metric \cite{Deser:2006ht}. In Goenner's variant of this principle, the connection is taken to be the Levi-Civita connection of another metric. Contrary to the claims made in Ref.\cite{Goenner:2010tr} however, both variants of the Palatini method yield an equivalent theory when applied to an $f(\R)$ action.
The concept of a connection as a fundamental variational degree of freedom but subjected to metric-compatibility appears to us somewhat vague. It remains unclear how to apply this restriction of variation in practice to a generic action, whilst the problem is absent in the $f(\R)$ case at the classical level since the extremals of the action turn out a posteriori to belong to a metric subspace. In any case, it is natural and technically straightforward to instead promote the underlying metric to a fundamental field with independent variations. For an obvious reason, we call this approach {\it the bimetric variational principle}. This indeed results in theories that are different from the corresponding metric as well as metric-affine ones, even in the case of Einstein-Hilbert action
where the two latter approaches are both well-known to lead to general relativity.
It is illuminating to view this from the wider perspective of C-theories \cite{Amendola:2010bk}. This is a unified framework which includes as special cases all theories that emerge from the metric and the Palatini variational methods, and describes also Goenner's variant of the Palatini principle, or alternatively, its bimetric version, at a certain decoupling limit.
As far as we know, Ref. \cite{Amendola:2010bk} was the first application of a bimetric variational principle, or ''formulation'' for short. Here we point out a subtlety which was missed there: whereas the metric-affine formulation of C-theories (trivially) reduces to the usual Palatini theory in the special decoupling limit, the bimetric formulation in general doesn't. The latter carries additional dynamics due to the second order property of the field equation. We also go further to generalize these results to actions beyond $f(\R)$.
\section{On metric-affine variations}
In order to clarify these issues, it is useful to first distinguish two different roles of a connection in gravitational theories. In the definition of the basic curvature object, the Riemann tensor, $\hat{R}^\alpha_{\beta\gamma\delta}$, one refers to a connection $\hat{\Gamma}$ as
\be \label{riemann}
\hat{R}^\alpha_{\beta\gamma\delta}=\hat{\Gamma}^\alpha_{\beta\gamma,\delta}-
\hat{\Gamma}^\alpha_{\beta\delta,\gamma}+
\hat{\Gamma}^\alpha_{\lambda\gamma}\hat{\Gamma}^\lambda_{\beta\delta} -
\hat{\Gamma}^\alpha_{\lambda\delta}\hat{\Gamma}^\lambda_{\beta\gamma}\,.
\ee
This tensor can be constructed in a metric-independent way by parallel transporting a vector $v$ around a closed curve, $[\hat{\nabla}_\mu,\hat{\nabla}_{\nu}]v^{\alpha}=\hat{R}^\alpha_{\gamma\mu\nu}v^\gamma$. The Riemann tensor is needed to compose invariants of the curvature. Thus, the connection $\hat{\Gamma}$ inevitably enters any covariant gravitational action and so affects the form of the left hand side of the gravitational field equations ensuing from that action. We therefore call $\hat{\Gamma}$ the {\it geometric connection}.
A crucial point is that, as long as there is a metric, another connection is always present. This is of course the Levi-Civita connection $\Gamma$ generated by the metric as
\be \label{matter_c}
\Gamma^\alpha_{\beta\gamma} = \frac{1}{2}g^{\alpha\la}\lp g_{\beta\la,\ga}+g_{\la\ga,\bt}-g_{\bt\ga,\la}\rp\,.
\ee
In the gravity theories we are discussing the metric is formally providing the invariant volume element and the contractions of vectors in the matter action. The metric thence provides measure of physical distances. Consequently, the trajectories of material particles are the shortest paths with respect to the metric. Equivalently, they follow the geodesics determined by the connection (\ref{matter_c}), which we hence call the {\it matter connection}. This result, implied by the so called generalized Bianchi identity that is a consequence of the diffeomorphism invariance of the action, is completely independent of the left hand side of the field equations, in particular the geometric connection that generates them \cite{Koivisto:2005yk}.
Using our notation, the starting point of Ref.\cite{Goenner:2010tr} may then be written as\footnote{In the notation of Ref.\cite{Goenner:2010tr}, $\hat{g}_{\mu\nu} \rightarrow g_{\mu\nu}$, $g_{\mu\nu} \rightarrow h_{\mu\nu}$ and $\hat{\Gamma} \rightarrow \{\}_g$. The two latter fields are the variational degrees of freedom in the approach of Ref.\cite{Goenner:2010tr}.}
\be \label{action}
S = \int d^4 x \sqrt{-g}\lb f(\R) + \mathcal{L}_m(\Psi,g_{\mu\nu})\rb\,,
\ee
where $\R=g^{\bt\delta}\hat{R}_{\bt\delta}$ and the tensor $\hat{R}_{\bt\delta}$ is understood to be a functional of an independent metric $\hat{g}_{\mu\nu}$. More precisely, it is given by the expression (\ref{riemann}) when $\alpha=\gamma$, and
\be \label{geometric_c}
\hat{\Gamma}^\alpha_{\beta\gamma} = \frac{1}{2}\hat{g}^{\alpha\la}\lp \hat{g}_{\beta\la,\ga}+\hat{g}_{\la\ga,\bt}-\hat{g}_{\bt\ga,\la}\rp\,.
\ee
Matter fields are collectively denoted by $\Psi$. The significance of the metric $g_{\mu\nu}$ should be clear from the discussion above.
In fact, with respect to this metric, we have a metric theory of gravitation satisfying the three requirements that \cite{lrr-2006-3}
1) there exists a symmetric metric,
2) test bodies follow geodesics of the metric, and
3) in local Lorentz frames, the non-gravitational laws of physics are those of special relativity. On the other hand, then the independent metric $\hat{g}_{\mu\nu}$ may in a sense be regarded as an auxiliary field. It can be algebraically eliminated in favor of the observable metric $g_{\mu\nu}$, and it is most transparent to write the field equations in terms of the latter.
The opposite viewpoint is adopted in Ref.\cite{Goenner:2010tr}. The metric $g_{\mu\nu}$ is interpreted there as an auxiliary field devoid of physical significance and the field equations are written in terms of the other metric. Despite of being presented as a novel alternative to the previously considered gravity theories, the ensuing theory is nothing but the Palatini version of $f(\R)$ disguised in the Einstein frame. It was claimed that the equivalence principle is broken since matter stress tensor is not covariantly conserved with respect to $\hat{\Gamma}$, and that no matter gradients appear in the theory. The stress energy is however covariantly conserved with respect to $\Gamma$, and when written in terms of the metric that is compatible with this connection, the matter gradients appear explicitly into the theory. In this light, the misleading conclusions reached in Ref.\cite{Goenner:2010tr} stem from a confusion regarding the physical roles of the two connections, or in this case equivalently, the associated metrics\footnote{It is formally possible to identify the Einstein frame as the physical frame, and then the equivalence principle indeed appears to be violated. The Einstein frame version of Palatini-$f(\R)$ gravity (that is ''dynamically equivalent'' to the Jordan frame version) has been also considered in the literature \cite{Olmo:2011uz}. We are not concerned here with the choice of frame, which is a separate issue from the result that the two variational principles yield equivalent theories.}.
The rationale of C-theories, recently introduced by Amendola {\it et al} \cite{Amendola:2010bk}, is the possibility of a nontrivial relation between the matter and the geometric connections. As a simple example we first consider the following conformal relation: $\hat{g}_{\mu\nu}=C(\R)g_{\mu\nu}$. In particular, we look at a specific class of C-theory actions parameterized by $\al \in \mathbb{R}$ in such a way that when the parameter $\al=0$, the corresponding metric, and when $\al=1$, the corresponding Palatini theory is reproduced for an arbitrary function $f(\R)$. An example of such $\alpha$-parameterization is given by the exponential interpolating function as
\be \label{ctheory_c}
S_\alpha = S+\int d^4x\sqrt{-g}\la^{\mu\nu}_\rho\lp \hat{\Gamma}^\rho_{\mu\nu}-
\left\{ {\rho \atop \mu\nu} \right\}_{\lp f'(\R)\rp^\alpha g}\rp\,.
\ee
Thus, we add to the action (\ref{action}) a Lagrange multiplier which constraints the geometric connection to be the Levi-Civita connection of the metric $\hat{g}_{\mu\nu}=\lp f'(\R)\rp^\alpha g_{\mu\nu}$. When $\al=0$, the two connections coincide and we have a metric theory.
When $\al=1$, the geometric connection is in the Einstein frame, which is the peculiar relation of Palatini-$f(\R)$ theories. However, in general also the lagrangian multiplier contributes to the dynamics. This can be of course changed by rescaling $\la^{\mu\nu}_\rho\rightarrow (1-\al)\la^{\mu\nu}_\rho$, and then we recover precisely Goenner's action (\ref{action}) at the limit $\al=1$. This is a discontinuous limit of the theory, since a degree of freedom becomes nondynamical there. This decoupling is the culprit for the pathology of the $\omega=-3/2$ Brans-Dicke theory that was discovered decades ago and whose disturbing consequences have surfaced in many different contexts more recently \cite{Olmo:2011uz}.
The action (\ref{ctheory_c}), due to the presence of the lagrange multiplier may display no discontinuity in the propagating degrees of freedom in the limit $\al=1$. Still, the peculiar relation $C(\R)=f'(\R)$ holds and furthermore, $\la=0$ is always a consistent solution of this version of the theory. Thus the solutions of theory (\ref{action}) form a subset of the solutions of the new theory, which nevertheless seems to avoid the notorious theoretical and observational problems of the former. Hence the action (\ref{ctheory_c}) at $\alpha=1$ can realize the motivation of Ref. \cite{Goenner:2010tr} by introducing a phenomenologically viable alternative to the Palatini method.
\section{On bimetric variations}
As discussed in the introduction, one can also consider the set-up where the metric $\hat{g}_{\mu\nu}$ rather than the connection $\hat{\Gamma}$ is the fundamental field \cite{Amendola:2010bk}.
Instead of the action (\ref{ctheory_c}), where the gravitational degrees of freedom consisted of the triplet $({g}_{\mu\nu},\hat{\Gamma},\lambda)$, one would write the action
\be \label{ctheory}
S_\alpha = \int d^4x\sqrt{-g}\lb f(\R) + \la^{\mu\nu}\lp \hat{g}_{\mu\nu}-\lp f'(\R)\rp^\alpha g_{\mu\nu} \rp\rb\,,
\ee
the independent fields being now $({g}_{\mu\nu},\hat{\Gamma},\lambda)$. By erasing the lagrangian multiplier constraint one then obtains a bimetric reformulation of Goenner's starting point (\ref{action}).
One may also ask whether the conclusions persist beyond the $f(\R)$ theories. In particular, one could suspect that the conformal relation appearing in this special class of theories is necessary to guarantee the degeneracy of the two variational methods. Starting from more general forms of action, one obtains a more complicated relation between the independent and the metric connection by applying the metric-affine variational principle \cite{Olmo:2011uz}. In the remainder of this communication, we will generalize the bimetric variational principle and $C$-theory field equations to such actions.
For this purpose, we write the action
\be \label{action2}
S = \int d^4 x \sqrt{-g}\lb f(\R,\Q) + \mathcal{L}_m(\Psi,g_{\mu\nu})\rb\,,
\ee
allowing in there the invariant $\Q$ constructed from the two metrics via
\be
\Q = g^{\alpha\beta}g^{\gamma\delta}\hat{R}_{\al\ga}\hat{R}_{\bt\delta}\,,
\ee
where again $\hat{R}_{\mu\nu}=\hat{R}_{\mu\nu}[\hat{\Gamma}[\hat{g}]]$ when taking into account Eqs. (\ref{riemann}) and (\ref{geometric_c}).
The field equations for the metric $g_{\mu\nu}$ are
\be \label{field2}
\lp f_{,\R}\hat{R}_{\mu\nu}+2f_{,\Q}\hat{R}_{\mu\alpha}\hat{R}^{\nu\alpha} \rp - \frac{1}{2}f g_{\mu\nu} = T_{\mu\nu}\,,
\ee
where $T_{\mu\nu}$ is the matter stress energy tensor. The variation with respect to the metric $\hat{g}_{\mu\nu}$ yields
\be \label{field3}
\hat{D}^{\mu\nu}_ {\al\bt}\lb \sqrt{\frac{{g}}{\hat{g}}}\lp f_{,\R}g^{\al\bt} + 2f_{,\Q}\hat{R}^{\al\bt}\rp\rb = 0\,,
\ee
where we have defined the differential operator
\ba
\hat{D}^{\mu\nu}_{\al\bt} =
\frac{1}{2}\sqrt{\frac{\hat{g}}{g}}\Big(\hat{g}^{\mu\nu}\delta^\rho_\alpha\delta^\gamma_\beta
& + & \hat{g}^{\rho\gamma}\delta^{\mu}_\alpha\delta^\nu_\beta \nonumber \\
- \hat{g}^{\rho\nu}\delta^{\mu}_\al\delta^{\gamma}_\bt & - & \hat{g}^{\rho\nu}\delta^{\mu}_\bt\delta^{\gamma}_\al
\Big) \hat{\nabla}_{\gamma}\hat{\nabla}_\rho\,.
\ea
The equation (\ref{field3}) is solved by
\be \label{disformal}
\hat{g}_{\mu\nu} = f_{,\R}g_{\mu\nu} + 2f_{,\Q}\hat{R}_{\mu\nu} \equiv h_{\mu\nu}\,.
\ee
This is the same disformal relation one obtains \cite{Olmo:2011uz} in the metric-affine variation of the action corresponding to (\ref{action2}).
Since, obviously in the usual Palatini version of the theory the field equations have the same the form (\ref{field2}), the dynamics of the solution (\ref{disformal}) coincide with those\footnote{
We assume that in the metric-affine variation of the action (\ref{action2}) one is restricted to symmetric connections and Ricci tensors. Furthermore, in the bimetric variations discussed here, we assume the metric $\hat{g}_{\mu\nu}$ to be symmetric. Though, since this metric is related to the spin connection aspects of geometry and not to physical distances, it would be meaningful to relax this assumption, we omit exploring the possibility here.}.
There is however the following subtlety. Unlike from a metric-affine variation which yields first order equations of motion, we now obtained a second order equation which may have different solutions. Indeed, Eq.(\ref{field3}) appears to allow solutions where the nonmetricity of the connection $\hat{\Gamma}$ with respect to the metric $h_{\mu\nu}$ defined by (\ref{disformal}) is nonvanishing $Q_{\la\al\bt}=-\hat{\nabla}_\la h_{\al\bt} \neq 0$. By writing open eq.(\ref{field3}), one gets a nontrivial differential constraint on nonmetricity. In general then, the solutions of the theory need not coincide with the Palatini theory, either in its usual or in its C-theory form, since the connection need not be the Levi-Civita connection of $h_{\mu\nu}$. Even in the Einstein-Hilbert case, when $h_{\mu\nu}=g_{\mu\nu}$, there can be nonmetric solutions. They are classically distinguishable from general relativity (or from more general Palatini theories when $f\neq \R$) if the difference of the metrics doesn't correspond to a projective transformation of the geometric connection.
In general from Eq.(\ref{riemann}) one gets that
\be
\hat{R}_{\al\bt} = {R}_{\al\bt}(h)+\hat{\nabla}_\mu\Delta^\mu_{\al\bt}-\hat{\nabla}_\bt\Delta^\mu_{\mu\al}
+ \Delta^\la_{\la\rho}\Delta^\rho_{\al\bt}
- \Delta^\la_{\al\rho}\Delta^\rho_{\bt\la}\,.
\ee
where in the case at hand we have
\be
\Delta^\rho_{\al\bt} = \frac{1}{2}g^{\rho\la}\lp Q_{\al\bt\la}+Q_{\bt\al\la}-Q_{\la\al\bt}\rp\,.
\ee
Let us demonstrate explicitly the appearance of nonmetricity in the general relativistic Einstein-Hilbert action $f=\R$ within the bimetric formulation. For simplicity, we assume conformal nonmetricity described the function $b$ as $Q_{\la\al\bt}=-\hat{\nabla}_\la g_{\al\bt} = b_{,\la}g_{\al\bt}$. The field equations (\ref{field2}) and (\ref{field3}) for the two metrics $g_{\mu\nu}$ and $\hat{g}_{\mu\nu}$ become now, respectively
\be
G_{\mu\nu}+\frac{1}{2}b_{,\mu,}b_{,\nu}-\nabla_{\mu}b_{,\nu}+\lp \Box b+\frac{1}{4}(\partial b)^2\rp g_{\mu\nu} = T_{\mu\nu}\,, \label{1st}
\ee
\be \label{2nd}
\nabla_\mu b_{,\nu} - 2b_{,\mu}b_{,\nu} -
\lp \Box b - \frac{1}{2}(\partial b)^2 \rp g_{\mu\nu} = 0\,.
\ee
There is no matter source in the right hand side of (\ref{2nd}) since we have assumed that matter is minimally coupled to geometry.
In fact by using the trace of the latter equation and rescaling $b=\sqrt{2/3}\phi$, one confirms the expectation that conformal nonmetricity contributes a canonic massless scalar obeying $\Box\phi=0$,
\be
G_{\mu\nu} - \phi_{,\mu,}\phi_{,\nu}+\frac{1}{2}(\partial \phi)^2 g_{\mu\nu} = T_{\mu\nu}\,. \label{1st_b}
\ee
Even without sources we can have propagating nonmetricity in the bimetric formulation of the action $f=\R$. In this sense the resulting theory is richer than in the metric or in the metric-affine formulation.
\begin{center}
\begin{table}[t]
\begin{tabular}{|c||c|c|c| }
\hline
C-theory & $\al \rightarrow 0$ & $\al \rightarrow 1$ & $\la \rightarrow 0$ \\
\hline \hline Metric-affine & metric & (1st order ?) gen. P. & Goenner $\sim$ P. \\
\hline Bimetric & metric & (2nd order ?) gen. P. & New theories \\
\hline
\end{tabular}
\caption{A summary of our findings at the various limits of C-theories.
Metric-affine and bimetric formulations are considered. If we impose the metric constraint $\al=0$, the metric field equations follow in both cases.
We refer to theories where the $\hat{g}_{\mu\nu}=h_{\mu\nu}$ but there are more dynamics than in the usual Palatini theories as ''gen. P.''. In the decoupling limit $\la \rightarrow 0$ of the bimetric formalism, we may have solutions with $\hat{g}_{\mu\nu} \neq h_{\mu\nu}$. Such solutions are absent in the metric-affine formalism at this limit, where the theory reduces to the usual Palatini theory. \label{tab}}
\end{table}
\end{center}
Finally, we look also at the C-theory generalization of the theory (\ref{action2}) in the form
\be \label{ctheory_g}
S = \int d^4 x \sqrt{-g}\lb f + \hat{\lambda} - C\la - D\hat{\La} + \mathcal{L}_m(\Psi,g_{\mu\nu})\rb\,,
\ee
where for brevity of notation, we have suppressed the explicit dependence of the three functions $f$, $C$ and $D$ on the two scalars $\R$ and $\Q$, and introduced the invariants constructed by contractions of the field $\la^{\mu\nu}$ as
\be
\la = \la^{\mu\nu}g_{\mu\nu}\,, \quad
\hat{\la} = \la^{\mu\nu}\hat{g}_{\mu\nu}\,, \quad
\hat{\La} = \la^{\mu\nu}\hat{R}_{\mu\nu}\,.
\ee
Variations with respect to the three tensor fields $g^{\mu\nu}$, $\hat{g}_{\mu\nu}$ and $\la^{\mu\nu}$ give, respectively, the field equations
\ba
T_{\mu\nu} & = & L_\R \hat{R}_{\mu\nu} + L_\Q \hat{R}_{\mu\al}\hat{R}^{\al}_{\ph{\al}\nu}-\frac{1}{2}L g_{\mu\nu} + C\la_{\mu\nu}\,, \\
\la^{\mu\nu} & = & \hat{D}^{\mu\nu}_{\al\bt}\lb \sqrt{\frac{g}{\hat{g}}}\lp L_\R
g^{\al\bt}+2L_\Q\hat{R}^{\al\bt}-D\la^{\al\bt}
\rp\rb\,, \label{fe2} \\
\hat{g}_{\mu\nu} &= & Cg_{\mu\nu}+D\hat{R}_{\mu\nu}\,. \label{fe3}
\ea
We have employed the shorthand notations for the gravity lagrangian and its derivatives with respect to the two scalars
\ba
L &=& f + \hat{\lambda} - C\la - D\hat{\La}\,, \\
L_\R & = & f_{,\R} - C_{,\R}\la - D_{,\R}\hat{\La}\,, \\
L_\Q & = & f_{,\Q} - C_{,\Q}\la - D_{,\Q}\hat{\La}\,.
\ea
Now the constraint (\ref{fe3}) dictates the relation of the metrics without any ambiguity. However, we should take into account that the lagrange multiplier has also other nontrivial consequences.
In the metric case $C=1$, $D=0$, $\la_{\mu\nu}$ contributes to the field equations in such a way that they reduce to the field equations one obtains from pure metric variation. In particular, $L=\R$ yields then pure Einstein gravity.
Choosing $C$ and $D$ suitably, we can obtain $\hat{g}_{\mu\nu} = h_{\mu\nu}$, as in (\ref{disformal}).
Also now, as in the metric-affine case above, the lagrangian multiplier can also contribute to the field equations.
Its equation of motion is given by (\ref{fe2}). In the $f(\R)$ case, these theories can be described as biscalar tensor gravity, which do not in general reduce to the $\omega_{BD}=-3/2$ theory when $C=f'(\R)$.
We can thus straightforwardly construct C-theories which generalize (\ref{ctheory}) and interpolate between the modified Palatini theories and the metric theories. A possible choice is $C= f_{,\R}^\alpha$ and $D= \alpha f_{,\Q}$.
The construction of the metric-affine C-theory generalizing (\ref{ctheory_c}) is completely analogous. There the rescaled theory $\lambda\rightarrow (1-\al)\la$ reduces to the usual Palatini theory at the decoupling limit $\al = 1$, but to recapitulate, in the bimetric framework this limit is different from both the usual Palatini and its C-theory generalisations.
\section{Conclusions}
We clarified the motivations and implications of recently introduced variational principles in gravity. Some erronous conclusions concerning (in)equivalences between theories were pointed out.
Starting from the distinction between the matter and the geometric connections, one may assume either the metric-affine or the bimetric formulation depending on whether the geometric connection or its underlying metric is the fundamental degree of freedom. We observed that in the latter case, even the Einstein-Hilbert action can support propagating torsion and nonmetricity, opening the possibility to observationally distinguish the correct fundamental assumption.
Furthermore, either formulation can be generalized by allowing an arbitrary relation between the two connections a'la C-theories. In particular, one can then obtain theories where the geometric connection is in the Einstein frame but which could avoid the worst problems of the usual Palatini theories. A summary of the different variants of the Palatini theory as special limits of C-theories is presented in the table \ref{tab}.
|
\section{Introduction}
The guiding principle of Einstein's principle of general relativity is that
the laws of physics should be expressed in a form that is independent of the
coordinate frame in which one might choose to express them. Weyl generalized
this concept in the realm of quantum physics by the Dirac operator for \ a
field by introducing a Riemann connection for space-time, so that the partial
derivative is replaced with a covariant derivative which yields a new Dirac
operator\cite{Frankel2004}
\begin{equation}
\bar{\partial}\Psi+\frac{1}{4}\Gamma_{ik}^{j}\gamma^{i}\gamma^{k}\Psi.
\end{equation}
Note $\Psi$ is the field, while $\Gamma\Psi$ is the \emph{interaction} term
that tells us how the gravity field interacts with the electron. When one can
neglect the gravitational interaction, \ one can consider a change of field
coordinates $\Psi$ rather than space-time coordinates, e.g. \emph{gauge
transformations.} (For a history, this summary is based upon the overview by
Gross\cite{Gross1992} .)
Quantum mechanics assigns a physical meaning to the absolute value $\left\vert
\Psi^{\alpha}\right\vert $, however it does not assign a meaning to the phase
of $\Psi^{\alpha}$, so $\Psi^{\alpha}=\left\vert \Psi^{\alpha}\right\vert
e^{i\theta_{\alpha}}$, and $\theta_{\alpha}$, the \emph{phase}, has always
been taken to have no physical meaning. Both the Dirac equation and Lagrangian
are invariant under transformations of the form:
\begin{equation}
\Psi\longmapsto e^{i\theta_{\alpha}}\Psi.
\end{equation}
This is a \emph{global} transformation since $\alpha$ is constant. A global
transformation is an invariance of the Dirac equation. Besides a global
transformation, physicists have also considered the effect of local
transformations of $\alpha$, local in this case means that $\alpha=\alpha(x)$
so it depends on the particular space-time coordinates. \ When coordinates are
included in $\alpha$, the Dirac equation and Lagrangian no longer are
invariant under transformations because a $d\alpha$ appears in the
transformation equations. Thus, one could conclude that there is a background
field that interacts with the electron. One can go through the usual argument
that associates a change of frame with a complex line bundle which leads
naturally to the association of the background field with electromagnetism and
yet another Dirac operator:
\begin{equation}
\bar{\partial}_{A}\triangleq\gamma^{i}\left( \partial_{i}+\omega_{j}\right)
=\bar{\partial}-ie\gamma^{i}A_{j}.
\end{equation}
Quantum mechanics mandates that $A$ be considered as a separate field, so one
gets a new Lagrangian associated with $A$. This brief review of the power of
invariance associated with the gauge concept now allows us to consider another
type of invariance that previously seems to have been over looked.
A gauge transformation can often be associated with the evolutionary path of a
quantum mechanical object. This is not observable directly, a difference can
be such as when a quantum mechanical object can take distinct paths, and then
be recombined to demonstrate a phase difference such as in the Aharonov-Bohm
effect (A-B effect). This has been interpreted as a geometric phase in a
number of papers by Aharonov and various collaborators. Aharonov's career can
be viewed as largely devoted to elucidating what phase means in quantum
mechanics. The wavefunction is invariant under $U(1)$ gauge transformations,
so the wavefunction can be written with a constant relative phase
function,$\Lambda$, so we have:
\begin{equation}
\Psi\left( x,t\right) \rightarrow e^{i\Lambda}\Psi\left( x,t\right) .
\end{equation}
This can be interpreted as associating a memory with a path. In general, the
phase can depend upon the topological features of the phase such as associated
with the A-B effect\cite{Aharonov1959} . There is no reason, however, to limit
$\Lambda$, it can be an arbitrary function $\Lambda\left( x\right) $ or it
could also be considered a random variable with an underlying probability
density function (PDF).
\section{Random Gauge Transformation}
Under some circumstances this "randomization" of $\hat{\Lambda}$ could be used
to explicitly bring in randomness associated with a path with an underlying
PDF. Thus, it is possible to define a "random gauge" in quantum mechanics.
Random path-dependent effects become a stochastic memory associated with the
quantum mechanical system. We explore this is some detail and suggest an
experiment to test the effects of randomness on path dependent phase
difference effects.
Since phase has no physical meaning, there is no reason to associate it with a
fixed parameter $\theta_{\alpha}$, instead one could assume that phase to be a
random variable $\hat{\theta}_{\alpha}$ with an underlying PDF $P_{\theta
_{\alpha}}(\theta)$. Then one could claim that $\Psi$, when interpreted as
either a field coordinate or $\left\vert \Psi^{\alpha}\right\vert ,$ has the
usual Born probability interpretation in conventional quantum mechanics and
must be invariant under a random gauge transformation:
\begin{equation}
\hat{\Psi}\longmapsto e^{i\hat{\theta}_{\alpha}}\Psi.
\end{equation}
This is a new type of invariance to consider, for both the Schrodinger
equation or the equations of quantum field theory. This type of transform is
define to be a \textbf{type 1} \emph{random gauge transformation}. \ \ If one
associates randomness with gauge invariance rather than with $\Psi$ directly,
it opens some alternative ways to think about the interpretation of quantum
mechanics, which will be examined in a subsequent paper.
One is immediately led to the question of meaning for this \textbf{type 1}
random gauge transformation. \ One concept of random phase can be connected to
the idea of what a "path" could mean in quantum mechanics. Empirically a path
is associated with the propagation of disturbance which has an intensity
pattern that can localized to a measurement device located at a point $x$ and
a time $t$. It is given that propagation of a disturbance within that medium
depends on the positions of the point and the direction of propagation of the
disturbance. Based on the wave nature of both mechanical and optical waves, a
principle was proposed by Huygens' (and latter others) that explains how wave
fronts propagate in time\cite{Baker1987} . \emph{Huygens' Principle} underpins
the analysis of classical propagation problems in electromagnetic and
acoustics\cite{Beckmann1987} , and much of modern physics\cite{Feynman2010} .
Empirically, the propagation of disturbances at a given point in a medium
depends on the position of the point and the direction of propagation of the
disturbance. This means that\cite{Gelfand2000} :
\begin{enumerate}
\item \textit{Each point can be in one of two states: excited or at rest. No
concept of intensity of the distribution needs to be introduced.}
\item \textit{If a disturbance arrives at the point }$P$\textit{\ at the time
}$t_{2}$\textit{, then when starting at time }$t_{1}$\textit{\ }$(t_{2}\geq
t_{1})$\textit{, the point }$P$\textit{\ itself serves as a source of
disturbances propagating in the medium.}
\end{enumerate}
The mathematical form of Huygens' Principle can be expressed in a number of
different ways. We choose the following:
\begin{definition}
(Huygens' Principle): Given a wave front $\Psi(x,t)$, it is the propagator of
small waves which collectively make up the next wave front. The mathematical
form of this statement is:
\begin{equation}
\Psi(x,t_{2})=\int G(x,y)\,\Psi(y,t_{1})\,dy.\;\;\;\;\;\;(t_{2}>t_{1})
\end{equation}
$G(x,y)\doteq$ kernel of the propagator.
\end{definition}
The intensity pattern of the path at a latter time is the superposition of all
possible paths from $x$ to $y$. Now since there is no reason to choose a
particular path, we can associate it with the random gauge transformation
$\hat{\theta}_{\alpha}$ so we have $\hat{\theta}_{\alpha}=\hat{\theta
_{\alpha}(x)$. Thus, because there is a random gauge associated with $\Psi$,
there is an underling PDF associated with it as well. Our generalization is
based on some work in scattering theory and other applied physics
applications. Methods of determining the probability density function (PDF)
are well known for the transformation:
\begin{equation}
\hat{x}=a\sin\hat{\theta
\end{equation}
where $a$ is a deterministic parameter and $\hat{\theta}$ is random variable
with a uniform PDF\cite{Papoulis1991} $f_{\theta}(\hat{\theta})$. The PDF's
for this distribution and others are found in the Appendix. Given the
characterization of the probability density function of the angle it is
possible to determine the moments of the random variable:
\begin{equation}
\hat{y}_{c}=\sum_{i=1}^{N}\hat{r}_{i}\cos\hat{\theta}_{i
\end{equation}
or
\begin{equation}
\hat{y}_{s}=\sum_{i=1}^{N}\hat{r}_{i}\sin\hat{\theta}_{i
\end{equation}
for an arbitrary $\hat{r}_{i}$ and $\hat{\theta}_{i}$ provided their
characteristic functions exist\cite{Gray} . (Note these sums could also be
viewed as discrete random Fourier series\cite{Lanczos1975}\cite{Marcus1981} .)
Now, the two sums can be written as:
\begin{equation}
\hat{z}=\hat{y}_{c}+i\hat{y}_{s}=\sum_{j=1}^{N}\hat{A}_{j}e^{i\hat{\theta
_{j}
\end{equation}
which is the same mathematical form as the random gauge transformation
$\hat{\Psi}\longmapsto\sum_{\alpha}\hat{A}_{\alpha}e^{i\hat{\theta}_{\alpha
}\Psi$. Thus we have a means for associating different types of randomness
with the phase variable; some examples of such probability distributions are
the uniform distributions, Cauchy distributions, Normal distributions, etc. In
general, the mean $\left\langle \cdot\right\rangle $ of $\hat{z}$ can be
written as:
\begin{align}
\left\langle \hat{z}\right\rangle & =\left\langle \hat{y}_{c}\right\rangle
+i\left\langle \hat{y}_{s}\right\rangle =\sum_{j=1}^{N}\left\langle \hat
{A}_{j}\right\rangle \left\langle e^{i\hat{\theta}_{j}}\right\rangle
\nonumber\\
& =\sum_{j=1}^{N}\left\langle \hat{A}_{j}\right\rangle \left\langle \cos
\hat{\theta}_{j}\right\rangle +i\sum_{j=1}^{N}\left\langle \hat{A
_{j}\right\rangle \left\langle \sin\hat{\theta}_{j}\right\rangle
\end{align}
provided the random variables in the sums are uncorrelated. (The appendix
considers specific PDF's for these angular variables.) Also, we have
\begin{align}
Var\left( \hat{z}\right) & =Var\left( \hat{y}_{c}\right) +iVar\left(
\hat{y}_{s}\right) \nonumber\\
& =\sum_{j=1}^{N}Var\left( \hat{A}_{j}\right) Var\left( \cos\hat{\theta
}_{j}\right) +i\sum_{j=1}^{N}Var\left( \hat{A}_{j}\right) Var\left(
\sin\hat{\theta}_{j}\right) .
\end{align}
Thus, the mathematics of scattering theory can be drawn upon to explore how
the variances in the the paths between two paths can vary for different types
of random gauges. For example, most of the examples of Berry phase can be
readily generalized by exploring the variation in paths as instances of random
gauges. In the next section, we consider how to demonstrate this using a Berry
phase device. We then discuss how this might be related to variation in the
space-time metric as an instance of aspects of gravity as a gauge transformation.
It is possible to convert the $\sum_{i=1}^{N}r_{i}\cos\hat{\varsigma}_{i}$ to
a polynomial function $g(\theta,\varphi)$ using trigonometric identities:
\begin{equation}
\hat{G}(\hat{\theta},\hat{\varphi})=\sum_{j=1}^{N}r_{j}\cos\hat{\varsigma
_{j}.
\end{equation}
This formula can be interpreted as an antenna pattern, since any polynomial
function of variables, $\theta$ and $\varphi$ with $G\left( \theta
,\varphi\right) $ representing the gain of the antenna. In classical
electromagnetics, an antenna is the source to launch an electromagnetic wave
into space, and the antenna pattern characterizes how the electromagnetic wave
are distributed in the angular variables. This suggests a second type of
random gauge transformation. \
A \textbf{Type 2} \emph{random gauge transformation} associates the randomness
with the field coordinates so we have a family of PDF's associated with a
given gauge transformation
\begin{equation}
\hat{\Psi}\longmapsto\exp\left( i\theta\left( \alpha\left( \hat{x}\right)
\right) \right) \Psi.
\end{equation}
It is in this form that Huygens' principle can be interpreted as a statement
in terms of antenna theory. Note this is done by interpreting $G(x,y)$ as
\ the antenna gain, which can be written as a function of angle variables:
\begin{equation}
G=G(\theta,\vartheta)=f(\sin\theta,\cos\vartheta)
\end{equation}
Now each point on a spherical surface of propagation can be treated as a point
source of outward propagating radiation, which, in the language of antenna
theory, is equivalent to the statement that the gain of the antenna is
$G(\theta,\varphi)=\frac{1}{2}$. If we assume that $G(\theta,\varphi)\neq
\frac{1}{2}$, then radiation expands non-uniformly. Thus we have non-radial
expansion of radiation from a point source. Huygens' principle with the gain
equal to one (inward and outward expansion) was more or less an observation
rather than an experimental fact or a theoretical principle etched in stone.
Thus, in line with the antenna theory interpretation one might argue that a
non-spherical expansion should be considered, so this would have the
functional form:
\begin{equation}
G(\theta,\vartheta)=\sum_{j=1}^{m}\sum_{i=1}^{n}a_{i}\cos\vartheta_{i
\sin\theta_{j}.
\end{equation}
The type 2 random gauge can be viewed as a generalization of the concept of an
antenna gain pattern. The field coordinates in quantum field theory can be
used as the source, in the same way the gain pattern of an antenna is, of
random fields that arise in space time.
\section{Physics Instances of Random Gauge Transformations}
The type 2 random gauge concept can be associated with the Feynman's path
integral and by using the random gauge concept in Huygens' principle. \ Given
a wave front $\Psi(x,t)$, which is the propagator of quantum waves, the gauge
acts collectively as a source to constitute the next wave front. The
mathematical form of this statement is:
\begin{equation}
\Psi(\vartheta,t_{2})=\int G(\hat{\theta},\hat{\vartheta})\,\Psi(\theta
,t_{1})\,d\theta.\;\;\;\;\;\;(t_{2}>t_{1})
\end{equation}
$G(\hat{\theta},\hat{\vartheta})\circeq$ random antenna pattern which
propagates $\Psi(\theta,t_{1})$ from the time $t_{1}$ to the time $t_{2}$. The
point of this is the randomness we normally associate with $\Psi$ in quantum
mechanics can instead be associated with mechanism for propagating $\Psi$ in
time if the most general form of gauge invariance is supposed for quantum
mechanics. It also provides a physical interpretation of the path integral by
a natural connection with Huygens' principle becoming the foundational
principle for quantum mechanics. Furthermore, if one takes this definition of
a propagator, then one can use it to explain the gauge invariance of $\Psi$,
and use this fact to explain phase invariance for many of the different
"paths" that can be taken in quantum mechanics as well as such things as the
Aharonov-Bohm effect and Berry phase.
The observation of Aharonov-Bohm phase\cite{Aharonov1959} suggests to one a
means of both interpreting and demonstrating some aspects of the random gauge
concept. In the Figure of the AB-effect, the electron wave function can be
written as\cite{Silverman2002} :
\begin{equation}
\Psi\left( \mathbf{r},t\right) =\Psi_{1}\left( \mathbf{r},t\right)
e^{iS_{1}}+\Psi_{2}\left( \mathbf{r},t\right) e^{iS_{2}},
\end{equation}
where the numbers 1 and 2 refer to the slits and the phases for each path the
electron can take
\begin{equation}
S_{1,2}=\frac{e}{\hbar c}\int_{x_{0}}^{x_{1}}\mathbf{A}_{1.2}\cdot
d\mathbf{l}_{1,2}.
\end{equation}
The phase difference between the two components of the wave function is
\begin{equation}
S_{1}-S_{2}=\frac{e}{\hbar c
{\textstyle\oint\limits_{C}}
\mathbf{A}\cdot d\mathbf{l}=\frac{e\Phi}{\hbar c
\end{equation}
where $C$ is a closed contour about the solenoid and the magnetic flux $\Phi$
is the cumulative amount of flux in the interior of the solenoid. The solenoid
has a current passing through it. Instead of considering a normal current that
induces AB-phase difference, which assumes a current which is proportional to
$NI_{0}\cos\omega_{0}t$, where $N$ is the number of turns in the solenoid and
$\omega_{0}$ is the frequency of the current
\begin{figure}[ptb
\centering
\includegraphics[
natheight=3.634800in,
natwidth=4.802300in,
height=2.034in,
width=2.6783in
{LI4YDE02.bmp
\end{figure}
If we consider the frequency no longer fixed, but disturbed by noise, it is
evident how noise can manifest itself in the AB-effect. If we specifically
modulate the current with noise, $\hat{I}=I_{0}\cos\left( \omega_{0}t+\hat
{n}\right) $, the effect of the noise is made manifest by an additional phase
difference in the phase difference between the two components of the wave
function. We have induced an effective random variation in the gauge path that
is equal to
\begin{equation}
\Delta\hat{S}=\hat{S}_{1}-\hat{S}_{2}=\frac{e\Phi\exp\left( i\hat{\theta
}\right) }{\hbar c}.
\end{equation}
The expected value of $\Delta\hat{S}$, $E\left[ \Delta\hat{S}\right] $, is
proportional to $E\left[ \exp\left( i\hat{\theta}\right) \right] $. For
many, but not all, probability distributions in the appendix, $E\left[
\exp\left( i\hat{\theta}\right) \right] =0$. The probability distributions
for which $E\left[ \exp\left( i\hat{\theta}\right) \right] \neq0$, are
those for which the characteristic function of the probability distribution of
$\theta$ is not even. The variance in the phase difference between the two
components of the wave function is:
\begin{equation}
Var\left[ \Delta\hat{S}\right] =\frac{e\Phi}{\hbar c}\left[ Var\left(
\cos\hat{\theta}\right) +iVar\left( \sin\hat{\theta}\right) \right] .
\end{equation}
Thus, for zero mean Gaussian noise, the variance in the phase difference
between the two components of the wave function is $Var\left[ \Delta\hat
{S}\right] =\frac{e\Phi}{\hbar c}\left[ \left( 1-e^{-2\sigma_{\theta}^{2
}\right) \left( 1+i\right) \right] $ while for the Cauchy distribution, it
is $\frac{e\Phi}{\hbar c}\left[ \left[ 1-e^{-2\alpha}\right] \left(
1+i\right) \right] $. Higher order corrections can be found as well using
the material in the appendix. The same argument would also hold for both the
Aharonov-Casher effect and the gravitational equivalent to the AB-effect since
the phase difference between two paths for the quantum phenomena "$\circ$" can
be written as:
\begin{equation}
S_{1}-S_{2}=\frac{\alpha\Phi_{"\circ"}}{\hbar},
\end{equation}
where $\alpha$ is a phenomenological constant appropriate to the particular
situation and $\Phi_{"\circ"}$ is the flux associated with the phenomena
"$\circ$". An effective random variation in the gauge path for the phenomena
"$\circ$" is equal to
\begin{equation}
\Delta\hat{S}=\frac{\alpha\Phi_{"\circ"}e^{i\hat{\theta}_{"\circ"}}}{\hbar}.
\end{equation}
A third idea is couched in general relativity and related the space-time
metric. In general, the spatial metric of space-time can be expressed in terms
of direction cosines, for example in three dimensional space the
parameterization is:
\begin{align}
\hat{x} & =r\cos\left( \hat{\theta}\right) \sin\left( \hat{\varphi
}\right) ,\\
\hat{y} & =r\sin\left( \hat{\theta}\right) \sin\left( \hat{\varphi
}\right) ,\\
\hat{z} & =r\cos\left( \hat{\varphi}\right) .
\end{align}
Now,
\[
s^{2}=x^{2}+y^{2}+z^{2}=\hat{x}^{2}+\hat{y}^{2}+\hat{z}^{2}=\hat{s}^{2
\]
so the distance is invariant with respect to a replace of the angles by random
variables, so:
\begin{equation}
s\rightarrow\hat{s}\rightarrow se^{i\sigma\left( \theta,\varphi\right) }.
\end{equation}
where $\sigma$ is an arbitrary function of $\theta$ and $\varphi$, so
$\sigma=\sigma\left( \theta,\varphi\right) $. Therefore, it is possible to
associate a fluctuation of the space-time metric random variations of in the
angular variables. These random variations are equivalent to a random gauge,
which can be observed using the idea of a path length phase difference for any
quantum phenomena. Any phenomena which has an invariant which it is possible
to replace $s\rightarrow\hat{s}$ as in Eq (27) potentially has a hidden random
gauge associated with it, so the formula for a gauge path for the phenomena
"$\circ$" can be used to find it
\begin{equation}
\Delta\hat{S}=\frac{\alpha\Phi_{"\circ"}e^{i\hat{\theta}_{"\circ"}}}{\hbar}.
\end{equation}
A Mach-Zenhender interferometer, combined with weak amplification device such
as has been discussed by Aharonov\cite{Aharonov2005} , might be used to detect
such random variations in the metric. In a latter publication, the random
gauge concept will be used to reexamine what Aharonov has termed modular momentum.
\section{Discussion and Conclusions}
Two concepts of random gauge invariance have been introduced in this paper
with separate examples of what random gauge could mean in a physical setting.
Random phase probability density functions are explained in the appendix with
examples of a variety of probability density functions. One concept of random
phase is associated with the Feynman's path integrals and used to provide
another interpretation of the Feynman path integral. Another interpretation of
random gauge is introduced using the Aharonov-Bohm effect and a method is
proposed for detecting the randomness is proposed. Additionally, a gauge
proposed which provides an explanation for fluctuations in the space-time
metric and a method is proposed for detecting it.
\textbf{Acknowledgement:} Thanks to Joshua Bellamy for reading this document
and comments that significantly improved the document.
|
\section*{Introduction}
The systematic development of the theory of Riemannian almost
product manifolds, \ie Riemannian manifolds with almost product
structure, was started by K. Yano in \cite{Ya}. The geometry of a
Riemannian almost product manifold $(M,P,g)$ is a geometry of the
Riemannian metric $g$ and the almost product structure $P$. There
are important in this geometry the linear connections with respect
to which the parallel transport determines an isomorphism of the
tangent spaces of the manifold $M$ with structures $g$ and $P$.
Such connections are called natural in \cite{Mi}. They are an
analogue of the Hetmitian connections in almost Hermitian
geometry.
If the almost product structure $P$ is integrable, \ie the torsion
tensor of $P$ (the Nijenhuis tensor) is zero, then the manifold
$(M,P,g)$ is called a Riemannian product manifold.
In the present work we study a natural connection $D$ on a manifold $(M,P,g)$
belonging to the largest class of Riemannian product manifolds,
which is closed with respect to the group of the conformal transformations
of the metric $g$. This is the class $\W_1$ of the classification in \cite{Sta-Gri}.
The present paper is organized as follows.
In Section~\ref{sec1} we give some necessary facts about the
Riemannian almost product manifolds and the natural connections on
them.
In Section~\ref{sec2} we consider a natural connection $D$ from
the 2-parametric family of all natural connections on
$(M,P,g)\in\W_1$ obtained in \cite{Dobr}. We find a relation
between the curvature tensors of the Levi-Civita connection $\n$
and the considered connection $D$. As a corollary we obtain a
relation between the Ricci tensors and between the scalar
curvatures for $\n$ and $D$.
In Section~\ref{sec3} we find necessary and sufficient conditions
for a connection $D$ whose curvature tensor is a Riemannian
$P$-tensor. The notion of a Riemannian $P$-tensor is an analogue
of the notion of a K\"ahler tensor in Hermitian geometry.
In Section~\ref{sec4} we consider the case of a connection $D$
with parallel torsion.
In Section~\ref{sec5} we establish that the Weyl tensors for $\n$
and $D$ coincide.
In Section~\ref{sec6} we prove that the curvature tensor of $D$ is
invariant with respect to the usual conformal transformation of
the metric $g$.
In Section~\ref{sec7} we consider the case of flat connection $D$.
In Section~\ref{sec8} we construct an example of the considered
manifold by a Lie group, where $D$ is a flat connection with
non-parallel torsion.
\section{Preliminaries}\label{sec1}
Let $(M,P,g)$ be a \emph{Riemannian almost product manifold},
\ie{} a differentiable manifold $M$ with a tensor field $P$ of
type $(1,1)$ and a Riemannian metric $g$ such that
\begin{equation*}\label{2.1}
P^2x=x,\quad g(Px,Py)=g(x,y)
\end{equation*}
for arbitrary $x$, $y$ of the algebra $\X(M)$ of the smooth vector
fields on $M$. Obviously $g(Px,y)=g(x,Py)$.
Further $x,y,z,w$ will stand for arbitrary elements of $\X(M)$ or
vectors in the tangent space $T_pM$ at $p\in M$.
In this work we consider Riemannian almost product manifolds with
$\tr{P}=0$. In this case $(M,P,g)$ is an even-dimensional
manifold. Let $\dim{M}$ be $2n$. Then the \emph{associated metric}
$\tilde{g}$ of $g$, determined by $\tilde{g}(x,y)=g(x,Py)$, is an
indefinite metric of signature $(n,n)$. We suppose that $\dim
M\geq 4$.
In \cite{Nav} A.M.~Naveira gives a classification of Riemannian
almost product manifolds with respect to the tensor $F$ of type
(0,3), defined by
\begin{equation}\label{2.2}
F(x,y,z)=g\left(\left(\nabla_x P\right)y,z\right),
\end{equation}
where $\n$ is the Levi-Civita connection of $g$. The tensor $F$
has the following properties:
\begin{equation*}\label{2.3}
\begin{array}{c}
F(x,y,z)=F(x,z,y)=-F(x,Py,Pz),\\[4pt]
F(x,y,Pz)=-F(x,Py,z).
\end{array}
\end{equation*}
Using the Naveira classification, in \cite{Sta-Gri} M.~Staikova
and K.~Gribachev give a classification of Riemannian almost
product manifolds $(M,P,g)$ with $\tr P=0$. The basic classes of
the classification in \cite{Sta-Gri} are $\W_1$, $\W_2$ and
$\W_3$. Their intersection is the class $\W_0$ of the
\emph{Riemannian $P$-manifolds}, determined by the condition $F=0$
or equivalently $\n P=0$ \cite{Sta87}. This class is the analogue
of the class of K\"ahler manifolds in the geometry of almost
Hermitian manifolds.
A Riemannian almost product manifold $(M,P,g)$ is a
\emph{Riemannian product manifold} if it has a local product
structure. This means that the almost product structure $P$ is
integrable, \ie the Nijenhuis tensor $N$ determined by
\begin{equation*}\label{2.3'}
N(x,y)=[Px,Py]+[x,y]-P[Px,y]-P[x,Py]
\end{equation*}
is zero. The Riemannian product manifolds form the class
$\W_1\oplus\W_2$ from the classification in \cite{Sta-Gri}. This
class is an analogue of Hermitian manifolds in almost Hermitian
geometry.
The class $\W_1$ from the classification in \cite{Sta-Gri} consist
of the manifolds which are locally conformal equivalent to
Riemannian $P$-manifolds. This class plays a similar role of the
role of the class of the conformal K\"ahler manifolds in almost
Hermitian geometry \cite{Gray-Her}. The characteristic condition
for the class $\W_1$ is the following
\begin{equation}\label{2.4}
\begin{split}
& F(x,y,z)=\frac{1}{2n}\big\{ g(x,y)\ta (z)+g(x,z)\ta (y)-g(x,Py)\ta (Pz) \big.\\
& \phantom{F(x,y,z)=\frac{1}{2n}\big\{g(x,y)\ta (z)+g(x,z)\ta (y)
\big.} -g(x,Pz)\ta (Py)\big\},
\end{split}
\end{equation}
where $\ta$ is the associated Lee 1-form for $F$ determined by
\begin{equation}\label{2.4'}
\ta(x)=g^{ij}F(e_i,e_j,x).
\end{equation}
Here and further $g^{ij}$ will stand for the components of the
inverse matrix of $g$ with respect to a basis $\{e_i\}$ of $T_pM$
at $p\in M$.
The curvature tensor $R$ of $\n$ is determined by
$R(x,y)z=\nabla_x \nabla_y z - \nabla_y \nabla_x z -
\nabla_{[x,y]}z$ and the corresponding tensor of type (0,4) is defined as
follows $R(x,y,z,w)=g(R(x,y)z,w)$. We denote the Ricci tensor and
the scalar curvature for $\n$ by $\rho$ and $\tau$, respectively,
\ie $\rho(y,z)=g^{ij}R(e_i,y,z,e_j)$ and
$\tau=g^{ij}\rho(e_i,e_j)$.
In \cite{Mek1}, a tensor $L$ of type (0,4) with pro\-per\-tie
\begin{equation}\label{2.5}
L(x,y,z,w)=-L(y,x,z,w)=-L(x,y,w,z),
\end{equation}
\begin{equation}\label{2.6}
\mathop{\s} \limits_{x,y,z} L(x,y,z,w)=0,
\end{equation}
\begin{equation}\label{2.7}
L(x,y,Pz,Pw)=L(x,y,z,w),
\end{equation}
is called a \emph{Riemannian $P$-tensor}. This notion is an
analogue of the notion of a K\"ahler tensor in Hermitian geometry.
The linear connections in our investigations have a torsion.
Let $\nn$ be a linear connection with a tensor $Q$ of the
transformation $\n \rightarrow\nn$ and a torsion $T$, \ie{}
\begin{equation}\label{2.8}
\nn_x y=\n_x y+Q(x,y),\quad T(x,y)=\nn_x y-\nn_y x-[x,y].
\end{equation}
The corresponding (0,3)-tensors are defined by
\begin{equation}\label{2.9}
Q(x,y,z)=g(Q(x,y),z), \quad T(x,y,z)=g(T(x,y),z).
\end{equation}
The symmetry of the Levi-Civita connection implies
\begin{equation}\label{2.10}
T(x,y)=Q(x,y)-Q(y,x),
\end{equation}
\begin{equation}\label{2.11}
T(x,y)=-T(y,x).
\end{equation}
\begin{defn}[\cite{Mi}]\label{defn-2.1}
A linear connection $\nn$ on a Riemannian almost product manifold
$(M,P,g)$ is called a \emph{natural connection} if $\nn P=\nn
g=0$.
\end{defn}
If $\nn$ is a linear connection with a tensor $Q$ of the
transformation $\n \rightarrow\nn$ on a Riemannian almost product
manifold, then it is a natural connection if and only if the
following conditions are valid \cite{Mi}:
\begin{equation*}\label{2.13}
F(x,y,z)=Q(x,y,Pz)-Q(x,Py,z),
\end{equation*}
\begin{equation}\label{2.14}
Q(x,y,z)=-Q(x,z,y).
\end{equation}
Let $R'$ be the curvature tensor of a natural connection $\nn$ on
Riemannian almost product manifold $(M,P,g)$. Then, according to
the definitional equalities for $R$ and $R'$, bearing in mind $\nn
g=0$ and equalities \eqref{2.8}, \eqref{2.9}, \eqref{2.10},
\eqref{2.11} and \eqref{2.14}, we have the following relation for
$R$ and $R'$:
\begin{equation}\label{2.15}
\begin{split}
R(x,y,z,w)&=R'(x,y,z,w)-Q(T(x,y),z,w)\\[4pt]
&- \left(\nn_x Q\right)(y,z,w)+\left(\n_yQ\right)(x,z,w)\\[4pt]
&+g\bigl(Q(x,z),Q(y,w)\bigr)-g\bigl(Q(y,z),Q(x,w)\bigr).
\end{split}
\end{equation}
\section{A natural connection $D$ on $(M,P,g)\in\W_1$}\label{sec2}
In the classification of Riemannian almost product manifolds given
in \cite{Sta-Gri}, the class $\W_1$ is the only class, where the
basic tensor $F$ is expressed in terms of the tensor
$g\otimes\ta$, namely \eqref{2.4}.
Let $T$ be the torsion of an arbitrary natural connection on a
Riemannian product manifold $(M,P,g)\in\W_1$. In \cite{Dobr}, it
is obtained the following expression of $T$ by $g\otimes\ta$:
\begin{equation}\label{3.1}
\begin{split}
&T(x,y,z)=\frac{1}{2n}\left\{g(y,z)\ta(Px)-g(x,z)\ta(Py)\right\}\\[4pt]
&+\lm\left\{g(y,z)\ta(x)-g(x,z)\ta(y)+g(y,Pz)\ta(Px)-g(x,Pz)\ta(Py)\right\}\\[4pt]
&+\mu\left\{g(y,Pz)\ta(x)-g(x,Pz)\ta(y)+g(y,z)\ta(Px)-g(x,z)\ta(Py)\right\},
\end{split}
\end{equation}
where $\lm,\mu\in\R$. Let us note that for $\lm=0$ and
$\mu=-\frac{1}{4n}$ we have the torsion of the canonical
connection investigated in \cite{Sta-Gri}. The canonical
connection on an arbitrary Riemannian almost product manifold is
introduced in \cite{Mi} as an analogue of the Hermitian connection
in Hermitian geometry (\cite{Li-2}, \cite{Li-1}, \cite{Ya}).
The goal of the present work is the investigation of the natural
connection $D$ whose torsion $T$ is determined by \eqref{3.1} for
$\lm=\mu=0$, \ie
\begin{equation}\label{3.2}
T(x,y,z)=\frac{1}{2n}\left\{g(y,z)\ta(Px)-g(x,z)\ta(Py)\right\}.
\end{equation}
\begin{prop}\label{prop-3.1}
The connection $D$ is determined by the following equality
\begin{equation}\label{3.3}
g\left(D_xy,z\right)=g\left(\n_xy,z\right)+Q(x,y,z),
\end{equation}
where
\begin{equation}\label{3.4}
Q(x,y,z)=\frac{1}{2n}\left\{g(x,y)\ta(Pz)-g(x,z)\ta(Py)\right\}.
\end{equation}
\end{prop}
\begin{proof}
Since $D$ is a natural connection, according to \cite{Hay}, there
is valid the following equality for the tensor $Q$ of the
transformation $\n\rightarrow D$:
\begin{equation}\label{3.4'}
Q(x,y,z)=\frac{1}{2n}\left\{T(x,y,z)-T(y,z,x)+T(z,x,y)\right\}.
\end{equation}
Applying \eqref{3.2} to \eqref{3.4'}, we have \eqref{3.3}.
\end{proof}
For the corresponding tensor $Q$ of type (1,2), according to
\eqref{3.4}, we obtain
\begin{equation}\label{3.5}
Q(x,y)=\frac{1}{2n}\left\{g(x,y)P\Omega-\ta(Py)x\right\},
\end{equation}
where $\Omega$ is determined by $g(\Omega,x)=\ta(x)$. Then
\eqref{3.3} implies
\begin{equation}\label{3.6}
D_xy=\n_xy+\frac{1}{2n}\left\{g(x,y)P\Omega-\ta(Py)x\right\},
\end{equation}
which yields the following relation between the covariant
derivatives of $\ta$ with respect $D$ and $\n$:
\begin{equation}\label{3.7}
\left(D_x\ta\right)y=\left(\n_x\ta\right)y
-\frac{1}{2n}\left\{g(x,y)\ta(P\Omega)-\ta(Py)\ta(x)\right\}.
\end{equation}
We have the following immediately consequence of \eqref{3.5}:
\begin{equation}\label{3.8}
T(x,y)=\frac{1}{2n}\left\{\ta(Px)y-\ta(Py)\ta(x)\right\}.
\end{equation}
The equality \eqref{3.2} implies the properties
\begin{equation}\label{3.9}
\mathop{\s}_{x,y,z} T(x,y,z)=\mathop{\s}_{x,y,z}T(Px,Py,z)=0,
\end{equation}
where $\mathop{\s}_{x,y,z}$ is the cyclic sum by $x, y, z$.
Equalities \eqref{3.9} and \eqref{3.4'} yield
\begin{equation*}\label{3.9'}
Q(x,y,z)=T(z,y,x).
\end{equation*}
According to \eqref{3.8}, it is follows $\ta(PT(x,y))=0$ and then,
bearing in mind \eqref{3.2}, we obtain
\begin{equation*
T(T(x,y),z)=\frac{1}{4n^2}\left\{\ta(Py)\ta(Pz)x-\ta(Px)\ta(Pz)y\right\}.
\end{equation*}
Therefore we have the property
\begin{equation}\label{3.10}
\mathop{\s}_{x,y,z} T(T(x,y),z)=0.
\end{equation}
\begin{rem}
According to property \eqref{3.10}, the case $T(x,y)=-[x,y]$, \ie
$D_xy=D_yx$, is possible (\cite{Hel}, \cite{KamTon}). Then the
manifold has a structure of a Lie group.
\end{rem}
According to \eqref{2.15}, \eqref{3.4}, \eqref{3.5} and
\eqref{3.8}, we obtain the following identity for the curvature
tensors $R$ and $R'$ of the connections $\n$ and $D$:
\begin{equation}\label{3.11}
\begin{split}
R(x,y,z,w)&=R'(x,y,z,w)-\frac{\ta(\Omega)}{4n^2}\pi_1(x,y,z,w)\\[4pt]
&-\frac{1}{2n}\left\{g(y,z)\left(D_x\ta\right)Pw-g(x,z)\left(D_y\ta\right)Pw\right.\\[4pt]
&\phantom{-\frac{1}{2n}\left\{\right.}
\left.+g(x,w)\left(D_y\ta\right)Pz-g(y,w)\left(D_x\ta\right)Pz\right\},
\end{split}
\end{equation}
where
\begin{equation}\label{3.12}
\pi_1(x,y,z,w)=g(y,z)g(x,w)-g(x,z)g(y,w).
\end{equation}
Since $\mathop{\s}_{x,y,z} R(x,y,z,w)=0$ and $\mathop{\s}_{x,y,z}
\pi_1(x,y,z,w)=0$, then equality \eqref{3.11} implies
\begin{equation}\label{3.12'}
\mathop{\s}_{x,y,z} R'(x,y,z,w)=\frac{1}{2n}
\mathop{\s}_{x,y,z}
g(x,w)\left\{\left(D_y\ta\right)Pz-\left(D_z\ta\right)Py\right\}.
\end{equation}
In \cite{Sta-Gri}, it is introduced the following curvature-like
tensor $\psi_1$ on a Riemannian product manifold
\begin{equation}\label{3.13}
\begin{split}
\psi_1(S)(x,y,z,w)&=g(y,z)S(x,w)-g(x,z)S(y,w)\\[4pt]
&+S(y,z)g(x,w)-S(x,z)g(y,w),
\end{split}
\end{equation}
where
\begin{equation}\label{3.14}
S(x,y)=\left(D_x\ta\right)Py+\frac{\ta(\Omega)}{2n}g(x,y).
\end{equation}
Using \eqref{3.11}, \eqref{3.12}, \eqref{3.13} and \eqref{3.14},
we obtain the following
\begin{thm}\label{thm-3.2}
The curvature tensors $R$ and $R'$ of $\n$ and $D$ are related via
the formula
\begin{equation}\label{3.15}
R(x,y,z,w)=R'(x,y,z,w)-\frac{1}{2n}\psi_1(S)(x,y,z,w).
\end{equation}
\qed
\end{thm}
\begin{cor}\label{cor-3.3}
The Ricci tensors $\rho$ and $\rho'$ as well as the scalar
curvatures $\tau$ and $\tau'$ for $\n$ and $D$, respectively, are
related as follows
\begin{gather}
\rho(y,z)=\rho'(y,z)-\frac{1}{2n}\left\{g(y,z)\tr S+2(n-1)S(y,z)\right\},\label{3.16}\\[4pt]
\tau=\tau'-\frac{2n-1}{n}\tr S.\label{3.17}
\end{gather}
\qed
\end{cor}
\section{Connection $D$ with Riemannian $P$-tensor of curvature}\label{sec3}
Conditions \eqref{2.5} are satisfies for the curvature tensor $R'$
of the connection $D$ because of \eqref{3.11}. Moreover, condition
\eqref{2.7} is also valid for $R'$, because of $DP=0$. Therefore,
$R'$ is a Riemannian $P$-tensor if and only if $R'$ satisfies
condition \eqref{2.6}.
\begin{thm}\label{thm-4.1}
The connection $D$ has a Riemannian $P$-tensor of curvature if and
only if the following condition is valid
\begin{equation}\label{4.3}
\left(D_y\ta\right)Pz=\left(D_z\ta\right)Py.
\end{equation}
\end{thm}
\begin{proof}
Let the curvature tensor $R'$ of the connection $D$ be a
Riemannian $P$-tensor. Then \eqref{2.6} is valid for $R'$ and
because of \eqref{3.12'} we have
\begin{equation}\label{4.2}
\mathop{\s}_{x,y,z}
g(x,w)\left\{\left(D_y\ta\right)Pz-\left(D_z\ta\right)Py\right\}=0.
\end{equation}
Since we have supposed that $\dim M\geq 4$, then \eqref{4.2}
implies \eqref{4.3}.
Vice versa, if \eqref{4.3} is satisfied, then \eqref{3.12'} yields
condition \eqref{2.6} for $R'$ and therefore $R'$ is a Riemannian
$P$-tensor.
\end{proof}
Bearing in mind \eqref{4.3} and \eqref{3.7}, we have immediately
the following
\begin{cor}\label{cor-4.2}
The connection $D$ has a Riemannian $P$-tensor of curvature if and
only if the following condition is valid
\begin{equation*}
\left(\n_x\ta\right)Py=\left(\n_y\ta\right)Px,
\end{equation*}
\ie if and only if the 1-form $\ta\circ P$ is closed. \qed
\end{cor}
\section{Connection $D$ with parallel torsion}\label{sec4}
Equality \eqref{3.8} implies
\begin{equation*}
\left(D_xT\right)(y,z)=\frac{1}{2n}\left\{\left(D_x\ta\right)Py.z-\left(D_x\ta\right)Pz.y\right\},
\end{equation*}
which yields immediately the following
\begin{prop}\label{prop-5.1}
The connection $D$ has parallel torsion if and only if the
associated Lee 1-form $\ta$ is also parallel with respect to $D$.
\qed
\end{prop}
Because of \eqref{3.7}, it is valid the following
\begin{cor}\label{cor-5.2}
The connection $D$ has parallel torsion if and only if the
following condition is satisfied
\begin{equation*}
\left(\n_x\ta\right)y=\frac{1}{2n}\left\{g(x,y)\ta(P\Omega)-\ta(Py)\ta(x)\right\}.
\end{equation*}
\qed
\end{cor}
Let $D$ have a parallel torsion. Then, according to \eqref{3.11}
we obtain
\begin{equation}\label{5.1}
R(x,y,z,w)=R'(x,y,z,w)-\frac{\ta(\Omega}{4n^2}\pi_1(x,y,z,w),
\end{equation}
which implies condition \eqref{2.6} for $R'$. Therefore, it is
valid the following
\begin{prop}\label{prop-5.3}
If the connection $D$ has parallel torsion then the curvature
tensor of $D$ is a Riemannian $P$-tensor which satisfies condition
\eqref{5.1}. \qed
\end{prop}
\section{The Weyl tensor of the transformation $\n\rightarrow D$}\label{sec5}
Let $W$ and $W'$ be the Weyl tensors for the connections $\n$ and
$D$, respectively, \ie
\begin{equation}\label{6.1}
\begin{split}
&W=R-\frac{1}{2(n-1)}\left\{\psi_1(\rho)-\frac{\tau}{2n-1}\pi_1\right\},\\[4pt]
&W'=R'-\frac{1}{2(n-1)}\left\{\psi_1(\rho')-\frac{\tau'}{2n-1}\pi_1\right\}.
\end{split}
\end{equation}
\begin{thm}\label{thm-6.1}
The Weyl tensor is invariant with respect to the transformation
$\n\rightarrow D$.
\end{thm}
\begin{proof}
We determine $\tr S$ from \eqref{3.17} and replace it in
\eqref{3.16}. So we obtain
\begin{equation*}
\frac{n-1}{n}S=\left\{\rho'-\frac{\tau'}{2(2n-1)}g\right\}-\left\{\rho-\frac{\tau}{2(2n-1)}g\right\}.
\end{equation*}
After that we determine $S$ from the latter equality and replace
it in \eqref{3.15}. In the result we take into account
\eqref{3.13} and the equality $\psi_1(g)=2\pi_1$ and obtain $W=W'$
by appropriate regrouping.
\end{proof}
\section{Conformal transformation of the curvature tensor of $D$}\label{sec6}
Now we will establish the way of transforming of the curvature
tensor $R'$ of the connection $D$ using the usual conformal
transformation of the metric $g$. This transformation is
determined via the formula
\begin{equation}\label{7.1}
\bar{g}=e^{2u}g,
\end{equation}
where $u$ is a smooth function on the considered manifold.
It is known that the Levi-Civita connection is transformed by
\eqref{7.1} as follows
\begin{equation}\label{7.2}
\bar{\n}_xy=\n_xy+\dd u(x)y+\dd u(y)x-g(x,y)L,
\end{equation}
where $L=\grad u$.
In \cite{Sta-Gri}, it is proved that $(M,P,g)\in\W_1$ implies
$(M,P,\bar{g})\in\W_1$, such as the associated Lee 1-forms $\ta$
and $\bar{\ta}$ as well as their corresponding vectors $\Omega$
and $\bar{\Omega}$ are related via the formulas
\begin{gather}
\bar{\ta}(x)=\ta(x)+2u\dd u(Px),\label{7.3}\\[4pt]
\bar{\Omega}=e^{-2u}\left(\Omega+2nL\right).\label{7.4}
\end{gather}
Analogously to \eqref{3.6} it is valid the following
\begin{equation*}
\bar{D}_xy=\bar{\n}_xy+\frac{1}{2n}\left\{\bar{g}(x,y)P\bar{\Omega}-\bar{\ta}(Py)x\right\},
\end{equation*}
The latter equality and equalities \eqref{7.1}, \eqref{7.2},
\eqref{7.3} and \eqref{7.4} yield
\begin{equation}\label{7.5}
\bar{D}_xy=D_xy+\dd u(x)y.
\end{equation}
Using the definitional equality
$\bar{R}'(x,y)z=\bar{D}_x\bar{D}_yz-\bar{D}_y\bar{D}_xz-\bar{D}_{[x,y]}z$
for the curvature tensor $\bar{R}'$ of $\bar{D}$ and \eqref{7.5},
we obtain $\bar{R}'=R'$, \ie it is valid the following
\begin{thm}\label{thm-7.1}
The curvature tensor of the connection $D$ is invariant with
respect to the usual conformal transformation \eqref{7.1} of the
metric $g$. \qed
\end{thm}
\section{The case of a flat connection $D$}\label{sec7}
Let $D$ be a flat connection, \ie $R'=0$. Then $W'=0$ and because
of \thmref{thm-6.1} we have $W=0$. Therefore it is valid the
following
\begin{prop}\label{prop-8.1}
If $D$ is a flat connection then the manifold is conformally flat
with respect to $\n$. \qed
\end{prop}
Further, we have the following
\begin{thm}\label{thm-8.2}
Let the connection $D$ be a flat connection with parallel torsion.
Then the following propositions are valid:
\begin{enumerate} \renewcommand{\labelenumi}{(\roman{enumi})}
\item
$R=-\frac{1}{4n^2}\ta(\Omega)\pi_1$, $\quad\rho=-\frac{2n-1}{4n^2}\ta(\Omega)g$,
$\quad\tau=-\frac{2n-1}{2n}\ta(\Omega)$;
\item The tensor $R$ is parallel with respect to $D$;
\item The manifold is a space form;
\item The scalar curvature $\tau$ is negative.
\end{enumerate}
\end{thm}
\begin{proof}
Let the connection $D$ be a flat connection with parallel torsion.
Then, according to the \propref{prop-5.3}, we obtain the first
equality in (i), which implies the other two equalities of (i).
Bearing in mind \eqref{3.3}, we have
\begin{equation}\label{8.2}
\begin{split}
&R(x,y)z=R'(x,y)z -\left(D_x Q\right)(y,z)+\left(D_y Q\right)(x,z)
\\[4pt]
&\phantom{R(x,y)z=R'(x,y)z}
+Q\bigl(x,Q(y,z)\bigr)-Q\bigl(y,Q(x,z)\bigr).
\end{split}
\end{equation}
Since $DT=0$, using \eqref{3.4'}, we obtain $DQ=0$. Then relation
\eqref{8.2} implies $DR=DR'$. Hence, because of $DR'=0$, we obtain
$DR=0$, \ie (ii) is valid.
The formulas in (i) imply directly (iii).
Since $g$ is a Riemannian metric, then
$\ta(\Omega)=g(\Omega,\Omega)> 0$. Therefore, because of the
latter equality in (i), we have that $\tau<0$, \ie (iv) is valid.
\end{proof}
\section{Example}\label{sec8}
\subsection{A Lie group $G$ as a Riemannian product $\W_1$-manifold $(G,P,g)$}
Let $G$ be a 4-dimensional real connected Lie group and $\g$ be
its Lie algebra with a basis $\{X_i\}$.
We introduce a structure $P$ and left invariant metric $g$ as
follows
\begin{equation}\label{9.1}
PX_1=X_3,\quad PX_2=X_4,\quad PX_3=X_{1},\quad PX_4=X_{2},
\end{equation}
\begin{equation}\label{9.2}
g(X_i,X_j)=
\begin{cases}
\begin{array}{rl}
1, \quad & i=j;\\
0, \quad & i\neq j.
\end{array}
\end{cases}
\end{equation}
Thus, $(G,P,g)$ becomes a Riemannian almost product manifold with
$\tr{P}=0$.
We will consider the case when $P$ is an Abelian almost product
structure \cite{AnBaDoOv}, \ie
\begin{equation}\label{9.3}
[PX_i,PX_j]=-[X_i,X_j].
\end{equation}
Then the manifold $(G,P,g)$ has a zero Nijenhuis tensor and
therefore $(G,P,g)$ belongs to the class $\W_1\oplus\W_2$.
\begin{thm}\label{thm-9.1}
The manifold $(G,P,g)$ is a Riemannian product manifold belonging
to the class $\W_1$ if and only if the Lie algebra $\g$ is
determined by the conditions:
\begin{equation}\label{9.5}
\begin{array}{l}
[X_1,X_2]=-[X_3,X_4]=\lm_1 X_1+\lm_2 X_2+\lm_3 X_3+\lm_{4} X_4,\\[4pt]
[X_1,X_3]=[X_2,X_4]=\lm_4 X_1-\lm_3 X_2+\lm_2 X_3-\lm_1 X_4,\\[4pt]
[X_2,X_3]=[X_1,X_4]=0,\qquad (\lm_i\in \R;\; i=1,2,3,4).
\end{array}
\end{equation}
\end{thm}
\begin{proof}
Because of \eqref{9.2} we have
\begin{equation}\label{9.3'}
\begin{split}
2g\bigl(\n_{X_i} X_j,X_k\bigr)
=
g\left([X_i,X_j],X_k\right)&+g\left([X_k,X_i],X_j\right)\\[4pt]
&+g\left([X_k,X_j],X_i\right),
\end{split}
\end{equation}
which implies the following equality, according to \eqref{9.3} and
\eqref{2.2}:
\begin{equation}\label{9.4}
\begin{split}
2F\left(X_i,X_j,X_k\right)&=g\bigl([X_i,PX_j]-P[X_i,X_j],X_k\bigr)\\[4pt]
&+g\bigl([X_i,PX_k]-P[X_i,X_k],X_j\bigr)\\[4pt]
&+2g\bigl([X_k,PX_j],X_i\bigr).
\end{split}
\end{equation}
The comparing of condition \eqref{9.4} with the characteristic
condition \eqref{2.4} for the class $\W_1$, taking into account
\eqref{9.1}, \eqref{9.2}, \eqref{9.3} and the Jacobi identity for
the commutators $[X_i,X_j]$, yields conditions \eqref{9.5}.
\end{proof}
The comparing of \eqref{9.4} and \eqref{2.4} give also the
following formulas for the associated Lee 1-form $\ta$:
\begin{equation}\label{9.6}
\ta_1=4\lm_4,\quad \ta_2=-4\lm_3,\quad \ta_3=-4\lm_2,\quad
\ta_4=4\lm_1.
\end{equation}
Further, $(G,P,g)$ will stand for the manifold determined by
\eqref{9.5}.
\subsection{Some geometrical characteristics of the manifold $(G,P,g)$}
By virtue of \eqref{9.3'} and \eqref{9.5} we get the non-zero
components $\n_{X_i}X_j$ of the Levi-Civita connection $\n$:
\begin{equation}\label{9.7}
\begin{array}{l}
\n_{X_1} X_1=\n_{X_4} X_4=-\lm_1 X_2-\lm_4 X_3,\\[4pt
\n_{X_2} X_2=\n_{X_3} X_3=\lm_2 X_1+\lm_3 X_4,\\[4pt
\n_{X_1} X_2=-\n_{X_3} X_4=\lm_1 X_1+\lm_3 X_3,\\[4pt
\n_{X_2} X_1=-\n_{X_4} X_3=-\lm_2 X_2-\lm_4 X_4,\\[4pt
\n_{X_1} X_3=\n_{X_2} X_4=\lm_4 X_1-\lm_3 X_2,\\[4pt
\n_{X_3} X_1=\n_{X_4} X_2=-\lm_2 X_3+\lm_1 X_4.
\end{array}
\end{equation}
Using \eqref{9.7} and \eqref{9.5} we obtain the non-zero
components $R_{ijks}=R(X_i,\allowbreak X_j,X_k,\allowbreak{}X_s)$
of the curvature tensor $R$ for $\n$:
\begin{equation}\label{9.8}
\begin{array}{c}
\begin{array}{ll}
R_{1212}=\lm_1^2+\lm_2^2,\quad & R_{1313}=\lm_2^2+\lm_4^2,\\[4pt]
R_{1414}=\lm_1^2+\lm_4^2,\quad & R_{2323}=\lm_2^2+\lm_3^2,\\[4pt]
R_{2424}=\lm_1^2+\lm_3^2,\quad & R_{3434}=\lm_3^2+\lm_4^2,\\[4pt]
\end{array}\\[4pt]
\begin{array}{ll}
R_{1213}=R_{2134}=\lm_1\lm_4,\quad & R_{1214}=R_{2343}=\lm_2\lm_4,\\[4pt]
R_{1232}=R_{1434}=\lm_1\lm_3,\quad & R_{1242}=R_{1343}=\lm_2\lm_3,\\[4pt]
R_{1341}=R_{2342}=\lm_1\lm_2,\quad & R_{1332}=R_{1442}=\lm_3\lm_4.
\end{array}
\end{array}
\end{equation}
The rest of the non-zero components are obtained by the properties
\[
R_{ijks}=R_{ksij},\qquad R_{ijks}=-R_{jiks}=-R_{ijsk}.
\]
Using \eqref{9.8} for the non-zero components
$\rho_{ij}=\rho(X_i,X_j)$ of the Ricci tensor $\rho$ we compute:
\begin{equation}\label{9.9}
\begin{array}{c}
\begin{array}{ll}
\rho_{11}=-2\left(\lm_1^2+\lm_2^2+\lm_4^2\right),\quad
&
\rho_{22}=-2\left(\lm_1^2+\lm_2^2+\lm_3^2\right),\\[4pt]
\rho_{33}=-2\left(\lm_2^2+\lm_3^2+\lm_4^2\right),\quad
&
\rho_{44}=-2\left(\lm_1^2+\lm_3^2+\lm_4^2\right),\\[4pt]
\end{array}\\[4pt]
\begin{array}{lll}
\rho_{12}=2\lm_3\lm_4,\quad & \rho_{13}=-2\lm_1\lm_3,\quad & \rho_{14}=-2\lm_2\lm_3,\\[4pt]
\rho_{34}=2\lm_1\lm_2,\quad & \rho_{23}=-2\lm_1\lm_4,\quad & \rho_{24}=-2\lm_2\lm_4.
\end{array}
\end{array}
\end{equation}
The rest of the non-zero components are obtained by the property
$\rho_{ij}=\rho_{ji}$.
By \eqref{9.9} we obtain the following
\begin{prop}\label{prop-8.1'}
The manifold $(G,P,g)$ has a negative scalar curvature for $\n$:
\begin{equation}\label{9.10}
\tau=-6\left(\lm_1^2+\lm_2^2+\lm_3^2+\lm_4^2\right).
\end{equation}
\qed
\end{prop}
For the Riemannian sectional curvatures of the $P$-invariant basis
2-planes \allowbreak $(X_1,X_3)$ and $(X_2,X_4)$, i.e. for the
invariant sectional curvatures of the basis 2-planes, we get
\begin{equation}\label{9.11}
k_{13}=-\left(\lm_2^2+\lm_4^2\right),\qquad
k_{24}=-\left(\lm_1^2+\lm_3^2\right).
\end{equation}
The sectional curvatures of the rest of the basis 2-planes, i.e.
the anti-invariant sectional curvatures of the basis 2-planes,
are:
\begin{equation}\label{9.12}
\begin{array}{ll}
k_{12}=-\left(\lm_1^2+\lm_2^2\right),\qquad &
k_{14}=-\left(\lm_1^2+\lm_4^2\right),\\[4pt]
k_{23}=-\left(\lm_2^2+\lm_3^2\right),\qquad &
k_{34}=-\left(\lm_3^2+\lm_4^2\right).
\end{array}
\end{equation}
Conditions \eqref{9.11} and \eqref{9.12} imply the following
\begin{thm}\label{thm-9.2}
The manifold $(G,P,g)$ has:
\begin{enumerate}\renewcommand{\labelenumi}{(\roman{enumi})}
\item
a constant invariant sectional curvature if and only if
\[
\lm_1^2-\lm_2^2+\lm_3^2-\lm_4^2=0;
\]
\item
a constant anti-invariant sectional curvature if and only if
\[
\lm_1^2=\lm_3^2,\qquad \lm_2^2=\lm_4^2;
\]
\item
a constant sectional curvature if and only if
\[
\lm_1^2=\lm_2^2=\lm_3^2=\lm_4^2.
\]
In this case
$R=\frac{\tau}{12}\pi_1$. \qed
\end{enumerate}
\end{thm}
By virtue of \eqref{6.1}, using \eqref{9.8}, \eqref{9.9},
\eqref{9.10}, \eqref{3.12} and \eqref{3.13} for $S=\rho$, we
obtain the following
\begin{prop}\label{prop-8.2'}
The manifold $(G,P,g)$ has a zero Weyl tensor, \ie $(G,P,g)$ is
conformally flat for $\n$. \qed
\end{prop}
\subsection{The connection $D$ on the manifold $(G,P,g)$}
Further in our considerations we exclude the trivial case
$\lm_1=\lm_2=\lm_3=\lm_4=0$, \ie the case when $(G,P,g)$ is a
Riemannian $P$-manifold.
By virtue of \eqref{3.6}, \eqref{9.7} and \eqref{9.2} for the
components of the connection $D$ on $(G,P,g)$ we obtain:
\begin{equation}\label{9.13}
\begin{array}{ll}
D_{X_1} X_1=-\lm_3 X_4,\quad & D_{X_1} X_2=\lm_3 X_3,\\[4pt
D_{X_1} X_3=-\lm_3 X_2,\quad & D_{X_1} X_4=\lm_3 X_1,\\[4pt
D_{X_2} X_1=-\lm_4 X_4,\quad & D_{X_2} X_2=\lm_4 X_3,\\[4pt
D_{X_2} X_3=-\lm_4 X_2,\quad & D_{X_2} X_4=\lm_4 X_1,\\[4pt
D_{X_3} X_2=-\lm_1 X_3,\quad & D_{X_3} X_1=\lm_1 X_4,\\[4pt
D_{X_3} X_4=-\lm_1 X_1,\quad & D_{X_3} X_3=\lm_1 X_2,\\[4pt
D_{X_4} X_2=-\lm_2 X_3,\quad & D_{X_4} X_1=\lm_2 X_4,\\[4pt
D_{X_4} X_4=-\lm_2 X_1,\quad & D_{X_4} X_3=\lm_2 X_2.
\end{array}
\end{equation}
Using \eqref{9.13} and \eqref{9.5}, we obtain that all components
of the curvature tensor $R'$ of $D$ are zeros, \ie the following
proposition holds:
\begin{prop}\label{prop-8.2''}
The manifold $(G,P,g)$ has a flat connection $D$. \qed
\end{prop}
By virtue of \eqref{3.8} and \eqref{9.6} we get the non-zero
components $T_{ij}=T(X_i,X_j)$ of the torsion $T$ for the
connection $D$:
\begin{equation*}\label{9.14}
\begin{array}{ll}
T_{12}=-\lm_1 X_1-\lm_2 X_2,\quad & T_{13}=-\lm_4 X_1-\lm_2 X_3,\\[4pt
T_{14}=\lm_3 X_1-\lm_2 X_4,\quad & T_{23}=-\lm_4 X_2+\lm_1 X_3,\\[4pt
T_{24}=\lm_3 X_2+\lm_1 X_4,\quad & T_{34}=\lm_3 X_3+\lm_4 X_4.
\end{array}
\end{equation*}
The rest of the non-zero components are obtained by the property
$T_{ij}=-T_{ji}$.
According to \propref{prop-5.1}, the connection $D$ has a parallel
torsion if and only if $\left(D_{X_i}\ta\right)X_j=0$. The latter
equality is equivalent to $\lm_1=\lm_2=\lm_3=\lm_4=0$, because of
\eqref{9.6} and \eqref{9.13}. This case is excluded from our
investigations. Therefore, the following proposition holds:
\begin{prop}\label{prop-8.3}
The connection $D$ on the manifold $(G,P,g)$ is not parallel. \qed
\end{prop}
|
\section{Introduction}
\label{sec:intro}
The transition between plateaux in the Integer Quantum Hall Effect
(IQHE) is a quantum critical phenomenon, which was predicted
theoretically~\cite{LLP,Pruisken} and observed
experimentally~\cite{QHE-exp} a few decades ago. Although
experimentally there is no a priori reason to neglect
electron-electron interactions, it is usually modelled
theoretically by noninteracting particles in two dimensions (2d), in a
perpendicular magnetic field and a random potential. Despite the
apparent simplicity of this conceptual setup, it turns out to be
very difficult to derive analytically the critical exponents of
this transition. Important progress was achieved by the
introduction of a simple network model which retains the salient
features of guiding centre motion and quantum tunnelling in the
presence of disorder: the Chalker-Coddington (CC) model~\cite{CC}.
Extensive numerical studies based on the CC model
or other approaches have led to good estimates for the critical
exponents, notably the correlation-length exponent $\nu = 2.37 \pm 0.02$~\cite{Cain}
(a larger value $\nu=2.593\pm 0.006$ has also been reported~\cite{Slevin}).
Also, a semi-classical argument~\cite{QHE-semi} yields the prediction $\nu=7/3$.
The CC model is also the starting point for several analytical approaches, like
the description by a $\sigma$-model~\cite{Ludwig}, or
a mapping to a one-dimensional (1d) quantum many-body system~\cite{KM,MT}.
However, from the point of view of critical lattice models, no
exact solution of the CC model has been found so far.
The situation is very different for the {\it spin} Quantum Hall
Effect (SQHE): the generalisation of the CC model to
SQHE~\cite{CC-SU2} (which we shall call $\Sp(2)$-CC) maps exactly
to classical bond percolation, where a large class of exponents
are known~\cite{DS}. This mapping of $\Sp(2)$-CC to classical
percolation was first observed by Gruzberg {\it et al.}~\cite{Gruzberg}, who used
a supersymmetric (SUSY) spin-chain formulation. Later on, it was
realised~\cite{CC-path1,CC-path2} that the SUSY {\it lattice path integral}
maps $\Sp(2)$-CC to a statistical model of lattice paths, which
are exactly the hulls of bond-percolation clusters. Moreover, a
number of SQHE physical observables are expressed in terms of
percolation correlation functions, and this mapping is valid even
at the level of lattice models.
In this paper, we propose a treatment of the original CC model based
on the lattice path integral. Since the corresponding statistical model
involves paths which may pass through a given edge
infinitely many times, the number of configurations per unit surface
is infinite, and the model is not directly tractable by exact-solution
methods such as Yang-Baxter integrability and Conformal Field Theory (CFT).
We therefore introduce a {\it truncation procedure}, leading to a series of
finite statistical models, and focus on the first order of truncation.
The arising model is a two-colour loop model including vacancies, and with loop
fugacity $n=0$.
Integrable multi-colour loop models have been known for a long
time~\cite{multi-colour}. They were originally defined through
multi-dimensional height models, but they may as well describe
coupled copies of classical magnetism models, such as the Potts or
$\On$ models, and also the ground state of {\it quantum} loop
models. More specifically, in a two-colour, completely
packed ({\it i.e.} without vacancies) loop
model~\cite{Martins-Nienhuis,Paul-Jesper}, new integrable points
were identified through a mapping to a braid-monoid algebra: the Birman-Wenzl-Murakami
(BWM) algebra~\cite{BWM}. In the present paper, we use a similar approach on
the loop model arising from our truncation procedure, which is a
two-colour loop model including vacancies. Generalising to
arbitrary loop fugacity $n$, we obtain four critical branches in
the phase diagram of this loop model. We then study the critical
properties of these branches.
We find that two of these regimes (denoted 1 and 2) correspond to a pair
of decoupled Coulomb-Gas (CG) theories, whereas the other two (3 and 4)
relate to the $\SU(2)_r \times \SU(2)_r / \SU(2)_{2r}$ Wess-Zumino-Witten
coset model, for values $n= \pm 2 \cos \frac{\pi}{r+2}$ with $r\in\{1,2,3,\dots\}$.
We obtain analytically two critical exponents: one of them, $X_{\rm int}$, corresponds
to an elliptic deformation of the integrable weights, and the other one,
$X_{(1,1;{\rm adj})}$, is associated to a perturbation of the weight per monomer.
The truncated, modified CC model is realised by the $n=0$
point of regime 4, but this point is outside the validity range
for the analytic continuation of the WZW exponents.
Our numerical study gives the estimate $\nu \simeq 1.1$ for the correlation-length
exponent, and $d_f \simeq 1.71$ for the fractal dimension of paths.
This is clearly incompatible with the IQHE universality class, and hence our
integrable two-colour loop model is only a crude approximation to IQHE.
However, the truncation procedure may be carried out to higher orders,
possibly yielding more accurate, solvable approximations.
The plan of the paper is as follows. In Section 2, we recall
the definition of the CC model and its lattice SUSY path-integral formulation,
and explain our truncation procedure,resulting in a two-colour
loop model. This truncation is compared in detail with the one used in~\cite{KM,MT}.
In Section 3, we use a mapping to a dilute braid-monoid algebra to derive
the integrable Boltzmann weights of the two-colour loop model,
as well as the corresponding 1d Hamiltonian.
In Section 4, we identify the four critical regimes of the integrable
model and the corresponding CFTs. Numerical and analytical support for
the identification of these CFTs is given.
In Section 5, we examine in more detail regime 4, which contains the truncated,
modified CC model at $n=0$. We discuss the analytic continuation of CFT results,
and estimate numerically some critical exponents, including the correlation-length
exponent $\nu$.
The paper has three appendices. Appendix A contains the details of the mapping
to the dilute BWM (dBWM) algebra used in Section 3. In Appendix B, we exhibit a lattice
holomorphic parafermion $\psi_s(z)$ in the integrable model. In Appendix C, we expose
the exact solution of a particular point in regime 4, which is mapped to free fermions.
This mapping provides a valuable check on our results, and also gives a proof
that the ${\rm O}(n=1)$ loop model has central charge $c=\half$.
\section{Truncation of the Chalker-Coddington model}
\label{sec:CC}
\subsection{The Chalker-Coddington model}
\label{sec:intro-CC}
The Chalker-Coddington model~\cite{CC} is a simple lattice model for
the IQHE. The latter consists of a two-dimensional gas of non-interacting
electrons in a disordered medium, subject to a strong transverse magnetic
field. In the presence of the random potential, the Landau levels are
broadened, and eigenenergies are of the form $E=\left( k+\half \right) \hbar \omega_c + V_0$,
where $k$ is an integer, $\omega_c$ is the cyclotron energy
of the electron in the magnetic field, and $V_0$ is a random part.
Let us recall briefly the main ingredients of the CC model.
We consider an electron in the
eigenstate of energy $E$. The spatial trajectories
of the electron over finite time steps $\Delta t$ are modelled by paths on the
directed square lattice $\cal L$
(see Fig.~\ref{fig:cc-def}),
and the time-evolution operator over $\Delta t$ is denoted $\cal U$.
The operator $\cal U$ reads
\begin{equation}
{\cal U} = \bigotimes_{{\rm edge} \ e} U_e
\ \bigotimes_{{\rm vertex} \ v} U_v \,,
\end{equation}
with two types of factors:
\begin{itemize}
\item On each directed edge $e$, the operator $U_e$
takes the particle along $e$ and multiplies the wavefunction by a
random Aharonov-Bohm phase $\exp(i\phi_e)$, where the $\phi_e$ are
independent and uniformly distributed on the interval $[0,2\pi]$.
\item At each vertex $v$, the operator $U_v$ scatters the particle to one
of the outgoing edges.
In the bases $(1,2)$ and $(3,4)$ of Fig.~\ref{fig:cc-def}, $U_v$ is represented by the
unitary matrix:
\begin{equation}\label{eq:Smatrix}
S = \left( \begin{array}{cc}
\tanh \beta & 1/{\cosh \beta} \\
1/{\cosh \beta} & -\tanh \beta
\end{array} \right) \,.
\end{equation}
\end{itemize}
The parameter $\beta$ measures the distance to the plateau transition at
$E=E_c=\left(k+\half\right)\hbar \omega_c$. The critical value is
$\beta_c=\log(1+\sqrt{2})$, and the corresponding energy perturbation is assumed to behave
as~\cite{CC}
\begin{equation}
(E-E_c) \propto (\beta-\beta_c) \,.
\end{equation}
No exact solution of the CC model is known, in the sense that the critical
exponents have not been determined analytically. However, very
good numerical estimates exist for some of these
exponents~\cite{Huck,Cain,Slevin}. In particular, the correlation-length
exponent $\nu_{\rm CC}$, defined by the scaling of the correlation
length
\begin{equation} \label{eq:nu-CC}
\xi \propto |E-E_c|^{-\nu_{\rm CC}} \,,
\end{equation}
has been estimated as~\cite{Cain}
\begin{equation}
\nu_{\rm CC} \simeq 2.37 \pm 0.02 \,.
\end{equation}
\begin{figure}[th]
\begin{center}
\begin{tabular}{m{4cm}m{4cm}}
\includegraphics{cc-lattice.eps}
& \includegraphics{cc-vertex.eps} \\
(a) & (b)
\end{tabular}
\caption{
(a) Oriented square lattice $\cal L$ for the Chalker-Coddington model.
(b) Labelling of the edges adjacent to a vertex of $\cal L$.
}
\label{fig:cc-def}
\end{center}
\end{figure}
\subsection{Path integral representation}
\label{sec:path-int}
The problem of solving the CC model amounts to the diagonalisation
of a random time-evolution operator. We want to perform the
average over disorder, in order to turn this into a
translationally invariant 2d classical model. For
this purpose, we use the supersymmetric path integral
representation~\cite{Efetov}. The following derivation is very
analogous to what was done by one of us for the SQHE~\cite{CC-path2},
and we use the notations of~\cite{CC-path2} throughout this Section.
The Green's function between two edges $e_1$ and $e_2$ is:
\begin{equation}
G(e_2,e_1,z) := \langle e_2 | (1-z {\cal U})^{-1} | e_1 \rangle \,.
\end{equation}
Here $z$ is a parameter which plays the role of the energy in the
usual Green's function $(E-{\cal H})^{-1}$: roughly speaking
$z\sim e^{{\rm i}E}$, where $E$ is measured from the filled Landau
level. We label $e_L, e_R$ the ends of any edge $e$, with the
convention that it is directed in the sense $e_R \to e_L$, and we
introduce the complex variables $b_L(e), b_R(e)$. The Gaussian
measure is defined as:
\begin{equation}
\int [{\rm d} b] \ (\dots) :=
\frac{1}{\pi} \int {\rm d}({\rm Re} \ b) \ {\rm d}({\rm Im} \ b)
\ \exp(-b^*b) (\dots) \,,
\qquad [{\rm D}b]:=\prod_e [{\rm d}b_L(e)] [{\rm d}b_R(e)] \,.
\end{equation}
The Green's function can then be written as a Gaussian integral on the $b_L(e), b_R(e)$:
\begin{equation} \label{eq:G-boson}
G(e_2,e_1,z) = \frac{\int [{\rm D}b] \ b_L(e_2) b_L^*(e_1) \exp A_b}
{\int [{\rm D}b] \ \exp A_b} \,,
\end{equation}
where the action reads
\begin{eqnarray}
A_b &=& A_b^{({\rm edge})} + A_b^{({\rm vertex})} \,, \\
A_b^{({\rm edge})} &=& z \sum_{{\rm edge} \ e}
b^*_L(e) \exp(i \phi_e) b_R(e) \,, \label{eq:edge}\\
A_b^{({\rm vertex})} &=& \sum_{{\rm vertex} \ v}
\ \sum_{\begin{smallmatrix} i &\to&j \\ &v& \end{smallmatrix}}
b^*_R(e_i) S_{ij} b_L(e_j) \,,
\end{eqnarray}
and the notation $\begin{smallmatrix} & \phantom{v} & \\ i &\to&j \\ &v& \end{smallmatrix}$ means
that $i$ ({\resp} $j$) is an incoming
({\resp} outgoing) edge adjacent to $v$.
The next step is to express the denominator in~\eqref{eq:G-boson} as the inverse of
a Gaussian integral over Grassmann variables $f_{L,R}(e), \bar{f}_{L,R}(e)$:
\begin{equation} \label{eq:G-SUSY}
G(e_2,e_1,z) = \int [{\rm D}b] [{\rm D}f] \ b_L(e_2) b_L^*(e_1) \exp (A_b + A_f) \,,
\end{equation}
with the measure
\begin{equation}
\int [{\rm d} f] \ (\dots) :=
\int {\rm d}\bar{f} \ {\rm d}f \ \exp(-\bar{f} f) (\dots) \,,
\qquad {[{\rm D}f]} := \prod_e [{\rm d}f_L(e)][{\rm d}f_R(e)] \,,
\end{equation}
and $A_f$ is the analog of $A_b$, with $b,b^*$ replaced by $f,\bar{f}$.
We denote by an overbar the quenched average over the variables $\phi_e$. A useful formula
for this computation is
\begin{equation} \label{eq:quench}
\frac{1}{2\pi} \int_0^{2\pi} {\rm d}\phi \exp \left(
u e^{i\phi} + v^* e^{-i\phi}
\right) = \sum_{m=0}^\infty \frac{(u v^*)^m}{(m!)^2} \,.
\end{equation}
It easy to see, for instance, that $\overline{G(e_2,e_1,z)} = \delta(e_1,e_2)$.
When studying transport properties, the main quantity of interest is $\overline{|G|^2}$.
We write
\begin{eqnarray}
|G(e_2,e_1,z)|^2 &=&
\int [{\rm D}b] [{\rm D}f] \ b_L(e_2) b_L^*(e_1) \ e^{A_b + A_f}
\times \int [{\rm D}b] [{\rm D}f] \ b_L^*(e_2) b_L(e_1) \ e^{A_b^* + A_f^*} \nonumber \\
&=& \int [{\rm D}b_{1,2}] [{\rm D}f_{1,2}]
\ b_{L1}(e_2) b_{L1}^*(e_1) b_{L2}^*(e_2) b_{L2}(e_1)
\ e^{A_{b1} + A_{f1} + A_{b2}^* + A_{f2}^*} \,.
\end{eqnarray}
Using~\eqref{eq:quench}, we get:
\begin{eqnarray}
&& \overline{|G(e_2,e_1,z)|^2} =
\int [{\rm D}b_{1,2}] [{\rm D}f_{1,2}]
\ b_{L1}(e_2) b_{L1}^*(e_1) b_{L2}^*(e_2) b_{L2}(e_1) \times \nonumber \\
&& \quad \exp\left[ A^{({\rm vertex})}_{b1} + A^{({\rm vertex})*}_{b2}
+ A^{({\rm vertex})}_{f1} + A^{({\rm vertex})*}_{f2} \right] \times \nonumber \\
&& \quad \prod_e \sum_{m_e=0}^{\infty} \frac{(z^*z)^{m_e}}{(m_e!)^2}
\left\{
\left[ b_{L1}^*(e) b_{R1}(e) + \bar{f}_{L1}(e) f_{R1}(e)
\right]
\left[ b_{R2}^*(e) b_{L2}(e) + \bar{f}_{R2}(e) f_{L2}(e)
\right]
\right\}^{m_e} \,.
\label{eq:G2}
\end{eqnarray}
The expression~\eqref{eq:G2} for $\overline{|G|^2}$ can be
interpreted graphically as follows. Each term in the expansion of
the product corresponds to a pair of paths $(\gamma_1,\gamma_2)$,
where $\gamma_1$ respects the orientation of the lattice $\cal L$
(forward path) and $\gamma_2$ follows the reverse orientation
(backward path). The two paths must use each edge $e$ the same
number of times $m_e$. Paths configurations are weighted by the
elements of the vertex $S$-matrix, and an additional factor
$(z^*z)^{m_e}$. Note that closed loops have a vanishing weight,
because the bosonic and fermionic contributions cancel each other.
\subsection{Truncation procedure}
\label{sec:truncation}
In the form~\eqref{eq:G2}, $\overline{|G|^2}$ can be viewed as a
two-point correlation function in a classical, two-dimensional
statistical model for two-colour path configurations. No
approximation has been introduced so far, and thus \eqref{eq:G2}
is identical to the value of $\overline{|G|^2}$ in the original CC
model. The main difficulty in evaluating \eqref{eq:G2} is that the
paths $\gamma_1, \gamma_2$ may go through a given edge an
arbitrary number of times $m_e$, and thus the statistical model
has an infinite number of degrees of freedom per edge. This type
of problem is not usually tractable by exact solution methods, so
we need to {\it truncate} the statistical model to a finite loop
model in order to use these methods. This is very analogous to
what Nienhuis did for the $\On$ spin model~\cite{nienhuis-On} on
the hexagonal lattice: in that context, the spin model with
variables $( {\bf S}_j \in \mathbb{R}^n, {\bf S}_j^2=1 )$ was
formally mapped to a polygon model where edges could be used an
arbitrary number of times, but the substitution $e^{J {\bf S}_i
\cdot {\bf S}_j} \to 1 + J {\bf S}_i \cdot {\bf S}_j$ in the edge
interaction led to a finite loop model, while preserving the $\On$
symmetry of the original spin model.
The truncation we propose consists in keeping only the terms
of~\eqref{eq:G2} with $m_e \in \{0,1\}$, {\it i.e.} the
configurations where each of the paths $\gamma_1,\gamma_2$ visits
an edge at most once. This preserves the boson/fermion
supersymmetry, ensuring that closed loops still have a vanishing
weight in the truncated model. This can be seen as follows. In the
original expression~\eqref{eq:edge} for the action on the edges,
we can imagine choosing a different fugacity $z_e$ for each edge
(so that it now appears inside the summation over $e$.) This does
not affect the supersymmetry of the action. On expanding in powers
of all the $z_e$, the bosonic contribution to a given edge now
enters with a factor $(z_e^* z_e^{\phantom{*}})^{m_e}$. Thus our truncation to
$m_e\in\{0,1\}$ amounts to keeping only the terms up to first
order in the expansion of the partition function in powers of
$z_e^* z_e^{\phantom{*}}$, and then setting all the $z_e=z$ again.. Note that to
this order we have either nothing, or a pair of bosons of
different flavours (1 and 2), or a pair of fermions of different
flavours, propagating along each edge. The supersymmetry ensures
that each closed loop is counted with weight $0$. At this stage it
is simpler to switch to a replica formulation rather than using
supersymmetry explicitly: we have a model with two flavours of
boson, such that each edge is either unoccupied, or occupied by
each flavour exactly once. Each closed loop is counted with a
fugacity $n$, taking then $n=0$. The vertices are shown in
Fig.~\ref{fig:cc-model}.
We briefly comment on how higher order truncations would look in
this expansion. For example, at ${\rm O}\big((z_e^* z_e^{\phantom{*}})^2\big)$ we would
have either 2 pairs of bosons of each flavour, or 1 pair of bosons
and 1 pair of fermions. (We can never have more than one pair of
fermions because the Grassmann variables square to zero.) Note
that in such a truncation we could give such a configuration a
weight different from $(z_e^* z_e^{\phantom{*}})^2$ and still preserve the
supersymmetry. This points to the existence of an
infinite-dimensional space of possible supersymmetric truncations.
However in this paper we consider only the simplest.
We denote by $\overline{|G(e_2,e_1,z)|^2_{\rm tr}}$ the truncated analog of
$\overline{|G(e_2,e_1,z)|^2}$: $\overline{|G(e_2,e_1,z)|^2_{\rm tr}}$ is given by
the same expression as~\eqref{eq:G2}, but with the sum running
only over $m_e=0,1$. Then $\overline{|G(e_2,e_1,z)|^2_{\rm tr}}$ is interpreted as
a two-point function in the loop model defined by the loop vertices
of Fig.~\ref{fig:cc-model} and with loop weight $n=0$.
In the original CC model, the parameter~$\beta$ in the $S$-matrix
(\ref{eq:Smatrix}) is staggered. It is useful to consider an
anisotropic version of this, where it takes the value $\beta$ on
the even sublattice of $\cal L$ and $\beta'$ on the odd
sublattice. In this anisotropic CC model, the critical line
is~\cite{CC}:
\begin{equation} \label{eq:critical-beta}
\sinh \beta \ \sinh \beta' = 1 \,.
\end{equation}
The Boltzmann weights of the truncated loop model are defined in
Fig.~~\ref{fig:cc-model}. For general $\beta,\beta'$ they take
the values:
\begin{equation} \label{eq:weights-cc1}
\begin{array}{lcl}
t\phantom{'}, u_1, u_2, w_1, w_2, x\phantom{'}
= 1, \ a\phantom{'}, \ b\phantom{'},
\ a^2\phantom{'}, \ b^2\phantom{'}, \ -a \phantom{'} b
&\quad& \hbox{(even sublattice),} \\
t', u'_1, u'_2, w'_1, w'_2, x' = 1, \ b', \ a', \ b'^2, \ a'^2, \ -a'b'
&\quad& \hbox{(odd sublattice),}
\end{array}
\end{equation}
where
\begin{equation} \label{eq:weights-cc2}
\begin{array}{lcl}
a\phantom{'}:= z^2 \cosh^{-2} \beta\phantom{'} \,,
&\quad& b\phantom{'} := z^2 \tanh^2 \beta\phantom{'} \,, \\
a':= z^2 \cosh^{-2} \beta' \,,
&\quad& b':= z^2 \tanh^2 \beta' \,.
\end{array}
\end{equation}
Note that at the isotropic point these weights are
\begin{equation} \label{eq:isocritical}
1, \ z^2/2, \ z^2/2, \ z^4/4, \ z^4/4, \ -z^4/4 \,.
\end{equation}
\begin{figure}[th]
\begin{center}
\includegraphics{cc-model.eps}
\caption{Vertices of the loop model arising from the truncation of the CC model.}
\label{fig:cc-model}
\end{center}
\end{figure}
\subsection{Critical properties}
\label{sec:critical-prop}
We now discuss the observables of the model. The most important is
the mean square Green's function between two edges
$\overline{|G(e_2,e_1,z)|^2}$. At $z=1$ in the untruncated model
this gives the point conductance. For a system without any open
boundary contacts, this is identically equal to one by
conservation of probability. It is given by the sum over all pairs
of Feynman paths going out and back from $e_1$ to $e_2$, such that
each edge is traversed the same number of times in the forward
path as in the return path, and weighted by the appropriate
$S$-matrix elements of the CC model. It has been argued
\cite{gruzberg-unpub} that the weights for such `pictures' are all
positive. For $z \to 1^{-}$ and on the critical line
\eqref{eq:critical-beta} we expect a scaling form
\begin{equation} \label{eq:G2scaling}
\overline{|G(e_2,e_1,z)|^2} \sim
r^{-2X_{|G|^2}} \, F \left[ r(1-z)^{1/d_f} \right] \,,
\end{equation}
where $r=|e_1-e_2|$, $X_{|G|^2}=0$, and $d_f$ is the fractal dimension of
these pictures (whereby their total mass $M$ behaves as
$r^{d_f}$). The absence of the prefactor $r^{-2X_{|G|^2}}$ is a
consequence of the fact that $|G(z=1)|^2=1$ for a closed system.
In the truncated model, we no longer have probability conservation
and so the point $z=1$ is no longer special. Instead, in analogy
with other loop models, we expect to find a different critical
point, at $z=z_c$, say, such that the average loop length is
finite for $z<z_c$ and diverges for $z\geq z_c$. The point
conductance $\overline{|G(e_2,e_1,z)|^2_{\rm tr}}$ now corresponds to the
weighted sum of a pair of black and grey paths connecting $e_1$
and $e_2$. On the critical line and as $z\to z_c^{-}$ we expect the
same scaling form as in \eqref{eq:G2scaling} with $(1-z)$ replaced by $(z_c-z)$,
but not necessarily with same exponents as in the full model.
In Figure~\ref{fig:zc-trcc} and Table~\ref{table:zc}, we show the numerical determination
of $z_c$ using the two largest eigenvalues $\Lambda_0,\Lambda_1$ of the transfer matrix.
These eigenvalues define the thermal exponent $X_t$ through the CFT form
of the free-energy gap:
\begin{equation}
\log \frac{\Lambda_0}{\Lambda_1} \simeq \frac{2\pi X_t}{L} \,.
\end{equation}
\begin{figure}[th]
\begin{center}
\includegraphics{zc-trcc.eps}
\caption{Numerical determination of the critical monomer fugacity $z_c$ in
the model of Fig.~\ref{fig:cc-model}. On the $y$-axis is plotted the
effective thermal exponent $X_t(L,\beta_c,z) = \frac{L}{2\pi} \log \frac{\Lambda_0}{\Lambda_1}$.
}
\label{fig:zc-trcc}
\end{center}
\end{figure}
\begin{table}
\begin{center}
\begin{tabular}{c|r|r|r|r}
$L$ & 4 & 6 & 8 & 10 \\
\hline
$z_c(L,L+2)$ & 1.029885 & 1.030895 & 1.031454 & 1.031695
\end{tabular}
\caption{Finite-size estimates of the critical monomer fugacity $z_c$ in the
model of Fig.~\ref{fig:cc-model}. The value $z_c(L,L+2)$ is defined as the solution of
$X_t(L,\beta_c,z)=X_t(L+2,\beta_c,z)$, where $X_t(L,\beta,z)= \frac{L}{2\pi} \log \frac{\Lambda_0}{\Lambda_1}$ is the
effective thermal exponent.}
\label{table:zc}
\end{center}
\end{table}
More generally, it is possible to consider `watermelon' exponents
$X_{\ell_1,\ell_2}$ corresponding to $\ell_1$ black and $\ell_2$
grey paths originating from the vicinity of a given edge. The
truncation constraint of course implies that these cannot
originate on the same edge for $\ell>1$, but we imagine taking the
scaling limit where edges a finite distance apart on the lattice
are mapped to the same point. These operators are well suited for
a transfer-matrix-based numerical analysis~\cite{ds-tm}.
In particular we see that $X_{1,1}$ corresponds to $X_{|G|^2}$ in~\eqref{eq:G2scaling}.
Also, since $z^*z$ counts the number of edges connected to 2 black and 2 grey paths,
we have
\begin{equation}\label{eq:X22}
d_f = 2-X_{2,2}\,.
\end{equation}
Using transfer-matrix diagonalisation, we obtain the value
\begin{equation}
X_{2,2} = X_t \simeq 0.3 \,.
\end{equation}
Finally, we evaluate the correlation-length exponent $\nu$ associated to
a perturbation of the parameter $\beta$ away from $\beta_c$. For the lowest
free-energy gap we expect the scaling form
\begin{equation}
\log \frac{\Lambda_0}{\Lambda_1} \simeq \frac{2\pi}{L}
\ F\left[ (\beta-\beta_c) \ L^{1/\nu}
\right] \,.
\end{equation}
The best data collapse is obtained for the value (see Fig.~\ref{fig:nu-trcc}):
\begin{equation}
\nu \simeq 1.1 \,.
\end{equation}
\begin{figure}[th]
\begin{center}
\includegraphics{nu-trcc.eps}
\caption{Data collapse for the effective thermal exponent $X_t(L,\beta,z_c)$,
under a perturbation of the parameter $\beta$. The value used for this plot is $1/\nu=0.9$.}
\label{fig:nu-trcc}
\end{center}
\end{figure}
\subsection{Relation to Hilbert-space truncation}
\label{sec:Marston-etal}
We close this Section by comparing our approach to earlier studies~\cite{KM,MT}
of the IQHE problem based on a different truncation procedure.
Our method consists in writing the lattice path integral representation for
the mean conductance using the supersymmetry
trick, and then truncating the infinite sum over the paths, to keep only the
self-avoiding paths. This gives us the well-defined loop model of
Fig.~\ref{fig:cc-model}, where we will tune slightly the Boltzmann weights
to obtain an integrable point (see Section~\ref{sec:integrable}).
In contrast, in~\cite{KM,MT}, one starts from a two-dimensional single-particle
Hamiltonian including Gaussian hopping coefficients, and computes its
supersymmetric path integral.
The resulting action is then interpreted as the action of a one-dimensional
many-body supersymmetric Hamiltonian $H_{\rm MB}$, given in Eqs.~(3--5) of~\cite{MT}. This
Hamiltonian is expressed in terms of the coefficients $S^a$ of a superspin matrix.
In this model, the Hilbert space for each site is infinite-dimensional
(each site can be occupied by an arbitrary number of bosons).
The idea is to truncate this Hilbert space down to dimension $D$, and follow
the behaviour of the energy gap as $D$ increases. The model is not critical for finite
$D$, but it becomes critical in the limit $D \to \infty$.
Let us shown how to relate the terms of $H_{\rm MB}$ in the truncated space of dimension
$D=5$, to the generators which encode the loop model of Fig.~\ref{fig:cc-model}.
We first get rid of the $(-1)^j$ factor in $H_{\rm MB}$~\cite{MT}. This is
done through the change
\begin{equation*}
c_{\up j} \to -c_{\up j} \qquad
\hbox{for $j \equiv 2 \mod 4$ \quad or \quad $j \equiv 3 \mod 4$} \,,
\end{equation*}
without affecting the
(anti-)commutation relations for the $b_j,c_j$. We obtain the Hamiltonian:
\begin{equation}
H_{\rm MB} = \sum_{j=1}^L \left[
\sum_{a=1}^{16} g_a S^a_j S^a_{j+1}
+\eta (S^1_j + S^2_j + S^5_j + S^6_j)
\right] \,,
\end{equation}
where the signs $g_a$ are given by
\begin{equation}
g_a = \begin{cases}
1 & \hbox{if $a = 1,2,10,12,14,16$} \\
-1 & \hbox{if $a = 3,\dots,9,11,13,15$.}
\end{cases}
\end{equation}
We decompose $H_{\rm MB}$ as a sum of generators
\begin{equation}
H_{\rm MB} = \sum_{j=1}^{L} \Bigg\{
-\dcup_j - \dcap_j + e_j + f_j
+ (1+\eta) \Big[ \idl_j + \idr_j + 2 \idlr_j \Big]
\Bigg\} \,,
\end{equation}
where we have defined
\begin{equation}
\begin{array}{rcl}
\dcup_j &:=& S_j^3 S_{j+1}^3 + S_j^7 S_{j+1}^7
+ S_j^{15} S_{j+1}^{15} - S_j^{16} S_{j+1}^{16} \\
\\
\dcap_j &:=& S_j^4 S_{j+1}^4 + S_j^8 S_{j+1}^8
+ S_j^{13} S_{j+1}^{13} - S_j^{14} S_{j+1}^{14} \\
\\
\idl_j + \idlr_j &:=& \half \left( S^1_j + S^2_j + S^5_j + S^6_j \right) \\
\\
\idr_j + \idlr_j &:=& \half \left( S^1_{j+1} + S^2_{j+1} + S^5_{j+1} + S^6_{j+1} \right) \\
\\
e_j &:=& \left( S_j^1-\half \right) \left( S_{j+1}^1-\half \right)
- \left( S_j^5+\half \right) \left( S_{j+1}^5+\half \right)
+ S_j^{10} S_{j+1}^{10} + S_j^{12} S_{j+1}^{12} \\
\\
f_j &:=& \left( S_j^2-\half \right) \left( S_{j+1}^2-\half \right)
- \left( S_j^6+\half \right) \left( S_{j+1}^6+\half \right)
- S_j^9 S_{j+1}^9 - S_j^{11} S_{j+1}^{11} \,.
\end{array}
\end{equation}
In terms of the creation/annihilation operators, the above generators read:
\begin{equation}
\begin{array}{rcl}
\dcup_j &=& (b_{j\up}^\dag b_{j+1\up}^\dag + c_{j\up}^\dag c_{j+1\up}^\dag)
(b_{j\dow}^\dag b_{j+1\dow}^\dag - c_{j\dow}^\dag c_{j+1\dow}^\dag) \\
\dcap_j &=& (b_{j\up} b_{j+1\up} + c_{j\up} c_{j+1\up})
(b_{j\dow} b_{j+1\dow} - c_{j\dow} c_{j+1\dow}) \\
e_j &=& (b_{j\up}^\dag b_{j+1\up}^\dag + c_{j\up}^\dag c_{j+1\up}^\dag)
(b_{j\up} b_{j+1\up} + c_{j\up} c_{j+1\up}) \\
f_j &=& (b_{j\dow}^\dag b_{j+1\dow}^\dag - c_{j\dow}^\dag c_{j+1\dow}^\dag)
(b_{j\dow} b_{j+1\dow} - c_{j\dow} c_{j+1\dow}) \\
\idl_j + \idlr_j &=& \half (b_{j\up}^\dag b_{j\up} + c_{j\up}^\dag c_{j\up}
+ b_{j\dow}^\dag b_{j\dow} + c_{j\dow}^\dag c_{j\dow}) \\
\idr_j + \idlr_j &=& \half (b_{j+1\up}^\dag b_{j+1\up} + c_{j+1\up}^\dag c_{j+1\up}
+ b_{j+1\dow}^\dag b_{j+1\dow} + c_{j+1\dow}^\dag c_{j+1\dow}) \,.
\end{array}
\end{equation}
In the $D=5$ truncated space, each site is either empty or occupied by two particles of
opposite spins $(\up,\dow)$. If each spin is interpreted as a loop color, the above generators
(when restricted to the $D=5$ space) obey a dilute two-color Temperley-Lieb algebra with loop
weight $n=0$. Hence, they represent the vertices $u_1, u_2, w_1, x$ of the loop model defined in
Section~\ref{sec:truncation}. In particular, the $e_j$ and $f_j$ form two decoupled
Temperley-Lieb algebras.
Note that, in this context, the generator for the $w_2$ vertex, $E_j=e_j f_j$,
cannot be realised by a linear combination of the $S^a_j S^a_{j+1}$, but it may be
a linear combination of the $(S^a_j S^a_{j+1})^2$. So introducing $E_j$ terms in the Hamiltonian
leads to higher-order terms in $H_{\rm MB}$, and most probably it breaks the invariance
with respect to the supersymmetric charges $Q_{1,2}$.
However, we have shown that the supersymmetric model $H_{\rm MB}$, when restricted to the
$D=5$ space, corresponds to a particular manifold in the phase diagram of the two-color
loop model.
\section{Construction of an integrable critical loop model}
\label{sec:integrable}
In the preceding section, we truncated the the Chalker-Coddington
network model to yield a two-color loop model that is simpler to
analyze. To make further progress, we modify this model further. We
augment it by allowing the ``straight-line'' vertices with weight $v$
illustrated in Fig.~\ref{fig:model}. We also generalize it by
allowing the weight per loop $n$ to not only be zero, but to vary in
the range $n \in [-2,2]$. By utilizing the results of \cite{GW},
we will show in this section that
this modified model for all values of $n$ in this range
has an integrable line, and includes several critical points.
The remainder of the paper will be devoted to
the study of the critical behaviour.
When the straight-line vertices are allowed, the loop model can no
longer be related directly to electron trajectories in a potential.
In the original CC model, the `checkerboard' structure of the lattice
(or, equivalently, the alternation of arrows on the edges of $\cal L$)
is essential to the interpretation of the paths as electron
trajectories along the contour lines of the random potential. However,
several arguments indicate that the truncated but unmodified loop model
of Fig.~\ref{fig:cc-model} is in the same universality class as that
of the modified model. In other words, one can obtain the unmodified
model by perturbing the critical line with irrelevant operators.
\begin{figure}[th]
\begin{center}
\includegraphics{model.eps}
\caption{Vertices of the augmented dilute two-color loop model.}
\label{fig:model}
\end{center}
\end{figure}
One argument for the equivalence of the two stems from the relation of
this two-color loop model to that studied in \cite{Paul-Jesper}. There
the completely packed version was studied; in the notation used here
this corresponds to setting the Boltzmann weights $t= u_1= u_2= v= 0$. It
was shown that at least for weight per loop $n\ge \sqrt{2}$, the model
has a critical point when $x/w_2$ is tuned appropriately. Moreover, at
this critical point, numerical evidence strongly suggests that
dilution (i.e.\ non-zero $t$, $u_1$ and $u_2$) is irrelevant. We will
provide additional evidence by finding that for certain discrete
values of $n\ge \sqrt{2}$, the critical point of the completely packed
model and that of the modified model studied here are described by the
same conformal field theory. Neither of these arguments
applies when $n=0$, but all the critical exponents we have computed ($X_t, X_{2,2}, \nu$)
for both the truncated CC model and the integrable model at $n=0$ in regime 4
(see Table~\ref{table:regimes}) agree, up to our numerical precision.
This strongly indicates that the integrable model at $n=0$ in regime 4
is in the universality class of the truncated CC model.
In this section, we give the Boltzmann weights of the integrable
critical line in the loop model of Fig.~\ref{fig:model}. These weights
are expressed in terms of the generators of the dilute
Birman-Wenzl-Murakami (dBWM) algebra, so that the solution of the
Yang-Baxter equation found in \cite{GW} can be used. In Appendix~A,
we review the BWM algebra and its graphical
presentation. The braid group can be represented in terms of the BWM
generators, and can then be used to find invariants of knots and links
generalizing the Jones polynomial \cite{BWM}.
An alternate way of obtaining the Boltzmann weights of the integrable critical
line is to search for holomorphic observables on the lattice. These
are operators whose expectation values satisfy the lattice analog of
the Cauchy-Riemann equations. This method is described in Appendix B,
and yields the same weights as those found in \cite{GW} using the dBWM
algebra.
\subsection{Critical completely packed loop models}
\label{sec:CPL}
We first review the critical completely packed loop model,
arising for example in the Fortuin-Kasteleyn expansion of the Potts
model \cite{FK}. Each vertex of this model has the two possible
configurations displayed in Fig.~\ref{fig:TL}.
\begin{figure}[ht]
\begin{center}
\includegraphics[scale=0.8]{TLtiles.eps}
\end{center}
\caption{Action of $\idop$ (left) and $e_j$ (right) on a pair of strands
at positions $j$ and $j+1$. The transfer matrix direction is
upwards.}
\label{fig:TL}
\end{figure}
The partition function is conveniently written
in terms of the generators of the Temperley-Lieb (TL) algebra \cite{TL}.
This algebra for a system of width $L$ has
$L$ generators $e_j$ acting at positions $j=1,2,\ldots,L$ as well as the
identity $\idop$, which obey the relations
\begin{equation} \label{TLalg}
e_j^2 = n \ e_j \,,
\qquad e_j e_{j\pm 1}e_j = e_j \,,
\qquad e_i e_j = e_j e_i \quad\hbox{for $|i-j|>1$.}
\end{equation}
The first of the relations encodes the fact that the weight for a
closed loop is $n$, while the second encodes the fact that
the weight does not depend on the length or the shape of the loop.
The Boltzmann weights of the integrable critical loop model are then
\begin{equation}
\Rc_j(u)= \sin(2\theta-u) \, \idop - \sin u \, e_j \,,
\end{equation}
where $n=-2\cos 2\theta$ and $|n|\le 2$.
The transfer matrix for an even number of sites $L$ is then
\begin{equation}
T = \Rc_1 \Rc_3 \dots \Rc_{L-1} \Rc_2 \Rc_4 \dots \Rc_L \, .
\end{equation}
It is straightforward to use the TL algebra to verify that
these Boltzmann weights satisfy the Yang-Baxter equation
\begin{equation}
\Rc_j(u) \Rc_{j+1}(u+v) \Rc_j(v) = \Rc_{j+1}(v) \Rc_j(u+v) \Rc_{j+1}(u)
\end{equation}
and the inversion relation
\begin{equation}
\Rc_j(u) \Rc_j(-u) = \sin(2\theta-u) \sin(2\theta+u) \ \idop \,.
\end{equation}
Braid group generators $b_j$ and $b_j^{-1}$ are found by taking $u \to \pm i\infty$:
$$
\Rc_j (i\infty)\propto b_j = e^{-i\theta}\, \idop + e^{i\theta}\, e_j,
\qquad
\Rc_j (-i\infty)\propto b^{-1}_j = e^{i\theta}\, \idop + e^{-i\theta}\, e_j \ .
$$
These satisfy the braid-group relations \eqref{eq:braid1} and
\eqref{eq:braid2} as a consequence of the Yang-Baxter equation and
the inversion relation respectively.
The critical completely packed loop model on the square lattice is in
the same universality class as what is usually known as the $\On$
model in its dense phase. Well-established results on the dense $\On$
model~\cite{nienhuis-On} give the central charge of the CFT
describing the scaling limit to be
\begin{equation}
c_{\On} = 1 - \frac{3(\pi -2\theta)^2}{\pi\theta} \ .
\label{eq:cOn}
\end{equation}
The Boltzmann weights of the completely packed doubled loop model
studied in \cite{Paul-Jesper,Martins-Nienhuis} can be written in terms
of the generators $e_i$ and $f_i$ of two independent TL
algebras. This model is displayed in Fig.~\ref{fig:model} with $t=u_1=
u_2=v=0$. In this picture, the $e_j$ acts on black loops while the
$f_j$ act on grey loops, while the transfer matrix goes to the
northeast. Thus the vertex with weight $w_1$ corresponds to the
generator $\idop$, the vertex with weight $w_2$ corresponds to $e_j f_j$,
while those with weight $x$ are $e_j$ and $f_j$. Since the $e_j$'s and
the $f_j$'s commute, we have immediately that the $B_j, B^{-1}_j$
defined by
\begin{equation} \label{eq:def-Bj}
\begin{array}{rcl}
B^{\phantom{-1}}_j &:=& \left(e^{-i\theta} \, \idop + e^{i\theta} \, e_j \right)
\left(e^{-i\theta} \, \idop + e^{i\theta} \, f_j \right)\\
B^{-1}_j &:=& \left(e^{i\theta} \, \idop + e^{-i\theta} \, e_j \right)
\left(e^{i\theta} \, \idop + e^{-i\theta} \, f_j \right)
\end{array}
\end{equation}
also generate a braid group.
Similarly, TL generators with loop weight $N=n^2$
may be constructed as
\begin{equation}
E_j := e_j \ f_j \,.
\end{equation}
Using the relations (\ref{TLalg}) for the $e_j$'s and $f_j$'s, it is
straightforward to show that the $B_j, B^{-1}_j, E_j$ generate the BWM
algebra described in Appendix~A with parameters
$N=n^2=(-2\cos 2\theta)^2$, $\omega=e^{i6\theta}$
~\cite{Paul-Jesper}. The doubled lines here correspond to the single
lines displayed in Appendix~A, as is apparent by comparing
Figs.~\ref{fig:model} and \ref{fig:dBWM-gen}. Writing the Boltzmann weights
in terms of this algebra is useful because solutions of the
Yang-Baxter equation involving the BWM generators have long been known
\cite{Wadati}. From this solution, a critical point for the coupled
completely packed loop models for $n\ge \sqrt{2}$ was found
\cite{Paul-Jesper,Martins-Nienhuis}.
With the parameterisation
$$
n= 2\cos \frac{\pi}{r+2} \,,
$$
in the isotropic case $w_1=w_2$,
the critical point is at $x/w_1 = \lambda_c$, where
\begin{equation}
\lambda_c =
-\sqrt{2} \sin \left[ \frac{\pi(r-2)}{4(r+2)} \right] \,.
\label{eq:lambdac}
\end{equation}
At integer values $r=2,3,4 \dots$, this critical point was
identified with a particular
conformal field theory, the WZW coset model $\SU(2)_r \times \SU(2)_r/\SU(2)_{2r}$.
This conformal field theory has central charge
\begin{equation}
c_r = \frac{3r^2}{(r+1)(r+2)}\ .
\label{eq:cr}
\end{equation}
For $1/\lambda_c < x/w_1 < \lambda_c$, the doubled loop model has a critical
phase corresponding to two decoupled completely packed loop
models. The central charge is thus twice (\ref{eq:cOn}).
\subsection{The integrable critical line}
\label{sec:integrable-line}
We now can use the results of Grimm and Warnaar \cite{GW} to find an
integrable model involving all the vertices in Fig.\
\ref{fig:model}. We are interested mainly in the critical points, which
can be written in terms of the dilute BWM algebra. The dilute BWM algebra extends
the BWM algebra described in Appendix~A to include edges of the lattice uncovered by
strands. In the two-color loop model, these amount to allowing vertices to
be empty of both colors. The dilute generators act identically on the
two colors, and so include the remaining vertices in Fig.\
\ref{fig:model}. In an obvious notation, we then can write the
$\Rc$-matrix as
\begin{eqnarray}
\Rc_j(\varphi) &=& t(\varphi) \vac_j
+ u_1(\varphi) \big[ \idl_j + \idr_j \big]
+ u_2(\varphi) \big[ \dcup_j + \dcap_j \big] \nonumber \\
&& + v(\varphi) \big[ \vup_j + \vdown_j \big]
+ w_1(\varphi) I_j + w_2(\varphi) E_j + x(\varphi) X_j \,.
\label{eq:R-mat}
\end{eqnarray}
In terms of the TL generators introduced in the previous
section, $E_j=e_j f_j$ and $X_j \equiv e_j + f_j$, while $I_j$ takes value $0$ on the
dilute configurations and $1$ otherwise.
Since the non-dilute vertices satisfy the BWM algebra, it is simple to
show that the operators
$$
B_j, E_j, I_j, (\ \ )_j, \idl_j, \idr_j, \dcup_j, \dcap_j,
\vup_j, \vdown_j
$$
constructed from the two-color loop model satisfy a dilute BWM algebra.
Namely, with doubled lines here corresponding to single lines in
Appendix~A, and the $B_j$ defined in \eqref{eq:def-Bj}, these operators
generate the dilute BWM algebra with parameters $(N=(q+q^{-1})^2,
\omega=q^3)$, where $q=e^{2i\theta}$.
In~\cite{GW}, an integrable model based on the dBWM algebra was derived.
With $n=-2 \cos 2\theta$ as before, its Boltzmann weights are given by
\begin{equation} \label{eq:weights}
\begin{array}{rcl}
t(\varphi) &=& -\cos(2\varphi-3\theta) - \cos 5\theta + \cos 3\theta + \cos \theta \\
u_1(\varphi) &=& -2 \sin 2\theta \sin(\varphi-3\theta) \\
u_2(\varphi) &=& \phantom{-} 2 \sin 2\theta \sin \varphi \\
v(\varphi) &=& -2 \sin \varphi \sin(\varphi-3\theta) \\
w_1(\varphi) &=& \phantom{-} 2 \sin(\varphi-2\theta) \sin(\varphi-3\theta) \\
w_2(\varphi) &=& \phantom{-} 2 \sin \varphi \sin(\varphi-\theta) \\
x(\varphi) &=& \phantom{-} 2 \sin \varphi \sin(\varphi-3\theta) \,.
\end{array}
\end{equation}
We denote by $\varphi_0$ the isotropic value which is closest to zero:
\begin{equation} \label{eq:phi0}
\varphi_0 =\begin{cases}
\frac{3\theta}{2}
& \hbox{if $0< \theta < \frac{\pi}{3}$} \\
\frac{3\theta}{2}-\pi
& \hbox{if $\frac{\pi}{3} < \theta < \pi$.}
\end{cases}
\end{equation}
The universal properties are independent of the anisotropy parameter
$\varphi$ (as long as $\varphi$ lies between $0$ and $\varphi_0$), but depend very strongly on
$\theta$, as we shall see. At the isotropic point
$\varphi=3\theta/2$, the weights can be rescaled to
\begin{equation}
\begin{array}{l}
t = 2\cos 3\theta + 2 \cos 2\theta + 1 \\
u_1=u_2 = 4 \cos \frac{\theta}{2} \cos \theta \\
v = 2\cos \theta + 1 \\
w_1=w_2 = 1 \\
x = -(2\cos \theta + 1) \,.
\end{array}
\end{equation}
The integrable model defined by \eqref{eq:R-mat}--\eqref{eq:weights} obeys the
following properties:
\begin{itemize}
\item The isotropic weights are invariant under the transformations
$\theta\to 2\pi+\theta$, and $\theta\to -\theta$, so the range of
inequivalent couplings is $\theta\in [0,\pi]$. Each value of $n\in [-2,2]$
appears twice in this interval.
\item Since there are no loop ends, the
number of loops mod 2 is the same as the number of $x$ vertices mod
$2$. This allows us to change the sign of $n$ by absorbing the sign
in the weight $x$: $(n,x) \to (-n,-x)$. Thus there are four distinct
critical points for each value of $n\in (0,2)$, while there are two
for $n=0$ and $n=2$.
\item The weights satisfy the inversion relation
\begin{equation}
\label{eq:inversion}
\check{R}(\varphi)\check{R}(-\varphi)
= 4 \sin(2\theta-\varphi)\sin(2\theta+\varphi)
\sin(3\theta-\varphi)\sin(3\theta+\varphi) \ \idop \,.
\end{equation}
\item Rotating by 90$^{\rm o}$ is equivalent to sending
$\varphi \to 3\theta-\varphi$.
\item The weights are trivial when $u=0$:
$\check{R}(0)= 2 \sin 2\theta \sin 3\theta \ \idop \,.$
\item The eigenvalues of the transfer matrix are preserved under
$(u_1,u_2) \to (-u_1,-u_2)$ and $v \to -v$.
\end{itemize}
\subsection{The quantum Hamiltonian}
\label{sec:Hamiltonian}
To gain intuition into this doubled loop model, it is useful to find
the equivalent 1d quantum Hamiltonian by taking the very anisotropic
limit $\varphi \to 0$. The Hamiltonian is found from the transfer
matrix $T_L(\varphi)$ for $L$ sites by
$$
H := 2\sin 2\theta \sin 3\theta \ \left.
\frac{{\rm d}\log T_L(\varphi)}{{\rm d}\varphi}
\right|_{\varphi=0}
+ 2L \sin 5\theta \ \idop \ ,
$$
yielding
\begin{eqnarray}
H &=& \sum_{j=1}^L \Big\{
4 \cos 4\theta \sin \theta \ \vac_j
+ 2 \cos 2\theta \sin 3\theta \ \left[ \idl_j + \idr_j \right]
+ 2 \sin 2\theta \ \left[ \dcup_j + \dcap_j \right] \nonumber \\
&& \qquad +2 \sin 3\theta \ \left[ \vup_j + \vdown_j \right]
-2 \sin \theta \ E_j -2 \sin 3\theta \ X_j
\Big\} \,.
\label{eq:ham}
\end{eqnarray}
To find the Fermi velocity $v_f$, we assume that in the scaling limit
this Hamiltonian is that of a conformal field theory. In the next
Section, we will present much evidence in support of this assumption.
In a conformal field theory, the ground-state energy
(the lowest eigenvalue of $H$) is \cite{E-CFT}
\begin{equation} \label{eq:E-CFT}
E^0_L \simeq L e_\infty
-\frac{\pi c}{6L} \ v_f \,.
\end{equation}
where $c$ is the central charge. Let $\Lambda^0_L(\varphi)$ be the
dominant eigenvalue of the transfer matrix. The analysis of Appendix~B
indicates that the free energy of the loop model on a rhombic
lattice with angle $\alpha$ is given by $\left[-\log
\Lambda^0_L(\varphi) \right]$, where $\alpha = \pi\varphi/(2\varphi_0)$ and $\varphi_0$ is the isotropic value, as
defined in \eqref{eq:phi0}.
In a conformal field theory, one expects \cite{E-CFT}
\begin{equation} \label{eq:Lambda-CFT}
- \log \Lambda^0_L(\varphi)\simeq L f_\infty(\alpha)
-\frac{\pi c}{6L} \sin \alpha \,.
\end{equation}
Differentiating \eqref{eq:Lambda-CFT} around $\varphi=0$ and
comparing with \eqref{eq:E-CFT} yields
\begin{equation} \label{eq:vf}
v_f = \left|
\frac{2\pi \sin 2\theta \sin 3\theta}{2\varphi_0}
\right| \,,
\qquad \varphi_0 =\begin{cases}
\frac{3\theta}{2}
& \hbox{if $0< \theta < \frac{\pi}{3}$} \\
\frac{3\theta}{2}-\pi
& \hbox{if $\frac{\pi}{3} < \theta < \pi$.}
\end{cases}
\end{equation}
\section{Identifying the critical theories}
\label{sec:critical-th}
In this Section, we present what we believe is convincing evidence that
the doubled loop model with Boltzmann weights~\eqref{eq:weights} is
critical. We find the presumably exact central charge of the conformal
field theories describing the scaling limit, and also give some of the
dimensions of fields. We do this by a combination of calculations
exploiting the integrability, comparison to a similar integrable
model, and exact diagonalization of the transfer matrix and the Hamiltonian
for widths up to $L=14$ sites.
\subsection{The four regimes}
\label{sec:regimes}
This critical line is parametrized by the value of $\theta\in[0,\pi]$,
related the weight per loop by $n=-2\cos 2\theta$. Since the Fermi
velocity vanishes at $\theta=\frac{\pi}{3}, \frac{\pi}{2}$, and has a discontinuity at
$\theta=\frac{2\pi}{3}$, it is natural to expect that the physics is
discontinuous if $\theta$ is varied across these values. We thus
divide the critical line into four regimes, as described in
Table~\ref{table:regimes}.
All known integrable models with Boltzmann weights parameterized by
trigonometric functions of the anisotropy parameter $\varphi$ are
critical, and this is no exception. One argument for this is the
existence of the lattice holomorphic operator described in Appendix B.
Another is the inversion-relation calculation done below, which shows
that with standard assumptions about holomorphicity in $\varphi$, the free
energy is singular as this critical point. A numerical check is to use
exact diagonalization to find the largest eigenvalue of $T$ and/or the
ground-state energy of $H$, and then fit the results to \eqref{eq:E-CFT} or
\eqref{eq:Lambda-CFT}. To extract the central charge $c$, we use two
different-length systems to get rid of the extensive piece
$L e_\infty$. Doing this, we find the results given in
Fig.~\ref{fig:cc}. We see a very nice convergence to the
critical behavior as expected.
\begin{table}
\begin{center}
\begin{tabular}{|c|c|c|c|}
\hline
{\bf regime} & {\bf $\theta$-range} & {\bf parameterisation}
& {\bf central charge} \\ \hline
& & & \\
1 & $0<\theta<\frac{\pi}{3}$ & $n=-2\cos \frac{\pi}{r+2}$ &
$c=2 \left[ 1 - \frac{6}{(r+1)(r+2)} \right]+ \half$ \\
& & & \\ \hline
& & & \\
2 & $\frac{\pi}{3} < \theta < \halfpi$ & $n= \phantom{-} 2\cos \frac{\pi}{r+2}$ &
$c=2 \left[1 - \frac{6}{(r+1)(r+2)} \right]$ \\
& & & \\ \hline
& & & \\
3 & $\halfpi < \theta < \frac{2\pi}{3}$ & $n= \phantom{-} 2\cos \frac{\pi}{r+2}$ &
$c= \frac{3 r^2}{(r+1)(r+2)} + \half$ \\
& & & \\ \hline
& & & \\
4 & $\frac{2\pi}{3} < \theta < \pi$ & $n= -2\cos \frac{\pi}{r+2}$ &
$c= \frac{3 r^2}{(r+1)(r+2)}$ \\
& & & \\ \hline
\end{tabular}
\caption{The four regimes of the integrable loop model.}
\label{table:regimes}
\end{center}
\end{table}
We combine these results with other arguments to conjecture
exact formulae for the central charge for all $\theta$. We can also
identify precisely which conformal field theories describe some
critical lines. There are two types of conformal field theories known
to describe doubled loop models, and both occur along
this critical line. Unfortunately, the value of $n=0$ at
$\theta= \frac{3\pi}{4}$ of interest for the truncated CC model lies in one of
the regions where we do not understand the conformal field theory. As
is apparent from Fig.~\ref{fig:cc}, we do know that $c=0$ as required
there.
\begin{figure}[th]
\begin{center}
\begin{tabular}{cc}
\includegraphics[width=.4\textwidth]{cc1.eps} &
\includegraphics[width=.4\textwidth]{cc2.eps} \\
\includegraphics[width=.4\textwidth]{cc3.eps} &
\includegraphics[width=.4\textwidth]{cc4.eps}
\end{tabular}
\caption{Numerical estimates for the central charge in the four
critical regimes. Different symbols represent data points for
consecutive system sizes:
$L=4,6 \ (+), L=6,8 \ (\times), L=8,10 \ (*), L=10,12 \ ({\scriptstyle \square}), L=12,14 \ ({\scriptstyle \blacksquare})$.
Full lines represent the predicted
exact values from Table~\ref{table:regimes}.}
\label{fig:cc}
\end{center}
\end{figure}
At several special values of $\theta$, the model simplifies. Namely,
when $n=\pm 1$, all loop configurations receive the same weight (if
$n=-1$, we transform $(n,x) \to (-n,-x)$ as explained in Sec.~\ref{sec:integrable-line}).
Thus when computing the
partition function, we can sum up the four completely packed vertices
to give a single one with weight $w_1+w_2+ 2nx$.
For $\theta=\frac{2\pi}{3}$ at the isotropic point, $x=v=0$, so this
reduces to a six-vertex model with no staggering. Here the usual
parameter~\cite{Baxbook} has value
$$
\Delta= \frac{a^2+b^2-c^2}{2ab} = -1 \,,
$$
so this is in the same universality class as the antiferromagnetic
Heisenberg model. Thus the central charge is $c=1$ and the
first thermal exponent is $X_t= \half$,
in agreement with the numerical results in Figs.~\ref{fig:cc}--\ref{fig:xt}.
At $\theta=\frac{\pi}{6}$ and
$\theta=\frac{5\pi}{6}$, we obtain a staggered version of the eight-vertex
model. Ordinarily the staggered eight-vertex model is not solvable,
but as we detail in Appendix~C, this one is not only solvable, but can
be mapped onto a free-fermion theory. There we show that there are two
Majorana fermions present, but only one of the two is critical. Thus
the central charge is $c=\half$ here, again consistent with the numerics.
\subsection{Computation of an exact scaling dimension}
\label{sec:exact-dim}
Since the model is integrable, it is possible to derive some
quantities exactly. Here we extract the dimension of an operator in
the critical theory as a function of $\theta$. This is possible
because at certain discrete values of $\theta$, there exists a
deformation away from the critical point preserving the
integrability \cite{GW}.
The inversion-relation method \cite{Baxbook} yields the
free energy along this deformation, and by analyzing its expansion
around the critical point, we extract the value of the exponent
$\nu_{\rm int}$. This then yields the dimension of the operator which
when added to the action causes the deformation.
It is convenient to parameterize the loop weight $n$ within each
of the four regimes by a parameter~$r$,
\begin{equation}
n= 2 \epsilon \cos \frac{\pi}{r+2} \,,
\end{equation}
where $\epsilon=-1$ in regimes 1 and 4, and $\epsilon=1$ in regimes 2 and 3
(see Table~\ref{table:regimes}).
The integrable deformations resulting in unitary field theories occur
at integer values of $r$ in all four regimes. Here the dilute BWM algebra
admits a ``height'' or ``Restricted Solid-On-Solid'' (RSOS)
realization~\cite{GW}. Instead of treating the loops as the degrees of
freedom, on the dual lattice one places height variables, which are
integers restricted to a certain interval. The loops then play the
role of domain walls separating regions of different heights.
The inversion-relation method is a way of computing the free energy
exactly after making assumptions about its holomorphicity
properties as a function of $\varphi$. The free energy satisfies
constraints following from the inversion relation
\eqref{eq:inversion2} below, and the fact that sending $\varphi\to
3\theta-\varphi$ rotates the lattice by $90^{\rm o}$. The holomorphicity
assumptions then give a unique solution to these
constraints. Parameterizing the deformation in our case by $p$,
the inversion relation becomes \cite{GW}
\begin{equation}
\label{eq:inversion2}
\check{R}(\varphi,p)\check{R}(-\varphi,p)
= (4p)^{-1}\, \theta_1(2\theta-\varphi,p)\,\theta_1(2\theta+\varphi,p)\,
\theta_1(3\theta-\varphi,p)\,\theta_1(3\theta+\varphi,p) \ \idop \,,
\end{equation}
where $\theta_1(u,p)$ is the standard elliptic theta function. This
indeed reduces to (\ref{eq:inversion}) in the critical limit $p\to
0$. From this, it is simple to show that the inverse of the transfer
matrix in the diagonal direction is given by forming a transfer
matrix out of products of $\Rc_j(-\varphi,p)$.
Conveniently, both (\ref{eq:inversion2}) and the behavior under
rotational symmetry are identical to that of the model studied in
\cite{WPSN}, so we utilize these results. The singular part of the
free energy approaching the critical point depends on $p$ as $f_{\rm
sing} \sim p^{2-\nu_{\rm int}}$, so the operator perturbing the
critical theory in the integrable direction has scaling dimension
$X_{\rm int}=2-2/\nu_{\rm int}$. Then we find
\begin{equation} \label{eq:xint}
X_{\rm int} = \begin{cases}
\frac{r-1}{r+2}+1 & \qquad \hbox{in regime\ 1,}\\
\frac{r-1}{r+2} & \qquad \hbox{in regime\ 2,}\\
\frac{3}{r+2}+1 & \qquad \hbox{in regime\ 3,}\\
\frac{3}{r+2} & \qquad \hbox{in regime\ 4.}
\end{cases}
\end{equation}
We have written these results in terms of $r$ instead of $\theta$ to
emphasize that the derivation only applies to $r$ integer, since this
is where (\ref{eq:inversion2}) can be derived. However, we
expect that the results can be continued to all $r$ within a given
regime, since the equations themselves depend on $r$ as a continuous parameter.
A good check of the validity of~\eqref{eq:xint} for generic $\theta$ is that
it corresponds exactly to $X_{\rm int} = 2s$, where $s$ is the spin of the
discretely holomorphic parafermion $\psi_s(z)$ described in Appendix~B:
see~\eqref{eq:s}.
\subsection{Description by conformal field theory}
\label{sec:CFT}
Here we give formulae for the central charges in all four regimes that
are presumably exact. All are related to those occurring in
completely packed models. However, the doubled loop models are not
identical to completely packed models: we have checked that
the doubled loops have non-trivial fractal
dimension in regime 4 (see Section~\ref{sec:other-exp}).
One type of critical behavior possible for a doubled loop model is
simply to have the two colors decouple in the scaling limit. This
occurs in the isotropic completely packed version when $x=-w_1=-w_2$,
and is argued to persist in a region around this point
\cite{Paul-Jesper}. The central charge is simply twice $c_{O(n)}$
given in (\ref{eq:cOn}). Examining our numerical results for $c$, we
see that in regime 2, the central charge indeed is converging nicely
to $2c_{O(n)}$ with the appropriate dependence on $n$. When $r$ is an
integer, this is twice the central charge of the conformal minimal
models, and so the corresponding height model should scale to two
decoupled minimal models. An additional check on this comes from the
fact that the dimension of the integrable perturbing operator in
(\ref{eq:xint}) is twice that of an operator in a minimal
model. Namely, we have $X_{\rm int}=2X_{1,2}$, where $X_{1,2}=
\frac{r-1}{2(r+2)}$ is the scaling dimension of the $\Phi_{1,2}$ operator in
the minimal model with central charge $c=1- \frac{6}{(r+1)(r+2)}$.
It is thus natural to conjecture that in regime 2, the scaling limit
of our integrable loop model is indeed that of two decoupled
completely packed loop models. These conformal field theories have
been extensively studied~\cite{minimal}.
In regime 1, the numerics for the central charge are apparently
converging to $2c_{O(n)}+\half$. Thus here the loops apparently decouple as
well, but additional critical Ising degrees of freedom appear. This
is consistent with the mapping to the Ising model at $\theta=\frac{\pi}{6}$ described in
Appendix~C. The scaling dimension $X_{\rm int}$ here is $2X_{1,2}+1$,
leading to the natural interpretation that the operator is a product
of the $\Phi_{1,2}$ operators in the two minimal models with the
energy operator in the Ising model, the latter having dimension $1$.
These extra Ising degrees of freedom, which also occur in
a certain regime of the square-lattice $\On$ model~\cite{square-On},
appear through the following mechanism. The vertices of the loop
model obey a $\mathbb{Z}_2$ symmetry, in the sense that any vertex is
surrounded by an even number of empty edges. Thus empty edges form
polygons where each node has even degree, and so they respect the
geometry of Ising domain walls for Ising variables $\sigma$ lying on
the dual lattice. Depending on the values of the Boltzmann weights,
these Ising variables may become critical in the continuum limit. This
is evidently what happens in regime 1.
The critical behavior in regimes 3 and 4 is not that of two decoupled
models. As mentioned above in Sec.~\ref{sec:CPL}, in the
completely packed version of the doubled loop model, there occurs a
coupled critical point corresponding to the $\SU_r(2) \times \SU_r(2) / \SU_{2r}(2)$ WZW
coset model, with central charge $c_r$ \eqref{eq:cr}.
The numerical analysis in Fig.\ \ref{fig:cc} nicely fits to
$c_r$ in regime 4, and agrees with the Ising value $c=\half$ at
$\theta=\frac{5\pi}{6}$, derived in Appendix~C. Moreover, when $r$ is an
integer, the exponent $X_{\rm int}$ in (\ref{eq:xint}) belongs to the
above coset theory. Thus it is natural to conjecture that the central
charge throughout regime 4 is $c_r$.
As we see from our numerics at $n=0$ (see Sec.~\ref{sec:other-exp}),
the fractal dimension of a single loop is $d_f<2$, so regime
4 represents a ``dilute branch'' of the coset theory.
Likewise, in regime 3 the data seem to be converging to $c_r+\half$,
agreeing with the six-vertex value $c=1$ at $\theta=\frac{2\pi}{3}$. The
exponent $X_{\rm int}$ is 1 greater than the value in regime 4, so it
is natural to interpret that the operator is multiplied by the Ising
energy operator of dimension of 1. Thus like in regime 1, the critical
theory presumably includes an extra Ising piece.
Outside $r$ integer, the conformal field theory in regimes 3 and 4 is
not understood. Moreover, we will see in the subsequent section that
even though the formula for the central charge is applicable for all
$r$, it is not even clear whether dimensions of exponents can be
continued to values of $|n|<1$.
\section{Critical behaviour in regime 4}
\label{sec:regime4}
The main motivation of this paper is to explore a doubled loop model
arising in the truncation of the Chalker-Coddington network model. For
a connection to disordered systems, the weight per loop $n$ and the
central charge $c$ must be zero. We have two $n=0$ points, but for
$\theta=\frac{\pi}{4}$ inside regime 1, the corresponding critical field
theory seems to have nothing to do with the CC model. Not only do the
different colors of loop decouple, but the extra Ising degree of
freedom makes $c\ne 0$. We thus in this section focus on the behavior
in regime 4, which contains the other $n=0$ point at $\theta=\frac{3\pi}{4}$.
As noted above, we do not have a conformal field theory description
valid in regime 4 outside of integer $r$. It therefore seems a good
idea to exploit the fact that the associated height
description at these points is described by the coset conformal field
theory $\SU(2)_r \times \SU(2)_r / \SU(2)_{2r}$.
\subsection{Integrable perturbations}
\label{sec:integrable-pert}
This coset theory is known to have two integrable perturbations. One
of them, found by using level-rank duality on the results of~\cite{Vaysburd},
is by the operator with dimension $X_{\rm int}=\frac{3}{r+2}$ discussed
above. This perturbation describes the scaling limit of the height
model with elliptic Boltzmann weights~\cite{GW}. In terms of the loop
model, we have found that the discrete parafermion~$\psi_s(z)$ of Appendix~B
is the chiral part of the corresponding operator. This parafermion consists
of the insertion of a one-leg defect for each loop flavour.
The other integrable perturbation also has a very natural
meaning in terms of loops. This perturbing operator corresponds
to the (1,1;adjoint) operator of dimension $X_{(1,1;{\rm adj})}=\frac{2r}{r+1}$.
Several arguments imply that this perturbation corresponds to changing the
weight per unit length of the loops \cite{Fendley06}. This integrable
field theory describes the scaling limit of an integrable height model
\cite{DJMO}, and using the BWM algebra, it is described in
\cite{Fendley06} how to relate this height model to a dilute doubled
loop model very similar to the one we study here. Moving away from the
critical point in this similar model turns out to be effectively
changing the weight per unit length.
The second argument implying this result involves the $S$ matrices for
this integrable field theory, which decompose into the tensor product
of $S$ matrices of two minimal models ${\cal S}_{r} \times {\cal
S}_{r}$ \cite{Z91}. It is natural to interpret the worldlines of a
particle in a single minimal model as a loop in the $\On$ model
\cite{Z90}. Thus when the $S$ matrix is given by this tensor product,
it is natural to interpret the worldlines of such particles as doubled
loops; when two particles scatter they obey one of the four processes
in the vertices $w_1$, $w_2$ and $x$ pictured in Fig.~\ref{fig:model}.
In such an interpretation, the weight per unit length
of the loop is related to the mass of the particle. In the field
theory, moving along this integrable line corresponds precisely to
varying the mass of the particle.
We denote $X_t$ the thermal exponent, defined as the conformal dimension
for the first excited state in the zero-leg sector.
The numerical calculation of $X_t$ (see Fig.~\ref{fig:xt}) brings two observations.
In the region $\frac{5\pi}{6} \lesssim \theta < \pi$, the thermal exponent $X_t$ converges
to $X_{(1,1;{\rm adj})}=\frac{2r}{r+1}$ even for generic values of $\theta$, whereas $X_{(1,1;{\rm adj})}$
was derived only for integer values of $r$. This indicates that the results
from the $\SU(2)_r \times \SU(2)_r / \SU(2)_{2r}$ coset WZW model may be continued to arbitrary
$\frac{5\pi}{6} \lesssim \theta < \pi$. However, in the region $\frac{2\pi}{3} < \theta \lesssim \frac{5\pi}{6}$,
$X_t$ clearly deviates from the continued value $\frac{2r}{r+1}$: this shows that not
all exponents of the loop model are given by analytic continuation of the WZW
coset model in this region, including our point of interest $\theta=\frac{3\pi}{4}$.
\begin{figure}[th]
\begin{center}
\includegraphics{xt.eps}
\caption{Thermal exponent $X_t$ in regime 4. Data points were obtained
by transfer-matrix diagonalisation, and the solid line represents the
exact result for the integrable perturbation dimension $X_{(1,1;{\rm adj})}=\frac{2r}{r+1}$.}
\label{fig:xt}
\end{center}
\end{figure}
\subsection{Correlation length exponent $\nu$}
\label{sec:nu}
The correlation length exponent $\nu$ is defined as the analog for the loop model of
$\nu_{\rm CC}$~\eqref{eq:nu-CC}. In the CC model, the effect of perturbing
the energy level $E$ away from the transition value $E_c$ amounts to taking
$\beta,\beta'$ out of the critical line~\eqref{eq:critical-beta}. The analog of this
perturbation in the integrable loop model is to introduce a staggering of the
spectral parameter between the even and odd sublattices, with symmetric values around
the isotropic spectral parameter $\varphi_0$~\eqref{eq:vf}:
$\varphi = (1 \pm \lambda)\varphi_0$, in the range $-1 \leq \lambda \leq 1$.
The parameter $\lambda$ acts in a similar way to $(\beta-\beta_c)$ in the original CC model.
At $\lambda=-1$, the only allowed loops are those with minimal length, winding around the
vertices of one sublattice (say, the even one). At $\lambda=1$ loops also have minimal length,
but wind around the odd sublattice. The critical transition takes place at $\lambda=0$, where
the two sublattices become equivalent, and loops may be very long.
In the limit $\lambda \to 0$, we expect this perturbation to develop a correlation length,
scaling as
\begin{equation} \label{eq:nu}
\xi \sim |\lambda|^{-\nu} \,.
\end{equation}
Since this staggering does not
respect the rapidity lines of the square lattice, it breaks integrability.
Let us first discuss the point $\theta=\frac{5\pi}{6}$, where the model maps to free fermions
and remains solvable when the $\lambda$ perturbation is included (see Appendix~C).
At this point, we get the analytical result $\nu=2$, whereas the energy operator of
the free-fermion theory is $X_t=1$. In Appendix~C, we show that the effective
theory is a massive Majorana fermion with mass proportional to $\lambda^2$
and not $\lambda$. Thus, at $\theta=\frac{5\pi}{6}$, we have the relation
between $\nu$ and the dimension of the perturbing operator:
\begin{equation} \label{eq:X-nu}
X_t = 2 - \frac{2}{\nu} \,.
\end{equation}
It is natural to assume that both $\nu$ and $X_t$ are continuous in $\theta$,
so the effective mass term should still be proportional to $\lambda^2$
outside $\theta=\frac{5\pi}{6}$, and the relation~\eqref{eq:X-nu} holds all along regime~4.
For $\theta \neq \frac{5\pi}{6}$, exponent $\nu$ is only
accessible numerically, through finite-size scaling.
The correlation-length exponent $\nu$
is obtained by assuming a one-parameter scaling
law for the energy gap in the presence of the staggered perturbation
$\lambda$. For $\lambda \simeq 0$, we expect the behaviour:
\begin{equation}
\log \frac{\Lambda_0(\lambda)}{\Lambda_1(\lambda)}
\simeq \frac{2\pi}{L} F\left( \lambda \ L^{1/\nu} \right)\,,
\end{equation}
where $F$ is a scaling function.
Since eigenvalues are unchanged under $\lambda \to -\lambda$,
$F$ must be an even function.
In particular, at $\theta=\frac{3\pi}{4}$, like for the truncated CC model
(see Sec.~\ref{sec:critical-prop}),
we get the best data collapse (see Fig.~\ref{fig:nu}) for the value
\begin{equation}
\nu \simeq 1.1 \,.
\end{equation}
To our numerical precision, this value may be related through~\eqref{eq:X-nu} to
the thermal exponent $X_t \simeq 1.71$. This is an indication that the
relation~\eqref{eq:X-nu} should hold throughout regime 4.
\begin{figure}[th]
\begin{center}
\includegraphics{nu.eps}
\caption{Data collapse for the effective thermal exponent $X_t(L,\lambda)$
in the presence of a $\lambda$ perturbation. The value used for this plot is $1/\nu=0.9$.}
\label{fig:nu}
\end{center}
\end{figure}
\subsection{Other exponents}
\label{sec:other-exp}
A finite-size scaling plot of the gap corresponding to
$\ell_1=\ell_2=1$ is shown in Fig.~\ref{fig:x1}. The fact that it
scales to zero faster than $1/L$ indicates that $X_{1,1}$ is
consistent with zero. This agrees with the value $X_{|G|^2}=0$ expected
for the untruncated model in (\ref{eq:G2scaling}), but as far
as we know there is no fundamental reason for them to agree, and
it would be interesting to investigate this further. The observed
slope in Fig.~\ref{fig:x1} suggests existence of an irrelevant
operator with scaling dimension $\approx 3.2$.
Moreover, we observe numerically that $X_t=X_{2,2}$. Hence, like for usual dilute polymers,
this means that $X_t$ is associated to a perturbation of the monomer fugacity
(but different from the coset-model continuation to $r=0$, which would yield
$X_{(1,1;{\rm adj})}=0$), and the fractal dimension of a path is
\begin{equation}
d_f = 2-X_{2,2} \simeq 1.71 \,.
\end{equation}
\begin{figure}[th]
\begin{center}
\includegraphics{x1.eps}
\caption{Log-log plot of the energy gap for the watermelon $(1,1)$ sector. Data points are
fitted by a line of slope $\simeq -2.2$, and hence the conformal dimension is $X_{1,1}=0$.}
\label{fig:x1}
\end{center}
\end{figure}
\section{Discussion}
\label{sec:discussion}
Regimes 3 and 4 of the integrable model are particularly
interesting, as the two loop colours remain coupled in the
continuum limit. They are described by the ``dilute branch'' of
the $\SU(2)_r \times \SU(2)_r / \SU(2)_{2r}$ WZW coset CFT (the same CFT as for the
completely packed case~\cite{Paul-Jesper}), with an additional
Ising degree of freedom in regime 3.
Strictly speaking, this theory is only valid at
the RSOS points, but some critical properties (including the
central charge) extend to the loop model for generic fugacity $n$.
However, differences between the loop and RSOS spectra exist, as
shown by our results on the thermal exponent $X_t$. An analytic
study of the loop model through the Bethe Ansatz Equations is
considered for future work. One also needs to understand if a
Coulomb-Gas construction (most probably with a three-dimensional
target space) could reproduce the coset results for generic $n$.
The original motivation for the present work was to propose an
exactly solvable approximation to the IQHE transition.
Unfortunately, our results for the correlation-length exponent
$\nu$ clearly indicate that the point $n=0$ of the integrable
model is not in the universality class of IQHE. However, this
model is a critical, integrable point in the phase diagram of our
modified CC model. It should really be considered as the first
order in a hierarchy of truncated models, converging to the IQHE
universality class. Higher-order truncated models should also
contain integrable points, which may be built by ``fusing'' the
edges of the first-order model, following~\cite{DJMO} or a more
recent approach based on additional $\mathbb{Z}_N$
symmetry~\cite{Z2}.
From the numerical point of view, we used the only known efficient method to study a generic
loop model: transfer-matrix diagonalisation. However, the inherent limitations on the system size
prevent us from obtaining sharp estimates for the exponents, especially for $\nu$. Recently,
new Monte-Carlo algorithms have been proposed to simulate 2d loop models~\cite{On-MC}.
We hope to adapt this new approach to two-colour models, and get more precise
estimates for the exponents of our truncated CC model.
\bigskip{\bf Acknowledgments}
The work of P.F. was supported in part by NSF grants
DMR/MPS 1006549 and 0704666.
\section*{Appendix A: The BWM algebra}
\renewcommand\thesection{A}
\setcounter{equation}{0}
\label{sec:BWM}
In this Appendix, we recall the motivation and definition of the BWM algebra in a graphical
language.
The BWM algebra~\cite{BWM} is a braid-monoid algebra, an object
relevant to knot theory. It was originally designed to compute a certain {\it link invariant},
and later it was realised that it could be represented by RSOS models related to affine
Lie algebras \cite{Wadati}. In~\cite{GW}, a dilute version of the BWM algebra was
constructed, together with the corresponding $R$-matrix.
In the context of knot theory, the basic objects under consideration are
{\it braids}. Let $(p_1, \dots, p_L)$ be $L$ distinct points in the
complex plane, and define two copies of each point in three-dimensional space,
$p'_j= p_j \times \{0\}, p''_j= p_j \times \{1\}$,
so that the points $\{p'_j\}$ and $\{p''_j\}$ lie in two parallel planes.
For all $j=1, \dots, L$, take a curve $\Gamma_j$ enclosed
between the two planes, and connecting $p'_j$ to $p''_j$.
Furthermore, impose that the $\Gamma_j$'s do not intersect each other.
Denote the multiplet $\Gamma = (\Gamma_1, \dots, \Gamma_L)$: a braid $\beta$ is then
an equivalence class of $\Gamma$'s, {\it modulo} continuous deformations of the curves.
A typical braid is depicted in Fig.~\ref{fig:braid}.
\begin{figure}[th]
\begin{center}
\includegraphics{braid.eps}
\caption{An element of the braid group (left) and a word of the braid-monoid algebra (right)
for $L=4$.}
\label{fig:braid}
\end{center}
\end{figure}
Multiplication of two braids $\beta, \beta'$
is defined by the concatenation of the two corresponding diagrams, with the convention
that diagrams act from bottom to top: the product $\beta \beta'$ corresponds
to $\beta$ above $\beta'$. The $\beta$'s form the braid group, generated by the
elementary braids $B_j, B_j^{-1}$, which satisfy the relations:
\begin{eqnarray}
B_j B_j^{-1} &=& B_j^{-1} B_j = \idop \label{eq:braid1} \\
B_j B_{j+1} B_j &=& B_{j+1} B_j B_{j+1} \label{eq:braid2} \\
B_j B_\ell &=& B_\ell B_j \qquad \text{if $|j-\ell| \geq 2$.} \label{eq:braid3}
\end{eqnarray}
Consider now multiplets $\Gamma = (\Gamma_1, \dots, \Gamma_L)$
of non-intersecting curves connecting all the elements of $\{p'_j\} \cup \{p''_j\}$,
without the restriction that a curve should go from a $p'_j$ to a $p''_k$.
The corresponding diagrams are then
words on the alphabet $\{B_j, B^{-1}_j, E_j, \ j=1, \dots, (L-1)\}$, where the meaning
of the letters $B_j, B^{-1}_j, E_j$ is given in Fig.~\ref{fig:bm-gen}.
The algebra on these words is called a braid-monoid algebra.
It has two parameters $(N, \omega)$, and is defined by the braid-group
relations~\eqref{eq:braid1}--\eqref{eq:braid2}, together with the additional
relations (see Fig.~\ref{fig:bm-rules}):
\begin{eqnarray}
E_j^2 &=& N \ E_j \label{eq:TL1} \\
E_j E_{j \pm 1} E_j &=& E_j \label{eq:TL2} \\
B_j E_j &=& E_j B_j = \omega \ E_j \label{eq:bm1} \\
B_j B_{j \pm 1} E_j &=& E_{j \pm 1} B_j B_{j\pm 1} = E_{j \pm 1} E_j \label{eq:bm2} \,.
\end{eqnarray}
Equations~\eqref{eq:TL1}--\eqref{eq:TL2} mean that the $E_j$ form a Temperley-Lieb algebra
with loop weight $N$.
\begin{figure}[th]
\begin{center}
\includegraphics{bm-gen.eps}
\caption{Generators of a braid-monoid algebra.}
\label{fig:bm-gen}
\end{center}
\end{figure}
\begin{figure}[th]
\begin{center}
\includegraphics{bm-rules.eps}
\caption{Algebraic rules in the braid-monoid algebra with parameters $(N,\omega)$.}
\label{fig:bm-rules}
\end{center}
\end{figure}
The BWM algebra is a braid-monoid algebra \eqref{eq:braid1}--\eqref{eq:bm2}
where one imposes a linear relation between $B_j$, $B_j^{-1}$ and $E_j$:
\begin{equation} \label{eq:BWM}
E_j = \idop + \frac{B_j - B_j^{-1}}{q-q^{-1}} \,,
\qquad \text{where}
\quad N = 1+ \frac{\omega-\omega^{-1}}{q-q^{-1}} \,.
\end{equation}
The reason for introducing such a constraint is that the resulting algebra
supports a linear form (the Markov trace) which is identical to a geometric invariant
of the diagrams $\Gamma$ (the Kauffman polynomial)~\cite{BWM}.
The dBWM algebra~\cite{GW} is obtained by allowing vacancies, or equivalently by taking
multiplets of curves $\Gamma=(\Gamma_1, \dots, \Gamma_\ell)$ with $0 \leq \ell \leq L$.
This amounts to adding the generators
\begin{equation*}
I_j, \vac_j, (\rangle \ )_j, (\ \langle )_j,
(\begin{smallmatrix} \cup \\ \phantom{\cap} \end{smallmatrix})_j,
(\begin{smallmatrix} \phantom{\cup} \\ \cap \end{smallmatrix})_j,
(\diagup)_j, (\diagdown)_j,
\end{equation*}
whose action is depicted in Fig.~\ref{fig:dBWM-gen}. In equations~\eqref{eq:braid1}
and \eqref{eq:BWM},
$\idop$ is replaced by $I_j$, so that the $B_j, B^{-1}_j, E_j$ still form a BWM algebra
on the set of occupied sites. Additional relations for the dilute generators should be
included, to implement invariance under continuous deformation of the curves in the
presence of vacancies. The full set of dBWM relations is given in~\cite{GW}.
\begin{figure}[th]
\begin{center}
\includegraphics{dBWM-gen.eps}
\caption{The generators of the dilute BWM algebra.}
\label{fig:dBWM-gen}
\end{center}
\end{figure}
\section*{Appendix B: Discretely holomorphic parafermion in the loop model}
\renewcommand\thesection{B}
\setcounter{equation}{0}
\label{sec:holo}
In this Appendix, we show that the two-colour loop model admits a
discretely holomorphic para\-fermion $\psi_s(z)$~\cite{Riva,Raj,IC} exactly on
the integrable manifold~\eqref{eq:weights}.
The parafermion $\psi_s(z)$ is defined on the midpoints of the dual lattice ${\cal L}^*$,
and inserts a one-leg defect for each color at point $z$.
In the two-point function $\langle \psi_s(0) \psi_s(z) \rangle$, there is a black
({\resp} grey) path $\gamma_1$ ({\resp} $\gamma_2$) connecting $0$ and $z$, and one
includes a phase factor involving the winding angles $W$ of the paths $\gamma_1, \gamma_2$:
\begin{equation}
\langle \psi_s(0) \psi_s(z) \rangle = \frac{1}{Z} \sum_{(\gamma_1,\gamma_2)_{0 \to z}}
\ \sum_{C|(\gamma_1,\gamma_2)} \Pi(C) \ e^{\frac{is}{2} [W(\gamma_1)+W(\gamma_2)]} \,,
\end{equation}
where the first sum is over all possible pairs of paths from $0$ to $z$, the
second sum is over the loop configurations $C$ compatible with $\gamma_1$ and $\gamma_2$,
and $\Pi(C)$ is the Boltzmann weight for a loop configuration $C$.
\begin{figure}[th]
\begin{center}
\includegraphics{parafermion.eps}
\caption{A loop configuration contributing to $\langle \psi_s(0) \psi_s(z)\rangle$.}
\label{fig:parafermion}
\end{center}
\end{figure}
We impose discrete Cauchy-Riemann (CR) on $\langle \psi_s(0) \psi_s(z) \rangle$:
\begin{equation} \label{eq:CR}
\sum_{z \in \square} \langle \psi_s(0) \psi_s(z) \rangle \ \delta z = 0 \,,
\end{equation}
where the sum is over the edges of an elementary plaquette of ${\cal L}^*$, and the
$\delta z$'s are the corresponding elementary displacements.
For the discrete CR equations~\eqref{eq:CR} to hold, it is sufficient to fix the external
loop configuration outside a given plaquette, and ask the total contribution
of internal configurations to vanish~\cite{Riva,IC}.
This determines a linear system of equations for the Boltzmann weights.
To get anisotropic solutions, we consider the analog problem on a rhombic
lattice of angle $\alpha$ \cite{Raj,IC}.
Setting
\begin{equation}
\lambda:= e^{\frac{i \pi s}{2}} \,, \qquad \mu:= e^{i\alpha(1+s)}\,,
\end{equation}
we get the $7 \times 7$ linear system for the unknowns $(t,u_1,u_2,v,w_1,w_2,x)$:
\begin{equation}
\label{eq:holo}
\begin{array}{rcl}
t+ \mu \lambda^{-2} u_1 - \mu u_2 - v &=& 0 \\
n^2 u_1 - \lambda^{-2} u_2 - \mu \lambda^2 v
+ \mu \lambda^{-2} (n^2 w_1 + w_2 + 2nx) &=& 0 \\
-\lambda^2 u_1 + n^2 u_2 + \mu \lambda^{-4} v
- \mu (w_1 + n^2 w_2 + 2nx) &=& 0 \\
-\mu \lambda^2 u_1 + \mu \lambda^{-4} u_2 + n^2 v
- \lambda^4 w_1 - \lambda^{-4} w_2 - 2x &=& 0 \\
n \lambda u_1 - n \lambda^{-1} u_2 - \mu \lambda^{-1} v + \mu \lambda^{-1}
[n(w_1+w_2) + (n^2+1)x] &=& 0 \\
n \lambda^{-1} u_1 - \lambda u_2 - n \mu \lambda v + \mu \lambda^{-3} (n w_1+x)
+ \mu \lambda (w_2+nx) &=& 0 \\
-u_1 + n \lambda^2 u_2 + n \mu \lambda^{-2} v - \mu \lambda^{-2} (w_1+nx)
- \mu \lambda^2 (n w_2+x) &=& 0 \,,
\end{array}
\end{equation}
As a first step, we need determine the spin $s$ by going back to the isotropic case
$\alpha=\pi/2$. Imposing $u_1=u_2$ and $w_1=w_2$, \eqref{eq:holo} reduces to a $5 \times 5$
system, whose determinant is:
\begin{equation} \label{eq:det}
D(n, \lambda) = -\lambda^{-4} (\lambda^2+1)(n\lambda^4-1)^2
(\lambda^4+\lambda^{-4} + n^3-3n) \,.
\end{equation}
Using the parameterisation $n= -2 \cos 2\theta$, this determinant vanishes when:
\begin{equation} \label{eq:s}
\exp(2i\pi s) = \exp(\pm 6i\theta) \,.
\end{equation}
For a general angle $\alpha$, the solution of~\eqref{eq:holo} is a set of $\alpha$-dependent
weights $t(\alpha), \dots x(\alpha)$. If we apply the substitution:
\begin{equation} \label{eq:alpha-phi}
\alpha \to \frac{\varphi}{1+s} \,,
\end{equation}
we observe that the solution of~\eqref{eq:holo} is identical to the integrable
weights~\eqref{eq:weights}. This is analogous to what was found for various
other integrable models with a discrete holomorphic parafermion~\cite{Riva,Raj,IC}.
Note that the relation~\eqref{eq:alpha-phi} is consistent
with the discussion on the Fermi velocity in Section~\ref{sec:Hamiltonian}.
Moreover, at the Ising points $\theta=\frac{\pi}{6}, \frac{5\pi}{6}$ (see Appendix~C), the spin $s=\frac{1}{2}$
is consistent with~\eqref{eq:s}.
\begin{figure}[th]
\begin{center}
\includegraphics[scale=0.8]{holo.eps}
\caption{External loop connectivities outside an elementary plaquette.}
\label{fig:holo}
\end{center}
\end{figure}
\section*{Appendix C: Free fermions at $\theta= \frac{\pi}{6}, \frac{5\pi}{6}$}
\renewcommand\thesection{C}
\setcounter{equation}{0}
\label{sec:free-fermions}
\subsection{Mapping to a staggered 8V model}
At $\theta=\frac{\pi}{6}$ and $\theta=\frac{5\pi}{6}$, the model maps to a free fermion
Hamiltonian through the mapping sequence:
\begin{center}
two-color loop $\to$ square-lattice ${\rm O}(n=1)$ $\to$ staggered 8V $\to$ free fermions.
\end{center}
In both cases, we start from the two-color loop model and do the sign change
$(n,x) \to (-n,-x)$ to get $n=1$. With this value of $n$, one may discard loop connectivities,
and the model is simply local, with occupied and empty edges. After the exchange of
occupied/empty edges, we get a square-lattice $\On$ model~\cite{square-On} with $n=1$
and weights~:
\begin{equation}
\wt{t} = w_1+w_2 -2x,
\quad \wt{u}_1=u_1,
\quad \wt{u}_2=u_2,
\quad \wt{v}=v,
\quad \wt{w}_1+\wt{w}_2= t \,.
\end{equation}
(The tilde is here to avoid confusion between the $\On$ and two-colour
loop model Boltzmann weights).
For $\theta= \frac{\pi}{6}$, we have the specific values
\begin{equation}
\begin{array}{rcl}
\wt{t}_{\phantom{1}} &=& \sin 2\varphi -\sqrt{3} \\
\wt{u}_1 &=& \phantom{-} \sqrt{3} \ \cos \varphi \\
\wt{u}_2 &=& - \sqrt{3} \ \sin \varphi \\
\wt{v}_{\phantom{1}} &=& \sin 2\varphi \\
\wt{w}_1 &=& - \cos(\frac{\pi}{6}-2\varphi) - \frac{\sqrt{3}}{2} \\
\wt{w}_2 &=& \phantom{-} \cos(\frac{\pi}{6}+2\varphi) - \frac{\sqrt{3}}{2} \,,
\end{array}
\end{equation}
which are exactly those of the integrable square-lattice $\On$ model~\cite{square-On}
with $n=1$ at the dilute critical point. The ${\rm O}(n=1)$ model maps in turn to an
eight-vertex model (see Fig.~\ref{fig:8V}) with staggered weights
\begin{equation} \label{eq:8V}
\begin{array}{rcl}
\omega_1=\omega_2 &=& \phantom{-}\sqrt{3} \cos\varphi \\
\omega_3=\omega_4 &=& -\sqrt{3} \sin\varphi \\
(\omega_5,\omega_6) &=& \begin{cases}
(-\sin 2\varphi -\sqrt{3}, \phantom{-}\sin 2\varphi -\sqrt{3}) & \hbox{on even sites} \\
(\phantom{-}\sin 2\varphi -\sqrt{3}, -\sin 2\varphi -\sqrt{3}) & \hbox{on odd sites}
\end{cases} \\
\omega_7=\omega_8 &=& \sin 2\varphi \,.
\end{array}
\end{equation}
At $\theta= \frac{5\pi}{6}$, one gets the same 8V model, up to irrelevant signs.
\begin{figure}[th]
\begin{center}
\includegraphics[scale=1]{8V.eps}
\caption{Correspondence between the 8V model and the ${\rm O}(n=1)$ model.
The mapping depicted here is valid on one sublattice, say even sites.
On odd sites, all arrows must be reversed.}
\label{fig:8V}
\end{center}
\end{figure}
\subsection{Very anisotropic limit: the XY chain in a magnetic field}
In terms of the Pauli matrices $\sigma_j$, the 8V $\Rc$-matrix reads:
\begin{eqnarray}
\Rc_j^{\rm 8V} &=& \frac{1}{4}(\omega_1+\omega_2+\omega_5+\omega_6) \idop
+ \frac{1}{4}(\omega_1+\omega_2-\omega_5-\omega_6) \sigma_j^z \sigma_{j+1}^z \nonumber \\
&&+ (\omega_3 \ \sigma_j^- \sigma_{j+1}^+ + \omega_4 \ \sigma_j^+ \sigma_{j+1}^-
+ \omega_7 \ \sigma_j^- \sigma_{j+1}^- + \omega_8 \ \sigma_j^+ \sigma_{j+1}^+) \nonumber \\
&&+ \frac{1}{4}(\omega_1-\omega_2) (\sigma_j^z+\sigma_{j+1}^z)
+ \frac{1}{4}(\omega_5-\omega_6) (\sigma_j^z-\sigma_{j+1}^z) \,.
\end{eqnarray}
We can now take the very anisotropic limit $\varphi \to 0$. Denoting by a prime the derivative
with respect to $\varphi$ at $\varphi=0$, and using the weights~\eqref{eq:8V}, we obtain:
\begin{equation}
\Rc_j^{'\rm 8V} = -\sqrt{3} (\sigma_j^- \sigma_{j+1}^+ + \sigma_j^+ \sigma_{j+1}^-)
+ 2(\sigma_j^- \sigma_{j+1}^- + \sigma_j^+ \sigma_{j+1}^+)
- (-1)^j (\sigma_j^z-\sigma_{j+1}^z) \,.
\end{equation}
The critical Hamiltonian is given by
\begin{equation} \label{eq:H0}
H_0 = -\half \sum_{j=1}^L \Rc_j^{'\rm 8V}
= - \frac{1}{4} \sum_{j=1}^L \left[
J_x \ \sigma_j^x \sigma_{j+1}^x - J_y \ \sigma_j^y \sigma_{j+1}^y
+ 2 h \ (-1)^j \sigma_j^z
\right] \,,
\end{equation}
where $J_x= (2-\sqrt{3})$, $J_y= (2+\sqrt{3})$ and $h=2$.
The alternating sign of the last term in~\eqref{eq:H0} can be eliminated
by the unitary change of basis:
\begin{equation}
H_0 \to U^\dag H_0 U \,,
\qquad \text{where} \qquad
U := \prod_{\ell=1}^{L/2} \sigma^x_{2\ell-1} \,.
\end{equation}
This maps $H_0$ to an XY chain in a magnetic field~\cite{XY,Ising}:
\begin{equation} \label{eq:H-XY}
H_{\rm XY} = U^\dag H_0 U = -\half \sum_{j=1}^L \left[
(1+\gamma) \ \sigma_j^x \sigma_{j+1}^x + (1-\gamma) \ \sigma_j^y \sigma_{j+1}^y
+ h \ \sigma_j^z
\right] \,,
\end{equation}
where $\gamma=-\frac{\sqrt{3}}{2}$ and $h=2$. In particular, we have learnt that this particular point
of the XY spin chain is exactly equivalent to the integrable dilute ${\rm O}(n=1)$ model.
\subsection{Staggered perturbation associated to $\nu$}
The setting of the 8V model also allows us to consider a staggered perturbation
like the one defining
exponent $\nu$ (see Sec.~\ref{sec:nu}). To do this, we introduce staggered
spectral parameters $(1 \pm \lambda) \varphi$ on the even/odd sites. We take the very
anisotropic limit $\varphi \to 0$,
with $\lambda$ fixed. This way, the parameter $\lambda$ controls the strength of
the perturbation.
The resulting Hamiltonian has the form
$$H(\lambda) = H_0 + \lambda H_1 \,$$
where
\begin{equation} \label{eq:H1}
H_1 = -\half \sum_{j=1}^L (-1)^j \Rc_j^{'\rm 8V}
= -\frac{1}{4} \sum_{j=1}^L (-1)^j \ \left(
J_x \ \sigma_j^x \sigma_{j+1}^x - J_y \ \sigma_j^y \sigma_{j+1}^y
\right) \,,
\end{equation}
and $J_x,J_y$ are the same as for $H_0$.
After the unitary change of basis defined by $U$, we get the
perturbing term:
\begin{equation} \label{eq:Hp}
H_p = U^\dag H_1 U =
- \frac{\lambda}{2} \sum_{j=1}^{L}
(-1)^j
\left[ (1+\gamma) \ \sigma_j^x \sigma_{j+1}^x
+ (1-\gamma) \ \sigma_j^y \sigma_{j+1}^y
\right] \,.
\end{equation}
\subsection{Exact free-fermion solution}
In this paragraph, we expose the exact solution of $H(\lambda)$ for general
values of $\gamma,h,\lambda$.
Following~\cite{XY}, we can solve the model $H(\lambda)$ by a Jordan-Wigner
transformation, mapping the Pauli matrices~$\sigma_j$ to fermion operators
\begin{equation}
c_j := \left( \prod_{\ell=1}^{j-1} \sigma^z_\ell \right) \ \sigma^+_j \,,
\qquad c^\dag_j := \left( \prod_{\ell=1}^{j-1} \sigma^z_\ell \right) \ \sigma^-_j \,,
\qquad c^\dag_j c^{\phantom \dag}_j = \half(1-\sigma^z_j) \,,
\end{equation}
obeying anti-commutation relations:
\begin{equation}
\{ c_j, c_\ell \}=0 \,,
\qquad \{ c_j, c^\dag_\ell \} = \delta_{j \ell} \,.
\end{equation}
In this language, the perturbed Hamiltonian reads
\begin{equation}
H(\lambda) =
- \sum_{j=1}^L \Bigg\{
\left[ 1 + (-1)^j \lambda \right]
\left[
c^\dag_j c^\nodag_{j+1} + c^\dag_{j+1} c^\nodag_j
+ \gamma (c^\dag_j c^\dag_{j+1} + c_{j+1} c_j)
\right]
+ h \left(\half - c^\dag_j c^\nodag_j \right)
\Bigg\} \,.
\end{equation}
We introduce two species of fermions
\begin{equation}
c_{1,\ell} := c_{2\ell} \,,
\qquad
c_{2,\ell} := c_{2\ell-1} \,,
\end{equation}
and their Fourier modes
\begin{equation}
c_{\mu,q} := \frac{1}{\sqrt{L/2}}
\sum_{\ell=1}^{L/2} e^{i\ell q} c_{\mu,\ell} \,,
\qquad q = \frac{2\pi m}{L/2} - \pi \,,
\quad m=1, \dots, \frac{L}{2} \,.
\end{equation}
We can now rewrite $H(\lambda)$ as
\begin{equation}
H(\lambda) = \sum_{-\pi < q \leq \pi}
\sum_{\mu,\nu} \left[
c^\dag_{\mu,q} A^\nodag_{\mu \nu, q} c^\nodag_{\nu,q}
+ \half \left( c^\dag_{\mu,q} B^\nodag_{\mu \nu, q} c^\dag_{\nu,-q}
- c_{\mu,-q} B_{\mu \nu, q} c_{\nu,q} \right)
\right] \,,
\end{equation}
where the matrices $A_q$ and $B_q$ read
\begin{equation}
A_q = \left( \begin{array}{cc}
h & \alpha_q^* \\
\alpha_q & h
\end{array} \right) \,,
\qquad B_q = \left( \begin{array}{cc}
0 & -\beta^*_q \\
\beta_q & 0
\end{array} \right) \,,
\qquad \begin{array}{rcl}
\alpha_q &:=& -[(1+\lambda)e^{iq} + (1-\lambda)] \,, \\
\beta_q &:=& \gamma[(1+\lambda)e^{iq} - (1-\lambda)] \,.
\end{array}
\end{equation}
Like in~\cite{XY}, the energies $\epsilon_\mu(q)$ are the square roots of
the eigenvalues of $(A_q+B_q)(A_q-B_q)$:
\begin{equation} \label{eq:epsq}
\epsilon_{1,2}(q) = 2 \sqrt{
\frac{h^2}{4} + (1+ \gamma^2 \lambda^2) \cos^2 \frac{q}{2}
+ (\gamma^2 + \lambda^2) \sin^2 \frac{q}{2}
\mp \sqrt{
4 \gamma^2 \lambda^2 + h^2
\left| \cos \frac{q}{2} + i\lambda \sin^2 \frac{q}{2} \right|^2
}
} \,.
\end{equation}
This is the two-branch dispersion relation for arbitrary $\gamma,h,\lambda$.
To express the corresponding eigenmodes, we need the unitary $2 \times 2$
matrices $W_q,V_q$ defined by the linear relations
\begin{equation}
(A_q - B_q) W_q = V_q D_q \,,
\qquad (A_q + B_q) V_q = W_q D_q \,,
\qquad D_q := \left(\begin{array}{cc}
\epsilon_1(q) & 0 \\
0 & \epsilon_2(q)
\end{array} \right) \,.
\end{equation}
The Bogoliubov transformation diagonalising $H(\lambda)$ is
\begin{equation} \label{eq:eta}
\eta_{\mu,q} := \half \sum_{\nu} \left[(W+V)^\dag_{\mu\nu,q} \, c^\nodag_{\nu,q}
- (W-V)^\dag_{\mu\nu,q} \, c^\dag_{\nu,-q}
\right] \,.
\end{equation}
Unitarity of $V_q$ and $W_q$ ensures the canonical
anticommutation relations
\begin{equation}
\{ \eta_{\mu,q}, \eta_{\mu',q'} \} = 0 \,,
\qquad \{ \eta^\nodag_{\mu,q}, \eta^\dag_{\mu',q'} \}
= \delta_{\mu \mu'} \delta_{q q'} \,.
\end{equation}
In terms of the $\eta$'s, the Hamiltonian reads
\begin{equation}
H(\lambda) = \sum_{-\pi < q \leq \pi}
\ \sum_{\mu=1,2} \epsilon_\mu(q)
\, \eta^\dag_{\mu,q} \eta^\nodag_{\mu,q} \,.
\end{equation}
Note that there are $L/2$ distinct momenta $q$,
and that each momentum corresponds to two modes $\mu=1,2$. Thus,
we recover $L$ independent modes $\eta_{\mu,q}$.
As a final step, we perform the change
\begin{equation} \label{eq:etah}
\eta_{\mu,q} \to \wt{\eta}_{\mu,q} = \begin{cases}
\eta_{\mu,q} & \text{if $q \geq 0$,} \\
\eta^\dag_{\mu,q} & \text{if $q < 0$,}
\end{cases}
\end{equation}
so that the modes with $q<0$ are now considered as holes.
the Hamiltonian becomes
\begin{equation}
H(\lambda) = \sum_{-\pi < q \leq \pi}
\ \sum_{\mu=1,2} \wt{\epsilon}_\mu(q)
\, \wt{\eta}^\dag_{\mu,q} \wt{\eta}^\nodag_{\mu,q} \,,
\qquad \text{where}
\qquad \wt{\epsilon}_\mu(q) := {\rm sgn}(q) \, \epsilon_\mu(q) \,.
\end{equation}
\subsection{Critical Majorana fermion at $h=2, \lambda=0$}
The dispersion relation of the XY chain in
a magnetic field is obtained by setting $\lambda=0$ in~\eqref{eq:epsq}:
\begin{equation}
\epsilon_{1,2}(q) = 2 \sqrt{
\left(\frac{h}{2} \mp \cos \frac{q}{2} \right)^2
+ \gamma^2 \sin^2 \frac{q}{2}
}
\qquad \qquad (\lambda=0) \,.
\end{equation}
For $h=2$, $\epsilon_1$ is critical at $q=0$, whereas $\epsilon_2$ is not critical\footnote{
In the case $h=\gamma=0$, both modes $\epsilon_1,\epsilon_2$ are critical at $q=\pi$,
and the corresponding field theory is a critical Dirac fermion.
}:
\begin{eqnarray}
\wt{\epsilon}_1(q) &=&
4 \sin \frac{q}{4} \sqrt{\sin^2 \frac{q}{4}
+ \gamma^2 \cos^2 \frac{q}{4}} \\
\wt{\epsilon}_2(q) &=&
4 \, {\rm sgn}(q) \cos \frac{q}{4} \sqrt{\cos^2 \frac{q}{4}
+ \gamma^2 \sin^2 \frac{q}{4}} \,.
\end{eqnarray}
The dispersion relation $\wt{\epsilon}_1(q)$ is approximately linear at $q=0$
(see Fig.~\ref{fig:eps}).
In the ground state, all levels with $-\pi<q<0$ are filled: this is a
Fermi sea with only one Fermi level $q_f=0$, and thus it corresponds
to a Majorana fermion with central charge $c=\half$.
The critical eigenmodes are obtained from~\eqref{eq:eta}:
\begin{equation} \label{eq:etac}
\left\{ \begin{array}{rcl}
\eta_{1,q} &=& \cos \frac{\theta_{1,q}}{2} \, (c_{1,q} + c_{2,q})
+ i \sin \frac{\theta_{1,q}}{2} \, (c^\dag_{1,-q} + c^\dag_{2,-q}) \,, \\
\eta_{2,q} &=& \cos \frac{\theta_{2,q}}{2} \, (c_{1,q} - c_{2,q})
+ i \sin \frac{\theta_{2,q}}{2} \, (c^\dag_{1,-q} - c^\dag_{2,-q}) \,,
\end{array} \right.
\end{equation}
where
\begin{equation} \label{eq:thetaq}
\left\{ \begin{array}{rcl}
\theta_{1,q} &:=& {\rm Arg} \left(
1-\cos\frac{q}{2} - i \sin \frac{q}{2} \right) \,, \\
\theta_{2,q} &:=& {\rm Arg} \left(
1+\cos\frac{q}{2} + i \sin \frac{q}{2} \right) \,.
\end{array} \right.
\end{equation}
After the change $\eta \to \wt{\eta}$~\eqref{eq:etah}, the continuum limit is
described by the effective Hamiltonian
\begin{equation}
H_0 \simeq i v_f \int {\rm d}x \ \wt{\eta}_1^\dag \partial_x \wt{\eta}^\nodag_1 \,.
\end{equation}
\begin{figure}[th]
\begin{center}
\begin{tabular}{cc}
\includegraphics[scale=1]{eps.eps}
& \includegraphics[scale=1]{eps2.eps}
\end{tabular}
\caption{The dispersion relation for $h=2, \gamma=-\frac{\sqrt{3}}{2}$.
Left: critical case $\lambda=0$. Right: $\lambda=\half$. The full ({\resp} dotted) lines
represent $\wt{\epsilon}_1$ ({\resp} $\wt{\epsilon}_2$).}
\label{fig:eps}
\end{center}
\end{figure}
\subsection{Gapped theory at $h=2,\lambda>0$}
When the $\lambda$ perturbation is turned on, an energy gap opens at
$q=0$ (see Fig.~\ref{fig:eps}):
\begin{equation} \label{eq:DE}
\Delta E := 2 \epsilon_1(q=0) = 2 \left|
\sqrt{1+\gamma^2 \lambda^2} - 1
\right| \sim \gamma^2 \lambda^\nu \,,
\qquad \nu=2 \,.
\end{equation}
To understand this value of $\nu$, we shall analyse the perturbing term $H_p$ in
terms of the critical modes $\wt{\eta}_{1,q}$~\eqref{eq:etac}.
The relations~\eqref{eq:etac} can be inversed, to give:
\begin{equation} \label{eq:aq}
\left\{\begin{array}{rcl}
c_{1,q}+c_{2,q} &=& \cos \frac{\theta_{1,q}}{2} \, \eta_{1,q}
- i \sin \frac{\theta_{1,q}}{2} \, \eta^\dag_{1,-q} := a_{1,q} \,, \\
c_{1,q}-c_{2,q} &=& \cos \frac{\theta_{2,q}}{2} \, \eta_{2,q}
- i \sin \frac{\theta_{2,q}}{2} \, \eta^\dag_{2,-q} := a_{2,q} \,.
\end{array} \right.
\end{equation}
The perturbing term has the expression
\begin{eqnarray}
H_p = \half \sum_{-\pi< q \leq \pi} \Bigg\{ &&
(1-\cos q) \, (a_{1,q}^\dag a^\nodag_{1,q} - a_{2,q}^\dag a^\nodag_{2,q}) \nonumber \\
&& + i\sin q \, \left[
a_{2,q}^\dag a^\nodag_{1,q} +
\frac{\gamma}{2} \, (a_{1,q}^\dag a_{1,-q}^\dag - a_{2,q}^\dag a_{2,-q}^\dag)
- {\rm h.c.}
\right] \nonumber \\
&& -\gamma (1+\cos q) \, (a_{2,q}^\dag a_{1,-q}^\dag + a_{1,-q} a_{2,q})
\quad \Bigg\} \,.
\label{eq:Hp2}
\end{eqnarray}
In the region $q \simeq 0$, the first term in~\eqref{eq:Hp2} is of order $q^2$,
and thus it generates irrelevant terms of the form $\wt{\eta}_1^\dag \partial_x^2 \wt{\eta}_1^\nodag$ in
the continuum limit. The second term is of order $q$, and corresponds to
$\wt{\eta}_1^\dag \partial_x \wt{\eta}_1^\nodag$, which renormalises the Fermi velocity.
At first order in $\lambda$, the third term has no effect on the continuum theory.
However, in second-order perturbation in $\lambda$, it generates terms of
the form $(\wt{\eta}^\dag_2 \wt{\eta}^\nodag_2)(\wt{\eta}^\dag_1 \wt{\eta}^\nodag_1)$, which are non-vanishing
since the lowest $\wt{\eta}_2$ modes are occupied in the ground state.
From this analysis, we obtain the effective Hamiltonian in the continuum limit
\begin{equation}
H_{\rm eff}(\lambda) \propto
\int {\rm d}x \ \left(i \, \wt{\eta}_1^\dag \partial_x \wt{\eta}^\nodag_1
+ {\rm const} \times \gamma^2 \lambda^2
\, \wt{\eta}_1^\dag \wt{\eta}^\nodag_1 \right) \,.
\end{equation}
The ``mass term'' has dimension $X_t=1$, and the energy gap thus scales as
$$
\Delta E \propto \left(\lambda^2 \right)^\frac{1}{2-X_t} \,.
$$
Comparing with~\eqref{eq:DE}, we get the scaling relation
\begin{equation}
X_t = 2 - \frac{2}{\nu} \,.
\end{equation}
|
\section*{Acknowledgements}}
\newcommand{\mathbb{C}}{\mathbb{C}}
\newcommand{\mathbb{Z}}{\mathbb{Z}}
\newcommand{\mathbb{R}}{\mathbb{R}}
\renewcommand{\P}{\mathbb{P}}
\newcommand{\mathcal{C}}{\mathcal{C}}
\newcommand{\mathcal{K}}{\mathcal{K}}
\newcommand{\mathcal{O}}{\mathcal{O}}
\newcommand{\mathrm{CAT}}{\mathrm{CAT}}
\newcommand{\mathrm{Gr}}{\mathrm{Gr}}
\newcommand{\mathrm{Hom}}{\mathrm{Hom}}
\newcommand{\mathrm{Inv}}{\mathrm{Inv}}
\newcommand{\mathrm{Irr}}{\mathrm{Irr}}
\newcommand{\mathrm{PGL}}{\mathrm{PGL}}
\newcommand{\mathrm{Sch}}{\mathrm{Sch}}
\newcommand{\mathrm{SL}}{\mathrm{SL}}
\renewcommand{\top}{\mathrm{top}}
\newcommand{{\mathrm{pt}}}{{\mathrm{pt}}}
\newcommand{\C_c}{\mathbb{C}_c}
\newcommand{\mathfrak{g}}{\mathfrak{g}}
\newcommand{{G^\vee}}{{G^\vee}}
\newcommand{\alpha^\vee}{\alpha^\vee}
\newcommand{\rho^\vee}{\rho^\vee}
\newcommand{{\vec{\lambda}}}{{\vec{\lambda}}}
\newcommand{{\vec{\gamma}}}{{\vec{\gamma}}}
\newcommand{{\vec{\mu}}}{{\vec{\mu}}}
\newcommand{{\vec{\nu}}}{{\vec{\nu}}}
\newcommand{\vec{p}}{\vec{p}}
\renewcommand{\tensor}{\otimes}
\newcommand{\longrightarrow}{\longrightarrow}
\newcommand{\hookrightarrow}{\hookrightarrow}
\newcommand{\twoheadrightarrow}{\twoheadrightarrow}
\newcommand{\stackrel{\mathrm{def}}{=}}{\stackrel{\mathrm{def}}{=}}
\newcommand{\epsilon}{\epsilon}
\renewcommand{\bar}{\overline}
\newcommand{\braket}[1]{{\langle #1 \rangle}}
\newcommand{\tilde{\times}}{\tilde{\times}}
\renewcommand{\min}{\textrm{min}}
\newcommand{\sqcup}{\sqcup}
\newcommand{\mathrm{\mathbf{hconv}}}{\mathrm{\mathbf{hconv}}}
\newcommand{\mathrm{\mathbf{econv}}}{\mathrm{\mathbf{econv}}}
\newcommand{\mathrm{\mathbf{fsp}}}{\mathrm{\mathbf{fsp}}}
\newcommand{\mathrm{\mathbf{perv}}}{\mathrm{\mathbf{perv}}}
\newcommand{\mathrm{\mathbf{rep}}}{\mathrm{\mathbf{rep}}}
\newcommand{\mathrm{\mathbf{spd}}}{\mathrm{\mathbf{spd}}}
\newcommand{\mathrm{\mathbf{vect}}}{\mathrm{\mathbf{vect}}}
\newcommand{\mathrm{\mathbf{svect}}}{\mathrm{\mathbf{svect}}}
\newtheorem{theorem}{Theorem}[section]
\newtheorem{proposition}[theorem]{Proposition}
\newtheorem{lemma}[theorem]{Lemma}
\newtheorem{question}[theorem]{Question}
\newtheorem{corollary}[theorem]{Corollary}
\newtheorem{conjecture}[theorem]{Conjecture}
\renewcommand{\theenumi}{\roman{enumi}}
\renewcommand{\labelenumi}{(\theenumi)}
\renewcommand{\figurename}{Figure}
\newenvironment{fullfigure}[2]
{\begin{figure}[htb]\begin{center}\def#1}\def\fullfigb{#2}{#1}\def\fullfigb{#2}}
{\vspace{\baselineskip}\caption{\fullfigb.}\label{#1}\def\fullfigb{#2}}
\end{center}\end{figure}}
\newenvironment{fullfigure*}[2]
{\begin{figure*}[htb]\begin{center}\def#1}\def\fullfigb{#2}{#1}\def\fullfigb{#2}}
{\vspace{\baselineskip}\caption{\fullfigb.}\label{#1}\def\fullfigb{#2}}
\end{center}\end{figure*}}
\newcommand{\eq}[2]{\begin{equation}\label{#1}#2\end{equation}}
\newcommand{\cor}[1]{Corollary~\ref{#1}}
\newcommand{\conj}[1]{Conjecture~\ref{#1}}
\newcommand{\fig}[1]{Figure~\ref{#1}}
\newcommand{\prop}[1]{Proposition~\ref{#1}}
\renewcommand{\sec}[1]{Section~\ref{#1}}
\newcommand{\thm}[1]{Theorem~\ref{#1}}
\newcommand{\lem}[1]{Lemma~\ref{#1}}
\newcommand{\emph{i.e.}}{\emph{i.e.}}
\newcommand{\emph{I.e.}}{\emph{I.e.}}
\newcommand{\vspace{-\baselineskip}}{\vspace{-\baselineskip}}
\colorlet{lightgray}{black!15!white}
\colorlet{darkgreen}{green!75!black}
\colorlet{darkblue}{blue!75!black}
\colorlet{darkred}{red!75!black}
\tikzset{midto/.style={postaction={decorate,
decoration={markings,mark=at position .5 with
{\draw (-.035,-.07) -- (.035,0) -- (-.035,.07);}}}}}
\tikzset{midfrom/.style={postaction={decorate,
decoration={markings,mark=at position .5 with
{\draw (.035,-.07) -- (-.035,0) -- (.035,.07);}}}}}
\tikzset{web/.style={darkblue,semithick}}
\begin{document}
\title{Buildings, spiders, and geometric Satake}
\author{Bruce Fontaine}
\affiliation{University of Toronto}
\author{Joel Kamnitzer}
\thanks{The first and second authors were partly supported by NSERC}
\affiliation{University of Toronto}
\author{Greg Kuperberg}
\thanks{The third author was partly supported by
NSF grants DMS-0606795 and CCF-1013079}
\affiliation{University of California, Davis}
\begin{abstract}
Let $G$ be a simple algebraic group. Labelled trivalent graphs called
webs can be used to product invariants in tensor products of minuscule
representations. For each web, we construct a configuration space of points
in the affine Grassmannian. Via the geometric Satake correspondence,
we relate these configuration spaces to the invariant vectors coming
from webs. In the case $G =\mathrm{SL}(3)$, non-elliptic webs yield a basis for
the invariant spaces. The non-elliptic condition, which is equivalent to
the condition that the dual diskoid of the web is $\mathrm{CAT}(0)$, is explained
by the fact that affine buildings are $\mathrm{CAT}(0)$.
\end{abstract}
\maketitle
\section{Introduction}
\subsection{Spiders}
Let $G$ be a simple, simply-connected complex algebraic group. In previous
work \cite{Kuperberg:spiders}, the third author defined a pivotal
tensor category with generators and relations called a ``spider", for
$G$ of rank 2. (The term ``spider" was originally intended to mean
any pivotal category, but in common usage only these categories are
called spiders.) The Karoubi envelope of this category is equivalent
to the category $\mathrm{\mathbf{rep}}^u(G)$ of finite-dimensional representations of $G$
with a modified pivotal structure. Actually, the spider comes with a
parameter $q$ making it equivalent to the quantum deformation $\mathrm{\mathbf{rep}}^u_q(G)$.
These results in rank 2 are analogous to the influential result of Kauffman
\cite{Kauffman:spinknot} and Penrose \cite{Penrose:negative} that the Karoubi
envelope of the Temperley-Lieb category (the category of planar matchings)
is equivalent to $\mathrm{\mathbf{rep}}^u_q(\mathrm{SL}(2))$. The Temperley-Lieb category can thus
be called the $\mathrm{SL}(2)$ spider. Conjectural generalizations of spiders
were proposed for $\mathrm{SL}(4)$ by Kim \cite{Kim:thesis} and for $\mathrm{SL}(n)$
by Morrison \cite{Morrison:diagram}.
In this article, for any $G$ as above, we will define the free spider for
$G$ generated by the minuscule representations of $G$. A morphism in the
free spider is given by a (linear combination) of labelled trivalent graphs
called webs. For each web $w$ with boundary edges labelled ${\vec{\lambda}}$,
there is an invariant vector
$$\Psi(w) \in \mathrm{Inv}(V({\vec{\lambda}})) = \mathrm{Inv}_G(V(\lambda_1) \tensor V(\lambda_2)
\tensor \cdots \tensor V(\lambda_n)).$$
If $G$ has rank 1 or 2, then the vectors $\Psi(w)$ coming from non-elliptic
webs $w$ (those whose faces have non-positive combinatorial curvature)
form a basis of each invariant space $\mathrm{Inv}(V({\vec{\lambda}}))$ of $G$, called
a web basis. The web basis for $\mathrm{SL}(2)$ is well-known as the basis
of planar matchings and it is known to be the same as Lusztig's dual
canonical basis \cite{FK:canonical}. On the other hand, the $\mathrm{SL}(3)$
web bases are eventually not dual canonical \cite{Kuperberg:notdual},
even though many basis vectors are dual canonical.
\subsection{Affine Grassmannians}
The goal of this article is to introduce a new geometric interpretation
of webs and spiders using the geometry of affine Grassmannians.
Let $\mathcal{O} = \mathbb{C}[[t]]$ and $\mathcal{K} = \mathbb{C}((t))$. In order to study the
representation theory of $G$, we will consider the affine Grassmannian of
its Langlands dual group
$$\mathrm{Gr} = \mathrm{Gr}({G^\vee}) = {G^\vee}(\mathcal{K})/{G^\vee}(\mathcal{O}).$$
The geometric Satake correspondence of Lusztig \cite{Lusztig:qanalog},
Ginzburg \cite{Ginzburg:loop}, and Mirkovi\'c-Vilonen \cite{MV:geometric}
will be our main tool in this article.
\begin{theorem} The category of equivariant perverse sheaves on the affine
Grassmannian $\mathrm{Gr}$ is equivalent as a symmetric and pivotal tensor category
to the tensor category $\mathrm{\mathbf{rep}}^u(G)$ of representations of $G$ with a modified
pivotal and symmetric structure.
\label{th:satake} \end{theorem}
As a consequence of this theorem, every invariant space $\mathrm{Inv}(V({\vec{\lambda}}))$
for every $G$ can be constructed from the geometry of $\mathrm{Gr}$. Given a
vector ${\vec{\lambda}}$ of dominant weights of $G$, there is a convolution morphism
$$m_{\vec{\lambda}}:\bar{\mathrm{Gr}({\vec{\lambda}})} = \bar{\mathrm{Gr}(\lambda_1)} \tilde{\times}
\bar{\mathrm{Gr}(\lambda_2)} \tilde{\times} \cdots \tilde{\times} \bar{\mathrm{Gr}(\lambda_n)}
\longrightarrow \mathrm{Gr},$$
where each $\mathrm{Gr}(\lambda)$ is a sphere of radius $\lambda$ (in the sense of
weight-valued distances \cite{KLM:generalized}) in $\mathrm{Gr}$. The fibre
$F({\vec{\lambda}}) = m_{\vec{\lambda}}^{-1}(t^0)$ is a projective variety that we call
the Satake fibre. In particular, we will use the following corollary of
the geometric Satake correspondence.
\begin{theorem} Every invariant space in $\mathrm{\mathbf{rep}}^u(G)$ is canonically
isomorphic to the top homology of the corresponding geometric Satake fibre
with complex coefficients:
$$\Phi:\mathrm{Inv}(V({\vec{\lambda}})) \cong H_\top(F({\vec{\lambda}}),\mathbb{C}).$$
Each top-dimensional component $Z \subseteq F({\vec{\lambda}})$ thus yields
a vector $[Z] \in \mathrm{Inv}(V({\vec{\lambda}}))$. These vectors form a basis, the
Satake basis.
\label{th:satbasis} \end{theorem}
A goal of this article is to understand how the invariant vectors coming
from webs expand in this basis. (Throughout, we will assume complex
coefficients for homology and cohomology.)
\subsection{Diskoids}
The orbits of $G(\mathcal{K})$ on the affine Grassmannian defines a notion
of distance on $\mathrm{Gr}$ with values in the set of dominant weights for
$G$. Thus, we can interpret $F({\vec{\lambda}})$ as the (contractive, based)
configuration space in $\mathrm{Gr}$ of an abstract polygon $P({\vec{\lambda}})$ whose
side lengths are
$${\vec{\lambda}} = (\lambda_1,\lambda_2,\ldots,\lambda_n).$$
One of our ideas is to generalize this type of configuration space
from polygons to diskoids. For us, a diskoid $D$ is a contractible
piecewise linear region in the plane; in many cases it is a disk. (See
\sec{s:config}.) If $D$ is tiled by polygons and its edges are labelled
by dominant weights, then its vertices are a weight-valued metric space.
We will define a (based) configuration space $Q(D)$ which consists of maps
from the vertices of $D$ to $\mathrm{Gr}$ that preserves the lengths of edges
of $D$. We will also define a special subset $Q_g(D)$ that consists of
maps that preserve all distances (globally isometric embeddings).
Assume that ${\vec{\lambda}}$ is a vector of minuscule highest weights. If $w$
is a web with boundary ${\vec{\lambda}}$, then it has a dual diskoid $D = D(w)$
(or possibly a diskoid with bubbles). The boundary of this diskoid is a
polygon $P({\vec{\lambda}})$ and so we get a map of configuration spaces $\pi:Q(D)
\to F({\vec{\lambda}})$. Our first main result is that we can recover the vector
$\Psi(w)$ using this geometry.
\begin{theorem} There exists a homology class $c(w) \in H_*(Q(D))$ such
that $\pi_*(c(w)) \in H_{\top}(F({\vec{\lambda}}))$ corresponds to $\Psi(w)$
under the isomorphism from \thm{th:satbasis}.
\label{th:homclass} \end{theorem}
We prove this theorem as an application of the geometric Satake
correspondence. In many cases, the class $c(w)$ is the fundamental class
of $Q(D)$, so that the coefficients of $\pi_*(c(w))$ (and hence $\Psi(w)$)
in the Satake basis are just the degrees of the map $\pi$ over the components
of $F({\vec{\lambda}})$.
\subsection{Buildings}
The affine Grassmannian $\mathrm{Gr}$ embeds isometrically into the affine
building $\Delta = \Delta({G^\vee})$. We can use this perspective to gain
greater insight into the variety $Q(D)$.
If $G = \mathrm{SL}(2)$, then a basis web is a planar matching (or cup diagram)
and its dual diskoid $D$ is a finite tree. The affine Grassmannian $\mathrm{Gr}$
is the set of vertices of the affine building $\Delta$, which is an infinite
tree with infinite valence. The configuration space $Q(D)$ is the space
of colored, based simplicial maps $f:D \to \Delta$; see \fig{f:a1example}.
It is known that
$$Q(D) = \P^1 \tilde{\times} \P^1 \tilde{\times} \cdots \tilde{\times} \P^1$$
is a twisted product of $\P^1$'s, and that these twisted products are
the components of the Satake fibre $F({\vec{\lambda}})$. Moreover, $Q_g(D)$ is
the open dense subvariety of points in $Q(D)$ which are contained in no
other component of $F({\vec{\lambda}})$. \fig{f:a1example} is an illustration
of the construction.
\begin{fullfigure*}{f:a1example}{From a non-elliptic $A_1$ web, to a tree,
to part of an affine $A_1$ building}
\begin{tikzpicture}[baseline]
\draw (0,0) circle (1);
\foreach \th in {0,3,5} {
\draw[web] (\th*45+67.5:1) .. controls (\th*45+67.5:.6)
and (\th*45+112.5:.6) .. (\th*45+112.5:1); };
\draw[web] (22.5:1) .. controls (22.5:.5)
and (157.5:.5) .. (157.5:1);
\end{tikzpicture}
{\large $\longrightarrow$}
\begin{tikzpicture}[baseline]
\coordinate (a) at (0,0); \coordinate (a0) at (90:.75);
\coordinate (a1) at (210:.75); \coordinate (a2) at (330:.75);
\coordinate (a00) at (90:1.5);
\draw (a00) -- (a) -- (a1); \draw (a) -- (a2);
\fill (a) circle (.075); \fill (a00) circle (.075);
\draw[fill=white] (a0) circle (.075);
\draw[fill=white] (a1) circle (.075);
\draw[fill=white] (a2) circle (.075);
\end{tikzpicture}
{\large $\longrightarrow$}
\begin{tikzpicture}[baseline]
\coordinate (a) at (0,0);
\foreach \x in {0,1,2} {
\coordinate (a\x) at (120*\x+90:.75);
\foreach \y in {0,1} {
\coordinate (a\x\y) at (120*\x+60*\y+60:1.5);
\foreach \z in {0,1} {
\coordinate (a\x\y\z) at (120*\x+60*\y+30*\z+45:2.25);
\foreach \w in {0,1} {
\coordinate (a\x\y\z\w) at (120*\x+60*\y+30*\z+10*\w+40:2.5);
}; }; }; };
\foreach \color/\r in {darkblue/.25,white/.225}
{
\path[fill=\color] (a) circle (\r);
\path[fill=\color] (a0) circle (\r);
\path[fill=\color] (a1) circle (\r);
\path[fill=\color] (a2) circle (\r);
\path[fill=\color] (a00) circle (\r);
\path[draw=\color,line width=\r*2cm] (a0) -- (a00);
\path[draw=\color,line width=\r*2cm] (a) -- (a0);
\path[draw=\color,line width=\r*2cm] (a) -- (a1);
\path[draw=\color,line width=\r*2cm] (a) -- (a2);
}
\fill[white] (a00) circle (.225);
\fill (a) circle (.075);
\foreach \x in {0,1,2} {
\foreach \r in {2.75,3,3.25} { \fill (120*\x+90:\r) circle (.035); };
\draw (a) -- (a\x);
\foreach \y in {0,1} {
\draw (a\x) -- (a\x\y);
\fill (a\x\y) circle (.075);
\foreach \z in {0,1} {
\draw (a\x\y) -- (a\x\y\z);
\foreach \w in {0,1} {
\draw (a\x\y\z) -- (a\x\y\z\w); };
\draw[fill=white] (a\x\y\z) circle (.075); }; };
\draw[fill=white] (a\x) circle (.075); };
\end{tikzpicture}
\end{fullfigure*}
\begin{fullfigure*}{f:a2example}{From a non-elliptic $A_2$ web, to a $\mathrm{CAT}(0)$
diskoid, to part of an affine $A_2$ building}
\begin{tikzpicture}[baseline]
\foreach \t in {0,90,180,270} {
\draw[web,midto] (\t+45:.75) -- (\t:.75);
\draw[web,midto] (\t-45:.75) -- (\t:.75);
\draw[web,midto] (\t:1.3) -- (\t:.75);
\draw[web,midto] (\t+45:.75) -- (\t+45:1.3); };
\draw[web,midto] (30:1.6) -- (45:1.3);
\draw[web,midto] (60:1.6) -- (45:1.3);
\draw[black] plot[smooth cycle] coordinates{(0:1.3) (30:1.6) (45:1.7)
(60:1.6) (90:1.3) (112.5:1.3) (135:1.3) (157.5:1.3) (180:1.3) (202.5:1.3)
(225:1.3) (247.5:1.3) (270:1.3) (292.5:1.3) (315:1.3) (337.5:1.3)};
\end{tikzpicture}
{\large $\longrightarrow$}
\begin{tikzpicture}[baseline]
\draw[fill=lightgray] (22.5:1) -- (45:1.6) -- (67.5:1);
\foreach \t in {0,45,...,315} {
\fill[lightgray] (0,0) -- (\t+22.5:1) -- (\t-22.5:1); };
\foreach \t in {0,45,...,315} {
\draw (0,0) -- (\t+22.5:1) -- (\t-22.5:1); };
\fill[blue] (45:1.6) circle (.07);
\fill[blue] (0,0) circle (.07);
\foreach \t in {0,90,180,270} {
\fill[red] (\t-22.5:1) circle (.07);
\fill[darkgreen] (\t+22.5:1) circle (.07); };
\end{tikzpicture}
{\large $\longrightarrow$}
\begin{tikzpicture}[baseline]
\coordinate (p31) at (1.861,0.589); \coordinate (p27) at (-0.433,0.422);
\coordinate (p05) at (1.926,0.997); \coordinate (p12) at (1.974,-1.439);
\coordinate (p09) at (-1.926,1.514); \coordinate (p28) at (0.816,-0.892);
\coordinate (p17) at (1.315,1.641); \coordinate (p21) at (-0.448,-2.385);
\coordinate (p10) at (1.131,-0.561); \coordinate (p15) at (-1.974,-1.073);
\coordinate (p25) at (0.448,-1.274); \coordinate (p08) at (-1.592,0.870);
\coordinate (p01) at (0.000,0.000); \coordinate (p33) at (-1.861,-0.756);
\coordinate (p14) at (-2.309,-0.429); \coordinate (p19) at (0.000,-1.450);
\coordinate (p20) at (-1.362,-1.716); \coordinate (p18) at (-0.269,1.254);
\coordinate (p04) at (1.592,-0.703); \coordinate (p32) at (-0.612,-2.070);
\coordinate (p23) at (0.269,2.406); \coordinate (p13) at (-1.131,-1.113);
\coordinate (p29) at (0.414,0.653); \coordinate (p06) at (-0.653,1.315);
\coordinate (p35) at (-0.816,1.059); \coordinate (p22) at (-1.315,2.019);
\coordinate (p30) at (0.612,-0.442); \coordinate (p16) at (0.000,1.450);
\coordinate (p02) at (0.653,0.359); \coordinate (p11) at (2.309,0.261);
\coordinate (p24) at (1.362,-1.944); \coordinate (p26) at (-0.414,-0.653);
\coordinate (p07) at (-1.545,0.175); \coordinate (p03) at (1.545,-0.175);
\coordinate (p34) at (0.433,2.090);
\fill[lightgray,nearly opaque] (p27) -- (p02) -- (p26);
\draw[semithick] (p27) -- (p02) -- (p26) -- cycle;
\fill[blue] (p27) circle (.07);
\fill[lightgray,nearly opaque] (p28) -- (p02) -- (p26);
\draw[semithick] (p28) -- (p02) -- (p26) -- cycle;
\fill[blue] (p28) circle (.07);
\fill[lightgray,nearly opaque] (p18) -- (p02) -- (p16);
\draw[semithick] (p18) -- (p02) -- (p16) -- cycle;
\fill[blue] (p18) circle (.07);
\fill[lightgray,nearly opaque] (p04) -- (p02) -- (p03);
\draw[semithick] (p04) -- (p02) -- (p03) -- cycle;
\fill[blue] (p04) circle (.07);
\fill[lightgray,nearly opaque] (p17) -- (p02) -- (p16);
\draw[semithick] (p17) -- (p02) -- (p16) -- cycle;
\fill[blue] (p17) circle (.07);
\fill[lightgray,nearly opaque] (p33) -- (p13) -- (p26);
\draw[semithick] (p33) -- (p13) -- (p26) -- cycle;
\fill[blue] (p33) circle (.07);
\fill[lightgray,nearly opaque] (p05) -- (p02) -- (p03);
\draw[semithick] (p05) -- (p02) -- (p03) -- cycle;
\fill[blue] (p05) circle (.07);
\fill[lightgray,nearly opaque] (p01) -- (p02) -- (p26);
\draw[semithick] (p01) -- (p02) -- (p26) -- cycle;
\fill[lightgray,nearly opaque] (p32) -- (p13) -- (p26);
\draw[semithick] (p32) -- (p13) -- (p26) -- cycle;
\fill[blue] (p32) circle (.07);
\fill[lightgray,nearly opaque] (p01) -- (p02) -- (p16);
\draw[semithick] (p01) -- (p02) -- (p16) -- cycle;
\fill[lightgray,nearly opaque] (p01) -- (p02) -- (p03);
\draw[semithick] (p01) -- (p02) -- (p03) -- cycle;
\fill[darkgreen] (p02) circle (.07);
\fill[lightgray,nearly opaque] (p01) -- (p13) -- (p26);
\draw[semithick] (p01) -- (p13) -- (p26) -- cycle;
\fill[red] (p26) circle (.07);
\fill[lightgray,nearly opaque] (p14) -- (p13) -- (p07);
\draw[semithick] (p14) -- (p13) -- (p07) -- cycle;
\fill[blue] (p14) circle (.07);
\fill[lightgray,nearly opaque] (p22) -- (p06) -- (p16);
\draw[semithick] (p22) -- (p06) -- (p16) -- cycle;
\fill[blue] (p22) circle (.07);
\fill[lightgray,nearly opaque] (p01) -- (p06) -- (p16);
\draw[semithick] (p01) -- (p06) -- (p16) -- cycle;
\fill[lightgray,nearly opaque] (p21) -- (p13) -- (p19);
\draw[semithick] (p21) -- (p13) -- (p19) -- cycle;
\fill[blue] (p21) circle (.07);
\fill[lightgray,nearly opaque] (p01) -- (p13) -- (p07);
\draw[semithick] (p01) -- (p13) -- (p07) -- cycle;
\fill[lightgray,nearly opaque] (p23) -- (p06) -- (p16);
\draw[semithick] (p23) -- (p06) -- (p16) -- cycle;
\fill[red] (p16) circle (.07); \fill[blue] (p23) circle (.07);
\fill[lightgray,nearly opaque] (p01) -- (p13) -- (p19);
\draw[semithick] (p01) -- (p13) -- (p19) -- cycle;
\fill[lightgray,nearly opaque] (p01) -- (p10) -- (p03);
\draw[semithick] (p01) -- (p10) -- (p03) -- cycle;
\fill[lightgray,nearly opaque] (p12) -- (p10) -- (p03);
\draw[semithick] (p12) -- (p10) -- (p03) -- cycle;
\fill[blue] (p12) circle (.07);
\fill[lightgray,nearly opaque] (p09) -- (p06) -- (p07);
\draw[semithick] (p09) -- (p06) -- (p07) -- cycle;
\fill[blue] (p09) circle (.07);
\fill[lightgray,nearly opaque] (p01) -- (p06) -- (p07);
\draw[semithick] (p01) -- (p06) -- (p07) -- cycle;
\fill[lightgray,nearly opaque] (p15) -- (p13) -- (p07);
\draw[semithick] (p15) -- (p13) -- (p07) -- cycle;
\fill[blue] (p15) circle (.07);
\fill[lightgray,nearly opaque] (p11) -- (p10) -- (p03);
\draw[semithick] (p11) -- (p10) -- (p03) -- cycle;
\fill[red] (p03) circle (.07); \fill[blue] (p11) circle (.07);
\fill[lightgray,nearly opaque] (p20) -- (p13) -- (p19);
\draw[semithick] (p20) -- (p13) -- (p19) -- cycle;
\fill[darkgreen] (p13) circle (.07); \fill[blue] (p20) circle (.07);
\fill[lightgray,nearly opaque] (p01) -- (p10) -- (p19);
\draw[semithick] (p01) -- (p10) -- (p19) -- cycle;
\fill[lightgray,nearly opaque] (p01) -- (p06) -- (p29);
\draw[semithick] (p01) -- (p06) -- (p29) -- cycle;
\fill[lightgray,nearly opaque] (p24) -- (p10) -- (p19);
\draw[semithick] (p24) -- (p10) -- (p19) -- cycle;
\fill[blue] (p24) circle (.07);
\fill[lightgray,nearly opaque] (p01) -- (p10) -- (p29);
\draw[semithick] (p01) -- (p10) -- (p29) -- cycle;
\fill[blue] (p01) circle (.07);
\fill[lightgray,nearly opaque] (p08) -- (p06) -- (p07);
\draw[semithick] (p08) -- (p06) -- (p07) -- cycle;
\fill[red] (p07) circle (.07); \fill[blue] (p08) circle (.07);
\fill[lightgray,nearly opaque] (p34) -- (p06) -- (p29);
\draw[semithick] (p34) -- (p06) -- (p29) -- cycle;
\fill[blue] (p34) circle (.07);
\fill[lightgray,nearly opaque] (p25) -- (p10) -- (p19);
\draw[semithick] (p25) -- (p10) -- (p19) -- cycle;
\fill[red] (p19) circle (.07); \fill[blue] (p25) circle (.07);
\fill[lightgray,nearly opaque] (p31) -- (p10) -- (p29);
\draw[semithick] (p31) -- (p10) -- (p29) -- cycle;
\fill[blue] (p31) circle (.07);
\fill[lightgray,nearly opaque] (p35) -- (p06) -- (p29);
\draw[semithick] (p35) -- (p06) -- (p29) -- cycle;
\fill[darkgreen] (p06) circle (.07); \fill[blue] (p35) circle (.07);
\fill[lightgray,nearly opaque] (p30) -- (p10) -- (p29);
\draw[semithick] (p30) -- (p10) -- (p29) -- cycle;
\fill[darkgreen] (p10) circle (.07); \fill[red] (p29) circle (.07);
\fill[blue] (p30) circle (.07);
\foreach \th in {90,210,330} {
\foreach \r in {2.8,3,3.2} { \fill (\th:\r) circle (.03); }; };
\end{tikzpicture}
\end{fullfigure*}
Our other main results are a generalization of this fact to $G = \mathrm{SL}(3)$.
In this case, $\mathrm{Gr}$ is again the vertex set of $\Delta$. If $w$ is a
non-elliptic web with boundary ${\vec{\lambda}}$, then $Q(D(w))$ is again the space
of colored, based simplicial maps $f:D \to \Delta$, as in \fig{f:a2example}.
Then:
\begin{theorem} Let $G = \mathrm{SL}(3) = A_2$ and let $w$ be a non-elliptic web
with minuscule boundary ${\vec{\lambda}}$ and dual diskoid $D$. Then the global
isometry configuration space $Q_g(D)$ is mapped isomorphically by $\pi$
to a dense subset of a component of the Satake fibre $F({\vec{\lambda}})$.
This inclusion yields a bijection between non-elliptic webs and the
components of $F({\vec{\lambda}})$.
\label{th:main} \end{theorem}
Our construction can be viewed as an explanation of why basis webs
are non-elliptic. A web is non-elliptic if and only if its diskoid is
$\mathrm{CAT}(0)$, essentially by definition. It is well-known that every affine
buildings is a $\mathrm{CAT}(0)$ space \cite{BT:local1}. Moreover, every convex
subset of a $\mathrm{CAT}(0)$ space, such as a diskoid which is isometrically
embedded in a building, is necessarily $\mathrm{CAT}(0)$. We will also show
that the image of each diskoid embedding $f:D \to \Delta$ in $Q_g(D)$
has a least area property. Likewise, the elliptic relations of the $A_2$
spider can be viewed as area-decreasing transformations.
Meanwhile, if $w$ is non-elliptic, then $Q(D)$ is sometimes the closure
of $Q_g(D)$ and hence maps to a single component of $F({\vec{\lambda}})$.
Eventually, $Q(D)$ has other components and maps to more than one
component of $F({\vec{\lambda}})$. These other components seem related to the
phenomenon that web bases are not dual canonical. However, we can get
an upper triangularity result as follows. In \sec{s:cyclic},
we will define a partial order $\le_S$ on the set of non-elliptic webs
using pairwise distances between boundary vertices of their dual diskoids.
\begin{theorem} The change of basis in $\mathrm{Inv}(V({\vec{\lambda}}))$ from non-elliptic
webs to the Satake basis is unitriangular, relative to the partial order
$\le_S$.
\label{th:unitriang} \end{theorem}
We have learned from Sergei Ivanov \cite{Ivanov:personal} that the partial
order in \thm{th:unitriang} refines the partial order on webs given
by the number of vertices.
Also, in \sec{s:notsatake}, we will show that the web basis, the Satake
basis, and the dual canonical basis for $\mathrm{SL}(3)$ are all eventually
different.
Finally, in \sec{s:euler}, we will propose a different formulation of the
geometric Satake correspondence based on convolution of constructible
functions rather than convolution of homology classes. (In
\thm{th:convperv}, we reinterpret geometric Satake in terms of convolution
in homology). We will prove this conjecture in the case of a tensor
product of minuscule representations of $\mathrm{SL}(3)$.
\subsection{Satake fibres and Springer fibres}
When $G = \mathrm{SL}(m)$ and ${\vec{\lambda}} = (\omega_1, \dots, \omega_1)$ is an $n =
mk$ tuple consisting of $\omega_1$ (the highest weight of the standard
representation), then $F({\vec{\lambda}})$ is isomorphic to the $(k, k, \dots,
k)$ Springer fibre. In other words, $F({\vec{\lambda}})$ is the variety of
flags in $\mathbb{C}^n$ invariant under a nilpotent endomorphism with $m$
Jordan blocks all of size $k \times k$. We have already mentioned the
well-known description of the components of the Springer or Satake fibre
in terms of planar matchings when $m=2$. This Springer fibre formalism
and this description of it have been used as a model of Khovanov homology
\cite{Khovanov:springer,Stroppel:springer}. One motivation for the present
work is to generalize this result to case $m = 3$ and obtain a description
of the components of the Springer or Satake fibre using non-elliptic webs.
\thm{th:main} accomplishes this task. (See also the end of the introduction
of \cite{Tymoczko:bijection}.)
\begin{fullfigure*}{f:joke}{Spiders and buildings}
\frame{\includegraphics[height=3in]{spiderman.jpg}}
\end{fullfigure*}
\acknowledgments
The authors would like to thank Dave Anderson, Charles Frohman, Dennis
Gaitsgory, Andr\'e Henriques, Misha Kapovich, Anthony Licata, John Millson,
Scott Morrison, Hiraku Nakajima, Alistair Savage, Petra Schwer, and Juliana
Tymoczko for useful discussions.
Another motivation for our work is shown in \fig{f:joke}.
\section{Spiders}
\label{s:spiders}
\subsection{Pivotal and symmetric categories}
\label{s:pivotal}
The definitions used in this section are nicely summarized in a survey by
Selinger \cite{Selinger:survey}; they are originally due to Freyd-Yetter
\cite{FY:braided} and Joyal-Street \cite{JS:braided}.
A \emph{pivotal category} $\mathcal{C}$ is a (strict) monoidal tensor category
such that each object $A$ has a two-sided dual object $A^*$. This means
that there is a contravariant functor $F(A) = A^*$ from $\mathcal{C}$ to itself
which is also an order-reversing tensor functor, \emph{i.e.},
$$(A \tensor B)^* = B^* \tensor A^*,$$
and which has these extra properties:
For each object $A$, there are ``cup" and ``cap" morphisms
$$b_A:I \longrightarrow A^* \tensor A \qquad d_A:A \tensor A^* \longrightarrow I$$
where $I$ denotes the unit object, such that
$$(1_A \tensor d_A)(b_A \tensor 1_A) = 1_A \qquad
(d_{A_*} \tensor 1_{A_*})(1_{A_*} \tensor b_{A_*}) = 1_{A^*}.$$
In addition, $*$ is an anti-involution of the category $\mathcal{C}$. (We assume
that $*$ is a strict involution of $\mathcal{C}$ that reverses both tensor products
and compositions of morphisms.) The axiom can be graphically summarized
as follows:
\eq{e:plumbing}{
\begin{tikzpicture}[scale=.75,baseline=-.5ex,allow upside down]
\draw[midto,web] (-1,-1) -- (-1,0);
\draw[web] (-1,0) arc (180:0:.5) arc (180:360:.5);
\draw[midto,web] (1,0) -- (1,1);
\node[anchor=south] at (-.5,.5) {$d_A$};
\node[anchor=north] at (.5,-.5) {$b_A$};
\node[anchor=east] at (-1,-.5) {$A$};
\draw[dashed] (-1.5,0) -- (1.5,0);
\end{tikzpicture} \;=\;
\begin{tikzpicture}[scale=.75,baseline=-.5ex]
\draw[web,midto] (1,-1) -- (1,0);
\draw[web] (1,0) arc (0:180:.5) arc (360:180:.5);
\draw[web,midto] (-1,0) -- (-1,1);
\node[anchor=south] at (.5,.5) {$d_{A^*}$};
\node[anchor=north] at (-.5,-.5) {$b_{A^*}$};
\node[anchor=west] at (1,-.5) {$A$};
\draw[dashed] (-1.5,0) -- (1.5,0);
\end{tikzpicture} \;=\;
\begin{tikzpicture}[scale=.75,baseline=-.5ex]
\draw[web,midto] (0,-1) -- (0,1);
\node[anchor=east] at (0,0) {$A$};
\end{tikzpicture}\;.}
A \emph{pivotal functor} is a tensor functor that preserves the above
structure.
Every object $A$ in a monoidal category has an \emph{invariant space}
$$\mathrm{Inv}(A) \stackrel{\mathrm{def}}{=} \mathrm{Hom}(I,A).$$
If the category is pivotal, then each invariant space has two other
important properties. First, every space of morphisms is an invariant
space by the relation
$$\mathrm{Hom}(A,B) \cong \mathrm{Inv}(A^* \tensor B).$$
Second, there is a cyclic action on the invariant spaces in tensor products
$$R:\mathrm{Inv}(A \tensor B) \stackrel{\cong}{\longrightarrow} \mathrm{Inv}(B \tensor A),$$
which we call a \emph{rotation map}. It extends to a rotation
of $n$ tensor factors:
$$R:\mathrm{Inv}(A_1 \tensor \cdots \tensor A_n) \stackrel{\cong}{\longrightarrow}
\mathrm{Inv}(A_2 \tensor \cdots \tensor A_n \tensor A_1).$$
Another way to describe a pivotal category, already suggested in equation
\eqref{e:plumbing}, is that it has the structure to evaluate a planar
graph $w$ drawn in a disk, if the edges of $w$ are oriented and labelled
by objects and the vertices are labelled by invariants. (The literature
uses the words ``labelled" and ``colored" interchangeably here; Selinger
\cite{Selinger:survey} calls an allowed set of colors a ``signature".)
The value of such a graph $w$ is another invariant, taking values in the
invariant space of the boundary of $w$. The graph is considered up to
isotopy rel boundary, and an edge labelled by $A$ is equivalent to the
opposite edge labelled by $A^*$. It is possible to write axioms for a
pivotal category using invariants and planar graphs rather than morphisms.
From this viewpoint, a word in a pivotal category is such a graph and it
can be called a \emph{web}.
A web is a special case of a \emph{ribbon graph} \cite{RT:ribbon}, the
difference being that a ribbon graph can also have crossings.
A \emph{braided category} is a monoidal category with crossing
isomorphisms
$$c_{A,B}:A \tensor B \to B \tensor A$$
that satisfy suitable axioms so that, among other things, a braid group acts
on the invariant space of a tensor product. If the crossing isomorphisms
are involutions, then the braid group action descends to a symmetric group
action and the monoidal category is called \emph{symmetric}. If a category
is both symmetric and pivotal, then there is an important compatibility
condition that together makes it a \emph{compact closed category}.
We require that the two involutions on $\mathrm{Inv}(A \tensor A)$, one coming from
the pivotal structure and the other from the symmetric structure, agree.
Equivalently, we require that
$$c_{A^*,A}(b_A) = b_{A^*}.$$
In a compact closed symmetric category, abstract graphs $w$ can be evaluated
whether or not they are planar.
Two other intermediate types of categories between pivotal and compact closed
are \emph{ribbon categories} and \emph{spherical categories}. A spherical
category is a pivotal category with the extra property that left traces
equal right traces, which allows the evaluation of a graph $w$ embedded in
the sphere rather than in the plane. A ribbon category is both pivotal and
braided in a compatible way, and allows the evaluation of a framed graph $w$
in $\mathbb{R}^3$. We will only need the pivotal category axioms in this article,
but all categories considered are actually ribbon or compact closed.
\subsection{Sign conventions}
\label{s:signs}
In many cases a pivotal category $\mathcal{C}$ which is linear over a field can
be modified to a different category $\mathcal{C}'$. We will be interested in two
modifications: Sign changes to the pivotal structure of $\mathcal{C}$ that do
not affect its tensor structure, and sign changes to the tensor structure
of $\mathcal{C}$. We want to restrict attention to those sign changes that allow
us to say that $\mathcal{C}$ and $\mathcal{C}'$ have the same algebraic information. For
simplicity, when discussing signs, we assume that $\mathcal{C}$ is abelian-linear
over an algebraically closed field $k$ not of characteristic 2, and
semisimple with irreducible trivial object.
Another objective of this section is to correctly interpreted labelled
graphs $w$ in a pivotal category with unoriented edges. Some edges in
a labelled graph $w$ for a pivotal or compact closed category can be
unoriented. Suppose that $A \cong A^*$ is self-dual, and suppose further
that the isomorphism $\phi_A \in \mathrm{Hom}(A,A^*)$ is cyclically invariant if
interpreted as an element of $\mathrm{Inv}(A \tensor A)$. In this case we say
that $A$ is \emph{symmetrically self-dual}. (This definition
does not require any linearity assumption.) Then an unoriented edge can
be defined by a replacement:
\eq{e:tag}{\begin{tikzpicture}[baseline=-.5ex]
\draw[web] (-.75,0) -- (.75,0) node[black,above,midway] {$A$};
\end{tikzpicture}\; \stackrel{\mathrm{def}}{=}\;
\begin{tikzpicture}[baseline=-.5ex]
\draw[web,midto] (-1,0) -- (0,0) node[black,above,midway] {$A$};
\draw[web,midto] (1,0) -- (0,0) node[black,above,midway] {$A$};
\fill[web] (0,0) circle (.07);
\end{tikzpicture},}
where the dot on the right side represents $\phi_A$. Algebraically, if
(and only if) $A$ is symmetrically self-dual, then $\mathcal{C}$ is equivalent
to a pivotal category in which $A = A^*$ outright, and $b_A = b_{A^*}$
and $d_A = d_{A^*}$. If every self-dual object in $\mathcal{C}$ is symmetrically
self-dual, then $\mathcal{C}$ is called \emph{unimodal} \cite{Turaev:quantum}.
If $A \cong A^*$ but $A$ is not symmetrically self-dual, then only the
right side of \eqref{e:tag} makes sense, and only if it is altered in
some way to break symmetry; Morrison denotes such a morphism by a ``tag"
\cite{Morrison:diagram}.
Suppose instead that $A \cong A^*$ but $A$ is not symmetrically self-dual,
and suppose that $\mathcal{C}$ is $k$-linear and semisimple and $A$ is irreducible.
Then by Schur's lemma, $\mathrm{Hom}(A,A^*)$ is 1-dimensional and rotation $R$
is multiplication by $-1$. In this case, $A$ is \emph{anti-symmetrically
self-dual}. Thus we can ask whether we can make $\mathcal{C}$ unimodal by
changing signs. This is what happens in our case (see \sec{s:examples}),
but there are also examples (namely, representation categories of finite
groups) that are not unimodal for any pivotal structure.
To understand the allowed sign changes to the pivotal structure of $\mathcal{C}$,
we first assume by category equivalence that $A$ and $A^*$ are different
objects for every $A$. Then by \eqref{e:plumbing}, we can negate $b_A$ and
$d_A$ for some irreducible $A$, without changing $b_{A^*}$ and $d_{A^*}$.
This yields a new pivotal category $\mathcal{C}'$, provided that the sign change
function $s(A)$ satisfies
$$s(A) = s(B)s(C)$$
whenever $\mathrm{Hom}(A,B \tensor C) \ne 0$. If $A$ is self-dual
and $s(A) = -1$, then this modification changes the sign
of the self-duality of $A$. It also negates the dimension of $A$,
by definition
$$\dim(A) = d_{A^*} \circ b_A.$$
Finally, since we are changing the pivotal structure by signs rather than
by other phases, $\mathcal{C}$ is spherical if only if $\mathcal{C}'$ is spherical.
We can change the sign of the tensor structure of $\mathcal{C}$ by a similar
but more complicated construction. We can assume, after passing to an
equivalent category, that the objects $\mathcal{C}$ are a free polynomial semiring
over the irreducible objects of $\mathcal{C}$ with respect to the operations $\oplus$
and $\tensor$. If
$$A = A_1 \tensor A_2 \tensor \cdots \tensor A_a$$
is a tensor product of irreducibles, and likewise $B$, $C$, and $D$
are also tensor products of irreducibles, then we can
change the sign of the tensor product map
$$\tensor:\mathrm{Hom}(A,B) \tensor \mathrm{Hom}(C,D) \to \mathrm{Hom}(A \tensor C,B \tensor D)$$
by some sign function $s(A,B;C,D) \in \{\pm 1\}$, defined when
$\mathrm{Hom}(A,B)$ and $\mathrm{Hom}(C,D)$ are both nonzero. In order for the result
$\mathcal{C}'$ to be another pivotal category, we need to check that compositions
and tensor products of morphisms are still both associative. In other
words, we need to check the equations
\begin{align*}
s(A,C;D,F) &= s(A,B;D,E)s(B,C;E,F) \\
s(A,B;C,D)s(A \tensor C,B \tensor D;E,F)
&= s(A,B;C \tensor E,D \tensor F)s(C,D;E,F)
\end{align*}
when the right sides are defined. It turns out that if $\mathcal{C}$
is pivotal or spherical, then the new tensor category $\mathcal{C}'$ can also be
made pivotal or spherical.
\subsection{Examples}
\label{s:examples}
A fundamental example of a pivotal category, indeed a compact closed
category, is the category $\mathrm{\mathbf{vect}}(k)$ of finite-dimensional vector spaces
over a field $k$. In this example, a web can be interpreted as the graph
of a tensor calculus expression (or a ``spin network"). For example,
if $\epsilon_{abc}$ is a trilinear determinant form on a 3-dimensional vector
space $V$, and if $\epsilon^{abc}$ is the dual form on $V^*$, then the tensor
$\epsilon_{abc} \epsilon^{cde}$ (with repeated indices summed) can be drawn as
$$\begin{tikzpicture}[baseline=-.5ex,scale=.65]
\draw[midto,web] (-.5,.866) -- (0,0);
\draw[midfrom,web] (0,0) -- (1,0);
\draw[midto,web] (1,0) -- (1.5,.866);
\draw[midto,web] (-.5,-.866) -- (0,0);
\draw[midto,web] (1,0) -- (1.5,-.866);
\node[anchor=south east] at (-.5,.866) {$a$};
\node[anchor=north east] at (-.5,-.866) {$b$};
\node[anchor=south west] at (1.5,.866) {$d$};
\node[anchor=north west] at (1.5,-.866) {$e$};
\end{tikzpicture},$$
with the convention in this case that the vertex labels can be inferred
from context. If the characteristic of $k$ is not 2, then another
fundamental example is the category $\mathrm{\mathbf{svect}}(k)$ of finite-dimensional
\emph{super vector spaces}, which are $\mathbb{Z}/2$-graded vector spaces with
a non-trivial symmetric and pivotal structure. Namely, if $v \in V$
and $w \in W$ are homogeneous elements of super vector spaces, then
$$c_{V,W}(v \tensor w) = (-1)^{(\deg v)(\deg w)} w \tensor v.$$
If $v \in V$ and $w \in V^*$ are homogeneous, then the cap $d_V$ is likewise
adjusted so that
$$d_V(v \tensor w) = (-1)^{(deg v)} w(v).$$
If $G$ is a group (or a Lie group, Lie algebra, or algebraic
group), then $\mathrm{\mathbf{rep}}(G,k)$, the category of finite-dimensional representations
(or continuous or algebraic representations) over $k$ is a pivotal category
with a pivotal functor to $\mathrm{\mathbf{vect}}(k)$. For the remainder of the article,
we let $G$ be a simple, simply connected algebraic group over $\mathbb{C}$ (and
later we will specialize to $G = \mathrm{SL}(3)$). We will study the pivotal
category $\mathrm{\mathbf{rep}}(G) = \mathrm{\mathbf{rep}}(G,\mathbb{C})$.
There is a deformation $\mathrm{\mathbf{rep}}_q(G)$ of $\mathrm{\mathbf{rep}}(G) = \mathrm{\mathbf{rep}}_1(G)$ that consists
of representations of the quantum group $U_q(\mathfrak{g})$, when the parameter $q$
is not a root of unity. (The deformation also exists when $q$ is a root of
unity, but there is more than one standard choice for it.) This deformation
is also a pivotal category, although it has no pivotal functor to $\mathrm{\mathbf{vect}}$,
because the cup and cap morphisms deform. Even though many ideas in this
article are clearly related to quantum representations, we will concentrate
on $\mathrm{\mathbf{rep}}(G)$, except in \sec{s:euler} when $\mathrm{\mathbf{rep}}_{-1}(G)$ will
also appear.
We are interested in two other variations of $\mathrm{\mathbf{rep}}(G)$. First, we want
to change its pivotal structure to make it unimodal. Recall that the
irreducible representations $V(\lambda)$ of $G$ are labelled by the set of
dominant weights. For a dominant weight $\lambda$, we write $\lambda^*$ for
the dominant weight such that $V(\lambda)^* \cong V(\lambda^*)$. We also
write $\rho$ for the Weyl vector and $\rho^\vee$ for the dual Weyl vector.
We make each $V(\lambda)$ a super vector space by giving it the grading
$\braket{2\lambda,\rho^\vee} \bmod 2$. In this way we realize $\mathrm{\mathbf{rep}}(G)$ as a
subcategory of $\mathrm{\mathbf{svect}}(\mathbb{C})$ with a different pivotal and symmetric structure,
and we call this version $\mathrm{\mathbf{rep}}^u(G)$. Likewise, it has a unimodal pivotal
deformation $\mathrm{\mathbf{rep}}^u_q(G)$.
Following \sec{s:signs}, $\mathrm{\mathbf{rep}}(G)$ and $\mathrm{\mathbf{rep}}_{-1}(G)$ differ only by sign
rules and all of them have equivalent information. To obtain $\mathrm{\mathbf{rep}}_{-1}(G)$
from $\mathrm{\mathbf{rep}}(G)$ in this fashion, we use the abbreviations
\begin{align*} {\vec{\lambda}} &= (\lambda_1,\lambda_2,\ldots,\lambda_n) \\
V({\vec{\lambda}}) &= V(\lambda_1) \tensor \cdots \tensor V(\lambda_n) \\
\lambda &= \sum_i \lambda_i.
\end{align*}
Then we define the sign rule
$$s(V({\vec{\lambda}}),V({\vec{\gamma}});V({\vec{\mu}}),V({\vec{\nu}})) =
(-1)^{\braket{2\lambda,\rho^\vee}\braket{\mu-\nu,\rho^\vee}}.$$
This sign rule takes $\mathrm{\mathbf{rep}}(G)$ to $\mathrm{\mathbf{rep}}_{-1}(G)$ and $\mathrm{\mathbf{rep}}^u(G)$ to
$\mathrm{\mathbf{rep}}_{-1}^u(G)$.
The other variation is a restriction to minuscule representations.
Recall that a dominant weight $\lambda$ is called \emph{minuscule} if
$\braket{\alpha^\vee, \lambda} \le 1$ for every positive coroot $\alpha^\vee$.
If $\lambda$ is a minuscule dominant weight, then $V(\lambda)$ is called a
\emph{minuscule representation}. These representations have the special
property that all of their weights are in the Weyl orbit of the highest
weight. We define $\mathrm{\mathbf{rep}}(G)_\min$ to be the monoidal subcategory of $\mathrm{\mathbf{rep}}(G)$
generated by minuscule representations. So the objects of $\mathrm{\mathbf{rep}}(G)_\min$
are tensor products of minuscule representations. It is a symmetric
category which is neither an additive nor an abelian category. If there
exists a minuscule $\lambda$ such that $\braket{2\lambda, \rho^\vee}$ is odd,
then $\mathrm{\mathbf{rep}}(G)_\min$ is also not a pivotal category, because it is skeletal
and yet has objects which are anti-symmetrically self-dual in $\mathrm{\mathbf{rep}}(G)$.
However, $\mathrm{\mathbf{rep}}^u(G)_\min$ is a well-defined pivotal category in which $*$
is a strict involution and $V(\lambda^*) = V(\lambda)^*$.
In the case $G = \mathrm{SL}(n)$, and in some other cases, $\mathrm{\mathbf{rep}}(G)$ can be recovered
as the Karoubi envelope of $\mathrm{\mathbf{rep}}(G)_\min$, although we will not use this
construction in this article.
The other main pivotal category which we will study in this paper is the
category of ${G^\vee}(\mathcal{O})$-equivariant perverse sheaves $\mathrm{\mathbf{perv}}(\mathrm{Gr})$ on $\mathrm{Gr}$.
This category has a relatively straightforward pivotal structure. It also
has a more delicate symmetric structure which is called a ``commutativity
constraint'' or ``braiding'', as defined by Ginzburg \cite{Ginzburg:loop}
and Mirkovi\'c-Vilonen \cite{MV:geometric} in two different ways.
(See also \cite[Sec. 5.3.8]{BD:hitchin}.)
\thm{th:satake} states that $\mathrm{\mathbf{perv}}(G)$ is equivalent to $\mathrm{\mathbf{rep}}^u(G)$ both as a
pivotal category and a symmetric category; we will be more interested in
the pivotal structure. We also will be more interested in the minuscule
analog of $\mathrm{\mathbf{perv}}(\mathrm{Gr})$, which we will analyze in \sec{s:conv}.
\subsection{Free spiders and presentations}
Pivotal categories can also be presented by generators and relations.
If the pivotal category is additive-linear over a ring or a field, then
it can presented in the same sense, using linear combinations of words in
the generators. In general there are generating objects (or edges) and
generating morphisms (or invariants or vertices), while the relations are
all morphisms. Relations in a pivotal category are also known as
\emph{planar skein relations}.
We now define the free spider $\mathrm{\mathbf{fsp}}(G)$ to be the free $\mathbb{C}$-linear pivotal
category generated by an edge for each minuscule representation of $G$
and a vertex for every triple $\lambda, \mu, \nu$ of minuscule dominant
weights such that
$$\mathrm{Inv}_G(V(\lambda, \mu, \nu)) \ne 0.$$
Note that the minuscule condition
forces this vector space to be at most one-dimensional. In $\mathrm{\mathbf{fsp}}(G)$,
we also impose that the dual of the $\lambda$ edge is $\lambda^*$.
In \cite{Morrison:diagram}, $\mathrm{\mathbf{fsp}}(\mathrm{SL}(n))$ was denoted $\text{Sym}_n$.
A free spider has the same relationship to webs as a free group has to
words in its generators. Namely, two webs are equal in $\mathrm{\mathbf{fsp}}(G)$ if and
only if they are isotopic rel boundary. (Selinger \cite{Selinger:survey}
also defines free categories of various kinds generated by signatures.)
Let us fix $q \in \mathbb{C}$, non-zero and not a root of unity (but possibly
equal to 1). There is a pivotal functor
$$\Psi:\mathrm{\mathbf{fsp}}(G) \to \mathrm{\mathbf{rep}}^u_q(G)_\min,$$
which is defined by choosing a non-zero element in each invariant space
$$\mathrm{Inv}_{U_q(\mathfrak{g})}(V(\lambda, \mu, \nu)).$$
In particular, for each web $w$ with boundary ${\vec{\lambda}}$, we obtain
an element
$$\Psi(w) \in \mathrm{Inv}_{U_q(\mathfrak{g})}(V({\vec{\lambda}})).$$
Actually, since webs are a notation for words in any pivotal category,
we could say also say that $w$ ``is" $\Psi(w)$, or that its value is
$\Psi(w)$. But the distinction between $w$ and $\Psi(w)$ will be useful
for us. The first result is that $\Psi$ is surjective when $G = \mathrm{SL}(n)$
\cite[Prop. 3.5.8]{Morrison:diagram}. (This follows from Weyl's fundamental
theorem of invariant theory.) Thus, the vectors $\Psi(w)$ of webs $w$
span the invariant spaces.
It is an open problem to generate the kernel of $\Psi$ with planar skein
relations in $\mathrm{\mathbf{fsp}}(G)$. This problem has been solved when $G$ has rank 1
or 2 by the third author \cite{Kuperberg:spiders}. Kim \cite{Kim:thesis}
has conjectured an answer for $\mathrm{SL}(4)$ in \cite{Kim:thesis} and Morrison
\cite{Morrison:diagram} has done so for $\mathrm{SL}(n)$. Once these planar
skein relations (which must depend on $q$) are determined, then the
resulting presented pivotal category can be called a spider and we denote
it $\mathrm{\mathbf{spd}}_q(G)$.
We now review the known solutions for $\mathrm{SL}(2)$ and $\mathrm{SL}(3)$. The
\emph{Temperley-Lieb category} or \emph{$A_1$ spider} $\mathrm{\mathbf{spd}}_q(\mathrm{SL}(2))$ is
the quotient of $\mathrm{\mathbf{fsp}}(\mathrm{SL}(2))$ by the single relation
\eq{e:tl}{\begin{tikzpicture}[baseline=-.5ex,web]
\draw (0,0) circle (.4);
\end{tikzpicture} \;\;= \;\;-q-q^{-1}.}
(Since $\mathrm{SL}(2)$ has a single, self-dual minuscule representation,
$\mathrm{\mathbf{fsp}}(\mathrm{SL}(2))$ and $\mathrm{\mathbf{spd}}_q(\mathrm{SL}(2))$ have unoriented edges with a single
color or label.) The \emph{$A_2$ spider} $\mathrm{\mathbf{spd}}_q(\mathrm{SL}(3))$ is the quotient
of $\mathrm{\mathbf{fsp}}(\mathrm{SL}(3))$ by the relations
\begin{align}
\begin{tikzpicture}[baseline=-.5ex,web,allow upside down]
\draw[midto,rotate=180] (0,0) circle (.4);
\end{tikzpicture}\;\; &= \;\; q^2+1+q^{-2} \nonumber \\
\begin{tikzpicture}[baseline=-.5ex,web]
\draw[midto] (-1.25,0) -- (-.5,0);
\draw[midto] (.5,0) -- (1.25,0);
\draw (.5,0) arc (30:150:.577); \draw[midto] (0,.288) -- (-.01,.288);
\draw (.5,0) arc (330:210:.577); \draw[midto] (0,-.288) -- (-.01,-.288);
\end{tikzpicture}\;\; &= \;\;
(-q-q^{-1})\;\;
\begin{tikzpicture}[baseline=-.5ex,web]
\draw[midto] (-.5,0) -- (.5,0);
\end{tikzpicture} \label{e:a2spider} \\
\begin{tikzpicture}[baseline=-.5ex,web]
\draw (.35,.35) arc (60:120:.7) arc (150:210:.7)
arc (240:300:.7) arc (-30:30:.7);
\draw[midfrom] (0,.444) -- (.01,.444);
\draw[midfrom] (-.444,0) -- (-.444,-.01);
\draw[midfrom] (0,-.444) -- (-.01,-.444);
\draw[midfrom] (.444,0) -- (.444,.01);
\draw[midto] (.35,.35) -- (.707,.707);
\draw[midfrom] (-.35,.35) -- (-.707,.707);
\draw[midto] (-.35,-.35) -- (-.707,-.707);
\draw[midfrom] (.35,-.35) -- (.707,-.707);
\end{tikzpicture}\;\; &=\;\;
\begin{tikzpicture}[baseline=-.5ex,web]
\draw (.5,.5) arc (315:225:.707);
\draw[midto] (0,.293) -- (.01,.293);
\draw (-.5,-.5) arc (135:45:.707);
\draw[midto] (0,-.293) -- (-.01,-.293);
\end{tikzpicture}\;\; + \;\
\begin{tikzpicture}[baseline=-.5ex,web]
\draw (.5,-.5) arc (225:135:.707);
\draw[midto] (.293,0) -- (.293,.01);
\draw (-.5,.5) arc (45:-45:.707);
\draw[midto] (-.293,0) -- (-.293,-.01); \nonumber
\end{tikzpicture}\;\;.
\end{align}
(Since $\mathrm{SL}(3)$ has two minuscule representations which are dual to each
other, $\mathrm{\mathbf{fsp}}(\mathrm{SL}(3))$ and $\mathrm{\mathbf{spd}}_q(\mathrm{SL}(3))$ have oriented edges with
one label or color. By convention, the edge is labelled by the first
fundamental representation $\omega_1$ in the direction that it is oriented.)
The other two known spiders, $\mathrm{\mathbf{spd}}_q(B_2)$ and $\mathrm{\mathbf{spd}}_q(G_2)$, have similar
but more complicated presentations.
\begin{theorem}[Kauffman \cite{Kauffman:spinknot}] If $q$ is not a root
of unity, then $\mathrm{\mathbf{spd}}_q(\mathrm{SL}(2))$ is equivalent to the pivotal category
$\mathrm{\mathbf{rep}}^u_q(\mathrm{SL}(2))_\min$ of minuscule representations.
\label{th:kauffman} \end{theorem}
\begin{theorem} \cite{Kuperberg:spiders} If $q$ is not a root of unity,
then $\mathrm{\mathbf{spd}}_q(\mathrm{SL}(3))$ is equivalent to the pivotal category
$$\mathrm{\mathbf{rep}}_q(\mathrm{SL}(3))_\min = \mathrm{\mathbf{rep}}^u_q(\mathrm{SL}(3))_\min$$
of minuscule representations.
\end{theorem}
(In the case of $\mathrm{SL}(3)$, it turns out that $\mathrm{\mathbf{rep}}_q(\mathrm{SL}(3))$
and $\mathrm{\mathbf{rep}}^u_q(\mathrm{SL}(3))$ are the same; see \sec{s:examples}.)
A main property of the spider relations \eqref{e:a2spider} is that they
are confluent or Gr\"obner type. In the free pivotal category generated
by the generating edges and vertices, each web can be graded by the
number of its faces. Then each relation has exactly one leading term, an
elliptic face. (In the $A_2$ spider, a face is \emph{elliptic} if it has
fewer than six sides. In the other two rank 2 spiders, a face is elliptic
if the total angle of the corresponding dual vertex is less than $2\pi$,
so that the vertex is $\mathrm{CAT}(0)$; see \sec{s:diskoids}.) A web that has
that face can be expressed, modulo the relation, as a linear combination of
lower-degree webs. The Gr\"obner property, proved using a diamond lemma,
is that any two sequences of simplifications of the same web lead to the
same final expression. This means that the webs that cannot be simplified,
\emph{i.e.}, the webs without elliptic faces or the non-elliptic webs, form a basis
of each invariant space. There is an extended version of this result,
but we will restrict our attention to the minuscule case, summarized in
the following theorem.
\begin{theorem} \cite{Kuperberg:spiders} If ${\vec{\lambda}}$ is a sequence of
dominant minuscule weights of $\mathrm{SL}(3)$, then the non-elliptic type $A_2$
webs with boundary ${\vec{\lambda}}$ are a basis of $\mathrm{Inv}(V({\vec{\lambda}}))$.
\label{th:sl3basis} \end{theorem}
\thm{th:unitriang} implies \thm{th:sl3basis} as a corollary. However,
it is much more complicated than other proofs of \thm{th:sl3basis}
\cite{Westbury:trivalent,Kuperberg:notdual}.
\section{Affine geometry}
\label{s:ageom}
\subsection{Weight-valued metrics and linkages}
\label{s:wmetric}
In the usual definition of a metric space, distances take values in
the non-negative real numbers $\mathbb{R}_{\ge 0}$. However, Kapovich, Leeb,
Millson \cite{KLM:generalized} have a theory of metric spaces in
which distances take values in the dominant Weyl chamber of $G$. Two of
the axioms of such a generalized metric space are easy to state:
$$d(x,x) = 0 \qquad d(x,y) = d(y,x)^*.$$
The third axiom, the triangle inequality, is different. The main results
of Kapovich, Leeb, and Millson are generalized triangle inequalities that
are satisfied in buildings and generalized symmetric spaces. On the one
hand, the triangle inequalities in the $A_1$ case are the usual triangle
inequality. On the other hand, the inequalities in higher rank cases are
decidedly non-trivial.
In this article, we will adopt the viewpoint of weight-valued metric spaces
in order to discuss isometries and distance comparisons. We will not need
the generalized triangle inequalities, but we will need isometries and
distance comparisons. The definition of an isometry is straightforward.
As for distance comparisons, we will say that $\mu \le \lambda$ as a
distance if and only if $\mu \le \lambda$ in the usual partial order on
dominant weights, namely that $\lambda - \mu$ is a non-negative integer
combination of simple roots. Thus, a ball of radius $\lambda$ is then a
finite union of spheres of radius $\mu \le \lambda$. For one construction
we will define distances that take values in the dominant Weyl chamber,
instead of integral weights; and then we say that $\mu \le \lambda$ when
$\lambda - \mu$ is a non-negative real combination of simple roots.
In addition to isometries, we will be interested in partial isometries in
which only some distances are preserved. For this purpose, we define a
\emph{linkage} to be an oriented graph $\Gamma$ whose edges are labelled by
dominant weights. As with webs, an edge labelled by $\lambda$ is equivalent
to the opposite edge labelled by $\lambda^*$. Let $v(\Gamma)$ be the set of
vertices of $\Gamma$. Then one may attempt to define a distance $d(p,q)$
between any two points $p,q \in v(\Gamma)$ by taking the shortest total
distance of a connecting path. However, since weights are only partially
ordered, this minimum may not be unique. We will say that $\Gamma$ has
\emph{coherent geodesics} if the minimum distance $\min(d(p,q))$ between
any two vertices $p$ and $q$ is unique, and if that minimum distance is
the length of the edge $(p,q)$ when $\Gamma$ has that edge. In this case
$\Gamma$ can be completed to another linkage $\Gamma_g$ which is a complete
graph, using all distances as weights.
\subsection{Configuration spaces}
\label{s:config}
Let $X$ be a weight-valued metric space, and let $\Gamma$ be a linkage as
in \sec{s:wmetric}. Let $v(\Gamma)$ be the set of vertices of $\Gamma$.
Then we define the \emph{linkage configuration space} $Q(\Gamma,X)$ to be
the set of maps
$$f:v(\Gamma) \to X$$
such that $d(f(p),f(q))$ equals the weight of the edge from $p$ to
$q$, when there is such an edge. If $X$ and $\Gamma$ both have a base
point, then $Q(\Gamma,X)$ is instead the configuration space of based maps.
Another possibility is that $\Gamma$ has a base edge of length $\lambda$
and $X$ has two base points at distance $\lambda$; then $Q(\Gamma,X)$
is again the configuration space of based maps. We will be interested in
four types of linkages $\Gamma$:
\begin{description}
\item[1] A path or \emph{polyline}.
\item[2] A cycle or polygon.
\item[3] The 1-skeleton $\Gamma(D)$ of a tiled diskoid $D$ (\sec{s:diskoids})
with edges labelled by weights.
\item[4] The complete linkage $\Gamma_g(D)$, if $\Gamma(D)$ has coherent
geodesics.
\end{description}
There is one final type of configuration space that is sometimes useful.
If an edge $(p,q)$ has weight $\lambda$, then we can ask that
$$d(f(p),f(q)) \le \lambda$$
instead of
$$d(f(p),f(q)) = \lambda.$$
The result is the contractive configuration space $Q_c(\Gamma,X)$.
Suppose that $X = G/H$ for some group $G$ with a subgroup $H$, and that
each sphere $X(\lambda)$ around the base point is a double coset of $H$.
Let $\Gamma$ be a linkage and let $\Gamma_0$ be the same linkage with a
chosen base point $0$. Then there is a fibration
$$Q(\Gamma_0,X) \longrightarrow Q(\Gamma,X) \longrightarrow X.$$
Similarly, if $\Gamma_e$ denotes the same linkage with a base edge $e$
of length $\lambda$ incident to $0$, then there is also a fibration
\begin{equation} \label{eq:basededge}
Q(\Gamma_e,X) \longrightarrow Q(\Gamma_0,X) \longrightarrow X(\lambda),
\end{equation}
where $X(\lambda) = Q(\lambda,X)$ is the sphere of radius $\lambda$ around
the (first) base point of $X$, and the second base point is an arbitrary
point in $X(\lambda)$.
If $f:\Gamma_2 \to \Gamma_1$ is a map between linkages, then there
is a restriction map,
\eq{e:pi}{\pi_{\Gamma_2}^{\Gamma_1}:Q(\Gamma_1,X) \to Q(\Gamma_2,X)}
between their configuration spaces. We will be particularly interested
in this map when $\Gamma_1$ is a sublinkage of $\Gamma_2$ (for example
its boundary).
Suppose now that $\Gamma = \Gamma_1 \cup \Gamma_2$, and that $\Gamma_1
\cap \Gamma_2$ is either an edge or a vertex. If we base $\Gamma_2$
(but not $\Gamma_1$) at this intersection, then the configuration space
$Q(\Gamma,X)$ is a twisted product:
$$Q(\Gamma,X) = Q(\Gamma_1,X) \tilde{\times} Q(\Gamma_2,X).$$
Informally, $\Gamma_2$ is either an arm attached to $\Gamma_1$ at a point
which can swing freely in any direction, or a flap attached to $\Gamma_1$
along a 1-dimensional hinge which can swing freely in the remaining
directions.
\subsection{Diskoids}
\label{s:diskoids}
Recall that a \emph{piecewise-linear diskoid} is a contractible, compact,
piecewise-linear region in the plane. (We will not need diskoids that are
not piecewise-linear. But if one were to consider them, the most natural
definition could be to make it a planar, cell-like continuum.) Any diskoid
$D$ has a polygonal boundary $P$ with a boundary map $P \to D$, which however
is not an inclusion unless $D$ is either a point or a disk. \fig{f:diskoid}
shows an example of a diskoid $D$ with its boundary $P$.
\begin{fullfigure}{f:diskoid}{A diskoid $D$ with boundary $P$}
\begin{tikzpicture}[scale=.5]
\coordinate (p1) at (-2,.5); \coordinate (p2) at (-3,2);
\coordinate (p3) at (-3.5,0); \coordinate (p4) at (0,0);
\coordinate (p5) at (.5,-2); \coordinate (p6) at (2.5,-.5);
\coordinate (p7) at (2,1); \coordinate (p8) at (-1,-3);
\coordinate (p9) at (1.5,-3); \coordinate (p10) at (3,2.5);
\coordinate (p11) at (2.98,2.52);
\draw[web,double distance = .2cm]
(p1) -- (p2) -- (p3) -- (p1) -- (p4) -- (p5) -- (p8) -- (p9) -- (p5) --
(p6) -- (p7) -- (p10) -- (p11) -- (p7) -- (p4) -- cycle;
\draw[fill=lightgray]
(p1) -- (p2) -- (p3) -- (p1) -- (p4) -- (p5) -- (p8) -- (p9) -- (p5) --
(p6) -- (p7) -- (p10) -- (p7) -- (p4) -- cycle;
\node at (1.25,-.375) {$D$};
\node[anchor=south west] at (-1,.5) {$P$};
\end{tikzpicture}
\end{fullfigure}
Note that since a diskoid comes with an embedding in the plane, its boundary
$P$ is implicitly oriented, so that the edges of $P$ are cyclically ordered.
We will assume a clockwise orientation in this article. Trees are diskoids,
and \fig{f:a1example} has an example of the polygonal boundary of a tree;
the polygon traverses each edge twice.
A diskoid $D$ can be tiled by polygons. Formally, a \emph{tiling} of $D$
is a piecewise-linear CW complex structure on $D$ with embedded 2-cells.
If $D$ is decorated in this way, then we define the graph $\Gamma(D)$ to be
its 1-skeleton. Then, as above, $\Gamma(D)$ can be made into a linkage,
which means, explicitly, that the edges of $D$ are labelled by distances.
In this article we will not need to the label the faces (or 2-cells)
of a tiled diskoid to define its configuration space, but only because
the corresponding representation theory is multiplicity-free. In future
work, the faces could also be labelled in order to define more restrictive
configuration spaces. We will write $Q(D)$ for $Q(\Gamma(D))$ and $Q_g(D)$
for $Q(\Gamma_g(D))$.
In some cases, although not the most important cases, we will be interested
in diskoids with bubbles. By definition, a \emph{diskoid with bubbles} is,
inductively, either a diskoid, or a one-point union of a smaller diskoid
with bubbles and either a line segment or a piecewise linear 2-sphere.
The extra line segments and 2-spheres are not embedded in the plane and
do not affect the boundary of the diskoid, even if the attachment point is
on the boundary. The discussion of the previous paragraph applies equally
well to diskoids with bubbles.
Our interest in diskoids arises from the fact that they are geometrically
dual to webs. As in the introduction, let $w$ be a web in $\mathrm{\mathbf{fsp}}(G)$ with
boundary ${\vec{\lambda}}$. Then it has a dual diskoid $D = D(w)$, with bubbles
if $w$ has closed components, and with a natural base point. To be precise,
$D$ has a vertex for every internal or external face of $w$; two vertices
are connected by an edge when the faces of $w$ are adjacent; and there
is a triangle glued to three edges whenever the dual edges of $w$ meet at
a vertex. We label the edges of $D$ using the labels of the corresponding
edges of $w$; also, if an edge of $w$ is oriented, we transfer it to an
orientation of the dual edge of $D$ by rotating it counterclockwise. As
a result, the boundary of the diskoid $D$ is the polygon $P({\vec{\lambda}})$.
\fig{f:a1example} shows an example of an $A_1$ web and its dual diskoid,
which in the $A_1$ case is always a tree. \fig{f:a2example} shows an
example of an $A_2$ web and its dual diskoid, which happens to be a disk
because the corresponding web is connected.
In this construction, $D$ is always triangulated because $w$ is always
trivalent. The vertices of $D$ are a weight-valued metric space, and by
linear extension the whole of $D$ is a Weyl-chamber-valued metric space.
We can also simplify this metric to an ordinary metric space by taking the
Euclidean length of the vector-valued distance. Finally, suppose that
$w$ is an $A_2$ web (or a $B_2$ or $G_2$ web). Then $w$ is non-elliptic
if and only if $D$, in its ordinary metric, is $\mathrm{CAT}(0)$ in the sense
of Gromov \cite{Gromov:hyperbolic}. This follows from the fact that $D$
is contractible and the condition that all complete angles in $D$ are at
least $2\pi$.
\subsection{Affine Grassmannians and buildings}
As before, let $G$ be a simple, simply-connected complex algebraic group
and let ${G^\vee}$ be its Langlands dual group. Let $\mathcal{O} = \mathbb{C}[[t]]$ be the
ring of formal power series over $\mathbb{C}$ and let $\mathcal{K} = \mathbb{C}((t))$ be its
fraction field. Then
$$\mathrm{Gr} = \mathrm{Gr}({G^\vee}) = {G^\vee}(\mathcal{K})/{G^\vee}(\mathcal{O})$$
is the \emph{affine Grassmannian} for ${G^\vee}$ with residue field $\mathbb{C}$. It is
an ind-variety over $\mathbb{C}$, meaning that it is a direct limit of algebraic
varieties (of increasing dimension). The affine Grassmannian $\mathrm{Gr}$ is
also a weight-valued metric space: The double cosets ${G^\vee}(\mathcal{O})\backslash
{G^\vee}(\mathcal{K})/{G^\vee}(\mathcal{O})$ are bijective with the cone $\Lambda_+$ of dominant
coweights of ${G^\vee}$, which is the same as the cone of dominant weights of $G$.
More precisely, for each coweight $\mu$ of ${G^\vee}$, there is an associated
point $t^\mu$ in the affine Grassmannian. If $p,q$ are two arbitrary
points of the affine Grassmannian, then we can find $g \in {G^\vee}(\mathcal{K})$
such that $gp = t^0$ and $gq = t^\mu$ for some unique dominant
coweight $\mu$. Under this circumstance, we write $d(p,q) = \mu$.
So the action of ${G^\vee}(\mathcal{K})$ preserves distances and $d(t^0, t^\mu) =
\mu$ for any dominant weight $\mu$.
The affine Grassmannian $\mathrm{Gr}$ is also a subset of the vertices $\mathrm{Gr}' =
v(\Delta)$ of an associated simplicial complex called an \emph{affine
building} $\Delta = \Delta({G^\vee})$ \cite{Ronan:buildings} whose type
is the extended Dynkin type of ${G^\vee}$. The simplices of this affine
building are given by parahoric subgroups of the affine Kac-Moody group
$\widehat{{G^\vee}}$. For a detailed description of affine buildings from
this perspective, see \cite{GL:cycles}.
An affine building $\Delta$ satisfies the following axioms:
\begin{description}
\item[1] The building $\Delta$ is a non-disjoint union of \emph{apartments},
each of which is a copy of the Weyl alcove simplicial complex of ${G^\vee}$.
\item[2] Any two simplices of $\Delta$ of any dimension are both contained
in at least one apartment $\Sigma$.
\item[3] Given two apartments $\Sigma$ and $\Sigma'$ and two simplices
$\alpha,\alpha' \in \Sigma \cap \Sigma'$, there is an isomorphism $f:\Sigma
\to \Sigma'$ that fixes $\alpha$ and $\alpha'$ pointwise.
\end{description}
The axioms imply that the vertices of $\Delta$, denoted $\mathrm{Gr}'$, are
canonically colored by the vertices of the extended Dynkin diagram $\hat{I}
= I \sqcup \{0\}$ of ${G^\vee}$, or equivalently the vertices of the standard
Weyl alcove $\delta$ of ${G^\vee}$. Moreover, every maximal simplex of
$\Delta$ is a copy of $\delta$; it has exactly one vertex of each color.
The affine Grassmannian consists of those vertices colored by $0$ and by
minuscule nodes of the Dynkin diagram of ${G^\vee}$.
The axioms also imply that $v(\Delta)$, and more generally the realization
$|\Delta|$ of $\Delta$, have a metric taking values in Weyl chamber.
(But not necessarily integral weights as one sees in $\mathrm{Gr}$.) Namely, if
$p,q \in |\Delta|$, then $p,q \in |\Sigma|$ for an apartment $\Sigma$, and
after a suitable automorphism $p = q + \lambda$ for some vector $\lambda$ in
the dominant Weyl chamber. We then define $d(p,q) = \lambda$. (The metric
has coherent geodesics, and it extends the metric defined above for $\mathrm{Gr}$.)
We will need the following fact.
\begin{lemma} If $p,q \in |\Delta|$, then every geodesic path $\gamma$
from $p$ to $q$ is contained in every apartment $\Sigma$ such that $p,q
\in |\Sigma|$.
\end{lemma}
A subtle feature of the above affine building $\Delta$ is that it has two
very different geometries. As an ordinary simplicial complex, its vertex
set $\mathrm{Gr}'$ is discrete, and $\mathrm{Gr}'$ has a combinatorial, weight-valued metric.
The vertex set $\mathrm{Gr}'$ is also naturally an algebraic ind-variety over $\mathbb{C}$,
as is the set of vertices of any given color or the set of simplices of
$\Delta$ of any given type. This second geometry endows $\mathrm{Gr}'$ with both
a Zariski topology and an analytic topology. Among the relations between
these two geometries, we will need the following fact.
\begin{proposition} The algebraic-geometric closure $\bar{\mathrm{Gr}'(\lambda)}$
of the sphere $\mathrm{Gr}'(\lambda)$ of radius $\lambda$ is the set of all points
in the metric ball of radius $\lambda$ that have the same color as $\lambda$.
\label{p:sphereball} \end{proposition}
An affine building $\Delta$ has a third geometry which is related to
the weight-valued metric but is not the same. Namely, we can give the
Weyl alcove $\delta$ its standard Euclidean structure, and consider the
induced metric on the realization $|\Delta|$ of $\Delta$. This locally
Euclidean metric can also be defined as $||d(p,q)||_2$, where $d(p,q)$ is
the weight-valued metric on $|\Delta|$.
\begin{theorem}[Bruhat-Tits \cite{BT:local1}] Every affine building is a
$\mathrm{CAT}(0)$ space with respect to its locally Euclidean metric.
\label{th:cat0} \end{theorem}
If $G = \mathrm{SL}(n)$ and thus ${G^\vee} = \mathrm{PGL}(n)$, then $\mathrm{Gr} = \mathrm{Gr}'$, and there
is a simple description of $\Delta$. Namely, a finite set of vertices
in $\mathrm{Gr}$ subtends a simplex if and only if the distances between them are
all minuscule.
Finally, to close a circle, let $L({\vec{\lambda}})$ be a polyline whose sides
are labelled by
$${\vec{\lambda}} = (\lambda_1,\lambda_2,\dots,\lambda_n),$$
based at the beginning. Let $P({\vec{\lambda}})$ be the corresponding polygon,
based between $\lambda_n$ and $\lambda_1$. Then the contractive polyline
configuration space
$$\mathrm{Gr}({\vec{\lambda}}) = Q_c(L({\vec{\lambda}}),\mathrm{Gr})$$
is the domain of the convolution morphism. The restriction map coming
from the projection onto the boundary $L({\vec{\lambda}}) \to {\mathrm{pt}}$, or
$$\pi_{\mathrm{pt}}^{L({\vec{\lambda}})}:Q_c(L({\vec{\lambda}}),\mathrm{Gr}) \to \mathrm{Gr},$$
is the convolution morphism. In keeping with the standard notation,
we will denote it by
$$m_{\vec{\lambda}} = \pi_{\mathrm{pt}}^{L({\vec{\lambda}})}.$$
Meanwhile the contractive polygon configuration space
$$Q_c(P({\vec{\lambda}}),\mathrm{Gr}) = F({\vec{\lambda}}) = m_{\vec{\lambda}}^{-1}(t^0)$$
is the Satake fibre. As another bit of notation, if $\Gamma$ is a linkage,
we will elide the $\mathrm{Gr}$ and write $Q(\Gamma)$ for $Q(\Gamma,\mathrm{Gr})$, etc.
\section{Geometric Satake for tensor products of minuscule representations}
\label{s:minuscule}
\subsection{Minuscule paths and components of Satake fibres}
\label{s:paths}
The full geometric Satake correspondence, \thm{th:satake}, simplifies
considerably when the weights are minuscule. In this special case, Haines
\cite[Thm. 3.1]{Haines:equidim} showed that all components of $F({\vec{\lambda}})$
are of maximal dimension. We can use his ideas to give an explicit
description of these components using minuscule paths. In addition to
previous notation, let $W$ be the Weyl group of $G$.
Let $\lambda$ be a minuscule dominant weight. Then there are no dominant
weights less than $\lambda$, so the sphere of radius $\lambda$ equals
the ball of radius $\lambda$. Hence the sphere $\mathrm{Gr}(\lambda)$ is closed
in the algebraic geometry of $\mathrm{Gr}$ by \prop{p:sphereball}, and thus it is
projective and smooth. In fact, ${G^\vee}$ acts transitively on $\mathrm{Gr}(\lambda)$.
The stabilizer of $t^\lambda$ is $M(\lambda)$, the opposite maximal proper
parabolic subgroup corresponding to the minuscule weight $\lambda$. Thus
$\mathrm{Gr}(\lambda)$ is isomorphic to the partial flag variety ${G^\vee}/M(\lambda)$.
More generally, if $\Gamma$ is a \emph{minuscule linkage}, meaning that
all of its edges are minuscule, then
$$Q(\Gamma) = Q_c(\Gamma) = \bar{Q(\Gamma)}.$$
Let
$${\vec{\lambda}} = (\lambda_1, \ldots, \lambda_n)$$
be a sequence of minuscule dominant weights. A \emph{minuscule path}
(ending at 0) of type ${\vec{\lambda}}$ is a sequence of dominant weights
$${\vec{\mu}} = (\mu_0,\mu_1,\mu_2,\ldots,\mu_n)$$
such that $\mu_k - \mu_{k-1} \in W\lambda_k$ for every $k$, and such that
$$\mu_0 = \mu_n = 0.$$
In other words, the $k$th step of the path ${\vec{\mu}}$ is a weight of
$V(\lambda_k)$, and the path is restricted to the dominant Weyl chamber
$\Lambda_+$. Minuscule paths are a special case of Littelmann paths
\cite{Littelmann:paths}, but it was much earlier folklore knowledge that
the number of minuscule paths of type ${\vec{\lambda}}$ is the dimension of
$\mathrm{Inv}(V({\vec{\lambda}}))$. (See Humphreys \cite[Ex. 24.9]{Humphreys:gtm}, and
use induction.)
\begin{fullfigure}{f:fan}{The fan diskoid $A({\vec{\lambda}},{\vec{\mu}})$}
\begin{tikzpicture}[scale=.75,draw=darkred,semithick]
\draw[midto] (0,0) to node[below left=-.5ex] {$\lambda_1=\mu_1$} (-2,1);
\draw[midto] (-2,1) to node[below left] {$\lambda_2$} (-3,3);
\draw[midto] (-3,3) to node[above left] {$\lambda_3$} (-2,5);
\draw[midto] (-2,5) to node[above] {$\lambda_4$} (0,6);
\draw (0,6) -- (1,6);
\draw (3,3) -- (3,2);
\draw[midto] (3,2) to node[below right] {$\lambda_{n-1}$} (2,0);
\draw[midto] (2,0) to node[below right] {$\lambda_n = \mu^*_{n-1}$} (0,0);
\draw[midto] (0,0) to node[above right] {$\mu_2$} (-3,3);
\draw[midto] (0,0) to node[above right] {$\mu_3$} (-2,5);
\draw[midto] (0,0) to node[above right] {$\mu_4$} (0,6);
\draw[midto] (0,0) to node[above=.75ex] {$\mu_{n-2}$} (3,2);
\fill (2.14,5.04) circle (.035); \fill (2.3,4.8) circle (.035);
\fill (2.46,4.56) circle (.035);
\fill (1.04,3.04) circle (.035); \fill (1.2,2.8) circle (.035);
\fill (1.36,2.56) circle (.035);
\end{tikzpicture}
\end{fullfigure}
Given a minuscule path ${\vec{\mu}}$ of type ${\vec{\lambda}}$, we define a based diskoid
$A({\vec{\lambda}},{\vec{\mu}})$ in the shape of a fan, whose the boundary is the polygon
$P({\vec{\lambda}})$ and whose ribs are labelled by ${\vec{\mu}}$, as in \fig{f:fan}.
Then there is a natural inclusion
$$Q(A({\vec{\lambda}},{\vec{\mu}})) \subseteq F({\vec{\lambda}}).$$
The following result is implicit in the work of Haines \cite{Haines:equidim}.
\begin{theorem} For each minuscule path ${\vec{\mu}}$, the fan configuration space
$Q(A({\vec{\lambda}},{\vec{\mu}}))$ is a dense subset of one component of $F({\vec{\lambda}})$.
The induced correspondence is a bijection between minuscule paths and
components of $F({\vec{\lambda}})$.
\label{th:haines} \end{theorem}
The key to the proof of this theorem is the following lemma.
\begin{lemma} Let
$$T_e(\mu,\lambda,\nu) \;\;=\;\;
\begin{tikzpicture}[baseline,semithick,darkred]
\draw[midto] (-1,0) -- (1,-0.5) node[black,midway,below] {$\mu$};
\draw[midto] (-1,0) -- (1.5,0.5) node[black,midway,above] {$\nu$};
\draw[midto] (1,-0.5) -- (1.5,0.5) node[black,midway,right] {$\lambda$};
\end{tikzpicture}$$
be a triangle with a minuscule edge $\lambda$, based at the edge $e$ of
length $\mu$. Then $Q(T_e(\mu,\lambda,\nu))$ is non-empty if and only if
there exists $w \in W$ such that $\mu + w\lambda = \nu$. If it is non-empty,
then it is smooth and has complex dimension $\braket{\nu-\mu+\lambda,\rho^\vee}$.
\label{l:smooth} \end{lemma}
\begin{proof}
Let $W(\mu)$ denote the stabilizer of $\mu$ in the Weyl group. It is a
parabolic subgroup of $W$.
Let us choose the base edge in $\mathrm{Gr}$ to be the edge connecting $t^{-\mu}$
and $t^0$. Then the edge based configuration space $Q(T_e(\mu, \lambda,
\nu))$ is a subvariety of $\mathrm{Gr}(\lambda)$ since there is only one free
vertex. In fact
$$Q(T_e(\mu, \lambda, \nu)) =
\{ p \in \mathrm{Gr}(\lambda) | d(t^{-\mu}, p) = \nu \}.$$
Let $A$ denote the set $W/W(\lambda)$, which we regard as a poset
using the opposite Bruhat order. With this order, $A$ becomes
the poset of $B$-orbits on $\mathrm{Gr}(\lambda) = {G^\vee}/M(\lambda)$, where $B$
is the Borel subgroup of ${G^\vee}$. We will be interested in the action of
$W(\mu)$ on $A$ by left multiplication. The quotient $W(\mu) \setminus
A$ is the set of $M_+(\mu)$ orbits on $\mathrm{Gr}(\lambda)$, where $M_+(\mu) =
\mathrm{Stab}_{G^\vee}(t^{-\mu})$ is the parabolic subgroup corresponding to
the minuscule weight $\mu$.
Hence we can write any point $p$ of $\mathrm{Gr}(\lambda)$ as $p = g t^{a \lambda}$
where $g \in M_+(\mu)$ and $a \in A$ is chosen to be a maximal length
representative for the orbit of $W(\mu)$. The action of
$M_+(\mu)$ on $\mathrm{Gr}$ stabilizes $t^{-\mu}$ so
$$d(t^{-\mu}, gt^{a \lambda})
= d(t^{-\mu}, t^{a\lambda}) = d(t^0, t^{\mu + a \lambda}).$$
Now, we claim that $\mu + a \lambda$ is always dominant. Let us write
$a = [w]$ for $w \in W$. We must check that
$$\braket{\mu + w\lambda, \alpha^\vee_i} =
\braket{\mu, \alpha^\vee_i} + \braket{\lambda, w \alpha^\vee_i} \ge 0$$
for all simple coroots $\alpha^\vee_i$. We break this calculation into
two cases.
First, suppose that $s_i \mu = \mu$. Then $\braket{\mu, \alpha^\vee_i} = 0$.
On the other hand $s_i w > w$ (in the usual Bruhat order) by the maximality
of $a$ in the $W(\mu)$-orbit. This implies that $w \alpha^\vee_i$ is a
positive coroot, which implies that $\braket{\lambda, w \alpha^\vee_i }$
is non-negative (since $\lambda$ is dominant). Hence
$$\braket{\mu, \alpha^\vee_i} + \braket{\lambda, w \alpha^\vee_i} \ge 0.$$
Next, suppose that $s_i \mu \ne \mu$. Then since $\mu$ is dominant,
$\braket{\mu, \alpha^\vee_i} \ge 1$. On the other hand, $|\braket{ \lambda,
w \alpha^\vee_i}| \le 1$ since $w \alpha^\vee_i$ is a coroot and $\lambda$
is minuscule. Hence
$$\braket{\mu, \alpha^\vee_i} + \braket{\lambda, w \alpha^\vee_i} \ge 0$$
in this case as well.
Since $\mu + a\lambda$ is always dominant, we conclude that
$$d(t^{-\mu}, gt^{a \lambda}) = \mu + a\lambda.$$
Hence, $Q(T_e(\mu,\lambda,\nu))$ is non-empty iff there exists $w \in
W$ such that $\mu + w\lambda = \nu$. (The above argument shows that
$[w]$ will necessarily be a maximal length representative for the
$W(\mu)$ action on $A$.) If such $w$ exists, then the configuration
space $Q(T_e(\mu,\lambda,\nu))$ is simply the $M(\mu)$-orbit through
$t^{w\lambda}$. Hence it is smooth and its dimension is given by
the length of $[w]$ in $A$ because it is of the same dimension as the
$B$-orbit through $t^{w\lambda}$. Since $\lambda$ is minuscule, this
equals $\braket{w \lambda + \lambda, \rho^\vee}$ as desired.
\end{proof}
\begin{proof}[Proof of \thm{th:haines}] It is easy to show by induction
that the fan configuration space
$$Q(A({\vec{\lambda}},{\vec{\mu}})) = Q(P_e(\mu_0,\lambda,\mu_1)) \tilde{\times}
\cdots \tilde{\times} Q(P_e(\mu_{n-1},\lambda_n,\mu_n))$$
is an iterated twisted product of triangle configuration spaces.
Since each factor has a minuscule edge, \lem{l:smooth} tells us that
$Q(A({\vec{\lambda}},{\vec{\mu}}))$ is also a smooth variety. Moreover, the dimensions
add to tell us that
$$\dim_\mathbb{C} Q(A({\vec{\lambda}},{\vec{\mu}})) = \braket{\lambda_1 + \dots + \lambda_n,\rho^\vee}
= \dim_\mathbb{C} F({\vec{\lambda}}).$$
On the other hand, $F({\vec{\lambda}}) = Q(P({\vec{\lambda}}))$ is partitioned as
a set by the subvarieties $Q(A({\vec{\lambda}},{\vec{\mu}}))$, simply by taking the
distances between the vertices of $P({\vec{\lambda}})$ and the origin. If $X$
is any algebraic variety with an equidimensional partition into smooth
varieties $X_1,\ldots,X_N$, then $X$ has pure dimension and its components
are the closures of the parts $X_k$. In our case, $X = F({\vec{\lambda}})$.
\end{proof}
It will be convenient later to abbreviate the dimension
of $F({\vec{\lambda}})$ as:
$$d({\vec{\lambda}}) \stackrel{\mathrm{def}}{=} \braket{\lambda_1 + \dots + \lambda_n,\rho^\vee}
= \dim_\mathbb{C} F({\vec{\lambda}}).$$
The same integers also arise in a different dimension formula:
$$\dim_\mathbb{C} \mathrm{Gr}({\vec{\lambda}}) = 2d({\vec{\lambda}}).$$
(Indeed, $\mathrm{Gr}({\vec{\lambda}})$ is a top-dimensional component of $F({\vec{\lambda}}
\sqcup {\vec{\lambda}}^*)$, given by collapsing the polygon $P({\vec{\lambda}} \sqcup
{\vec{\lambda}}^*)$ onto the polyline $L({\vec{\lambda}})$.)
Another important corollary of \lem{l:smooth} is the following:
\begin{theorem} Suppose that $D$ is a diskoid with boundary ${\vec{\lambda}}$
with no internal vertices, and suppose that all edges of $D$
(including the terms of ${\vec{\lambda}}$) are minuscule. Then $Q(D)$
is smooth and projective, and therefore a single component
of $F({\vec{\lambda}})$.
\label{th:smooth} \end{theorem}
\begin{proof} Let $T_e(\mu,\lambda,\nu)$ be a triangle of $D$ with
three minuscule edges, and let the base edge $e$ be any of the edges.
Then by \lem{l:smooth}, $Q(T_e(\mu,\lambda,\nu))$ is smooth. Likewise
$T_p(\mu,\lambda,\nu)$, based at a point $p$ instead, is smooth.
By construction, $Q(D)$ is a twisted product of configuration spaces of
this form, so it is also smooth. It is also projective since $D$ is a
minuscule linkage.
There is one delicate point in the inference that $Q(D)$ is a component of
$F({\vec{\lambda}})$: Is the restriction map $Q(D) \to F({\vec{\lambda}})$ injective?
As in the proof of \lem{l:smooth}, the restriction map
$$\pi:Q(T_e(\mu,\lambda,\nu)) \to \mathrm{Gr}(\lambda)$$
is injective, and so is the restriction map
$$\pi:Q(T(\mu,\lambda,\nu)) \to \mathrm{Gr}(\mu,\lambda).$$
The diskoid $D$ must have a triangle with at least two edges on the boundary,
so by induction its restriction map to $F({\vec{\lambda}})$ is also injective.
\end{proof}
\subsection{A homological state model}
\label{s:hstate}
This subsection discusses our motivation for the technical constructions
in the remainder of \sec{s:minuscule}.
We would like to use \thm{th:satake} as a state model or counting model
to evaluate webs in $\mathrm{\mathbf{rep}}^u(G)$. If $w$ is a web with dual diskoid $D$,
then there is a map of linkages
$$P({\vec{\lambda}}) = \partial D \longrightarrow \Gamma(D)$$
given by the inclusion of the boundary. This gives rise to a restriction map
$$\pi = \pi^{\Gamma(D)}_{P({\vec{\lambda}})}: Q(D) \to F({\vec{\lambda}}).$$
A point in $Q(D)$ is a ``state" of $D$ in the sense of mathematical
physics, in which each vertex of $D$ (or each face of $w$) is assigned
an element of $\mathrm{Gr}$. We would like to count the number of states of $D$
with some fixed boundary, or in other words the cardinality of a diskoid
fibre $\pi^{-1}(f)$ for $f \in F({\vec{\lambda}})$. If $f$ is chosen generically
in a top-dimensional component of $F({\vec{\lambda}})$, then optimistically this
cardinality will be the coefficient of $\Psi(w)$ in the Satake basis.
However, this sketch is naive. The diskoid fibre $\pi^{-1}(f)$ often
has a complicated geometry for which it is hard to define ``counting".
The first and main solution for us is to replace counting by a homological
intersection. (In \sec{s:euler} we will propose a second solution, in which
we count by taking the Euler characteristic of the fibre.) In particular,
for each web $w$, we will define a homology class $c(w) \in H_\top(Q(D))$
such that $\pi_*(c(w))$ equals $\Psi(w)$.
\subsection{The homology convolution category}
\label{s:conv}
If $M$ is an algebraic variety over $\mathbb{C}$, we will consider its
intersection cohomology sheaf $IC_M$ as a simple object in the category
of perverse sheaves on $M$. If $M$ is smooth, then $IC_M$ is isomorphic
to $\mathbb{C}_M[\dim_\mathbb{C} M]$, the constant sheaf shifted by the complex dimension
of $M$. For brevity, we will write this perverse sheaf as $\mathbb{C}[M]$.
The geometric Satake correspondence is a tensor functor that takes the usual
product on $\mathrm{\mathbf{rep}}^u(G)$ to the convolution tensor product on $\mathrm{\mathbf{perv}}(\mathrm{Gr})$.
In particular, the tensor product $V({\vec{\lambda}})$ of irreducible minuscule
representations corresponds to the convolution tensor product of the simple
perverse sheaves $\mathbb{C}[\mathrm{Gr}(\lambda_i)]$ on minuscule spheres, which are
closed in the algebraic geometry. By definition, this convolution tensor
product is given by the pushforward $(m_{{\vec{\lambda}}})_*(\mathbb{C}[\mathrm{Gr}({\vec{\lambda}})])$
along the convolution morphism.
Let $\mathrm{\mathbf{perv}}(\mathrm{Gr})_\min$ denote the subpivotal category of $\mathrm{\mathbf{perv}}(\mathrm{Gr})$
consisting of such pushforwards. By construction, $\mathrm{\mathbf{perv}}(\mathrm{Gr})_\min$
is equivalent to $\mathrm{\mathbf{rep}}^u(G)_\min$. Our goal is to study $\mathrm{\mathbf{perv}}(\mathrm{Gr})_\min$
using convolutions in homology, following ideas of Ginzburg. We begin by
reviewing some generalities, following \cite[Sec. 2.7]{CG:complex}.
Let $\{M_i\}$ be a set of connected, smooth complex varieties and let $M_0$
be a possibly singular, stratified variety with strata $\{U_\alpha\}$.
For each $i$, let $\pi_i:M_i \to M_0$ be a proper semismall map.
In this context, the statement that $\pi_i$ is semismall means that
$\pi_i$ restricts to a fibre bundle over each stratum $U_\alpha$ and
that the dimensions of these fibres is given by
$$\dim_\mathbb{C} \pi_i^{-1}(u) = \frac{\dim_\mathbb{C} M_i - \dim_\mathbb{C} U_\alpha}2$$
for $u \in U_\alpha$ (note that we have equality above). Let $d_i =
\dim_\mathbb{C} M_i$.
With this setup, let $Z_{ij} = M_i \times_{M_0} M_j$. The semismallness
condition implies that $\dim_\mathbb{C} Z_{ij} = \frac{d_i + d_j}{2}$. Let
$$H_\top(Z_{ij}) = H_{d_i + d_j}(Z_{ij})$$
be the top homology of $Z_{ij}$. If the $M_i$ are proper, which they
will be in our situation, then we will obtain a valid definition of the
convolution product using the ordinary singular homology of $Z_{ij}$.
(Otherwise the correct type of homology would be Borel-Moore homology.)
Define a \emph{homological convolution product}
$$*:H_\top(Z_{ij}) \tensor H_\top(Z_{jk}) \to H_\top(Z_{ik})$$
by the formula
$$c_1 * c_2 = (\pi_{ik})_* (\pi_{ij}^*(c_1) \cap \pi_{jk}^*(c_2)),$$
where ``$\cap$" denotes the intersection product (with support), relative
to the ambient smooth manifold $M_i \times M_j \times M_k$. This may
be defined using the cup product in cohomology via Poincar\'e duality.
For more details about this construction, see \cite[Sec. 2.6.15]{CG:complex}
or \cite[Sec. 19.2]{Fulton:intersect}. Note that because
$$\dim_\mathbb{C} Z_{ij} = \frac{d_i + d_j}{2},$$
the correct homological degree is preserved by the convolution product.
This construction is relevant for us because of a theorem of Ginzburg
that relates $H_\top(Z_{ij})$ to morphisms in the category $\mathrm{\mathbf{perv}}(M_0)$
of perverse sheaves on $M_0$.
\begin{theorem} \cite[Thm. 8.6.7]{CG:complex}
With the above setup, there is an isomorphism
$$H_\top(Z_{ij}) \cong
\mathrm{Hom}_{\mathrm{\mathbf{perv}}(M_0)} \bigl( (\pi_i)_*\mathbb{C}[M_i], (\pi_j)_*\mathbb{C}[M_j] \bigr).$$
This isomorphism identifies convolution products on the left side with
compositions of morphisms on the right side.
\label{th:ginzburg}
\end{theorem}
We will apply this setup by letting $M_0 = \mathrm{Gr}$ and by letting each $M_i$
be $\mathrm{Gr}({\vec{\lambda}})$ for a sequence ${\vec{\lambda}}$ of dominant minuscule weights.
The convolution morphism $m_{\vec{\lambda}}: \mathrm{Gr}({\vec{\lambda}}) \to \mathrm{Gr}$ is semismall.
(See \cite[Lem. 4.4]{MV:geometric}; it also follows from the proof of
\thm{th:haines}.) Then $Z_{ij}$ becomes
$$Z({\vec{\lambda}}, {\vec{\mu}}) = \mathrm{Gr}({\vec{\lambda}}) \times_\mathrm{Gr} \mathrm{Gr}({\vec{\mu}})
= Q(P({\vec{\lambda}}^* \sqcup {\vec{\mu}})),$$
where $P({\vec{\lambda}}^* \sqcup {\vec{\mu}})$ is this polygon:
$$P({\vec{\lambda}}^* \sqcup {\vec{\mu}}) \;\;=\;\; \begin{tikzpicture}[baseline]
\draw[darkred,semithick] (0,0) -- (1,1) -- (2,.5) -- (3,1) -- (3.5,0)
-- (2,-1) -- (1,-1) -- cycle;
\draw[darkred,semithick,midto] (1,1) -- (2,.5);
\draw[darkred,semithick,midto] (1,-1) -- (2,-1);
\fill[darkred] (0,0) circle (.07);
\path[draw=darkred,fill=white] (3.5,0) circle (.07);
\draw[anchor=north] (1.75,-1.1) node {${\vec{\lambda}}$};
\draw[anchor=south] (1.75,.9) node {${\vec{\mu}}$};
\end{tikzpicture}$$
\thm{th:ginzburg} motivates the following construction of a category
$\mathrm{\mathbf{hconv}}(\mathrm{Gr})$. The objects in $\mathrm{\mathbf{hconv}}(\mathrm{Gr})$ are the polyline varieties
$\mathrm{Gr}({\vec{\lambda}})$, where ${\vec{\lambda}}$ is a sequence minuscule weights.
The tensor product on objects is, by definition, given by convolution on
objects, so
$$\mathrm{Gr}({\vec{\lambda}}) \tensor \mathrm{Gr}({\vec{\mu}}) \stackrel{\mathrm{def}}{=} \mathrm{Gr}({\vec{\lambda}} \sqcup {\vec{\mu}}),$$
where $\sqcup$ denotes concatenation of sequences. So the identity object
is the point $\mathrm{Gr}(\emptyset)$. Finally the dual object $\mathrm{Gr}({\vec{\lambda}})^*
= \mathrm{Gr}({\vec{\lambda}}^*)$ of $\mathrm{Gr}({\vec{\lambda}})$ is given by reversing ${\vec{\lambda}}$
and taking the dual of each of its terms.
We define the morphism spaces of $\mathrm{\mathbf{hconv}}(\mathrm{Gr})$ as
$$\mathrm{Hom}_{\mathrm{\mathbf{hconv}}(\mathrm{Gr})}(\mathrm{Gr}({\vec{\lambda}}), \mathrm{Gr}({\vec{\mu}}))
\stackrel{\mathrm{def}}{=} H_\top(Z({\vec{\lambda}}, {\vec{\mu}})).$$
The composition of morphisms is given by the convolution product. Note that
the identity morphism $1_{{\vec{\lambda}}} \in H_\top(Z({\vec{\lambda}}, {\vec{\lambda}}))$
is given by the class $[\mathrm{Gr}({\vec{\lambda}})_\Delta]$ of the diagonal
$$\mathrm{Gr}({\vec{\lambda}})_\Delta \subseteq Z({\vec{\lambda}},{\vec{\lambda}}) \subseteq
\mathrm{Gr}({\vec{\lambda}}) \times \mathrm{Gr}({\vec{\lambda}}),$$
\emph{i.e.}, it is the configurations in which the polygon $P({\vec{\lambda}}^* \sqcup
{\vec{\lambda}})$ has collapsed onto the polyline $L({\vec{\lambda}})$.
To describe the tensor structure on morphisms, it is enough to describe
how to tensor with the identity morphism. So let ${\vec{\lambda}}, {\vec{\mu}}, {\vec{\nu}}$ be
three sequences of dominant minuscule weights and let $c \in H_\top(Z({\vec{\mu}},
{\vec{\nu}}))$. Our goal is to construct a class
$$1_{{\vec{\lambda}}} \tensor c \in
H_\top(Z({\vec{\lambda}} \sqcup {\vec{\mu}}, {\vec{\lambda}} \sqcup {\vec{\nu}}))$$
For the moment, let $\Gamma$ be a $\rho$-shaped graph with a tail of type
${\vec{\lambda}}$ and a loop of type ${\vec{\mu}}^* \sqcup {\vec{\nu}}$, based at the end of
the tail:
$$\Gamma \;\;=\;\; \begin{tikzpicture}[baseline]
\draw[darkred,semithick] (0,0) -- (1,1) -- (2,.5) -- (3,1) -- (3.5,0)
-- (2,-1) -- (1,-1) -- cycle;
\draw[darkred,semithick] (0,0) -- (-1,.5) -- (-2,0);
\draw[darkred,semithick,midto] (1,1) -- (2,.5);
\draw[darkred,semithick,midto] (1,-1) -- (2,-1);
\draw[darkred,semithick,midto] (-2,0) -- (-1,.5);
\fill[darkred] (-2,0) circle (.07);
\path[draw=darkred,fill=white] (0,0) circle (.07);
\path[draw=darkred,fill=white] (3.5,0) circle (.07);
\draw[anchor=north] (1.75,-1.1) node {${\vec{\mu}}$};
\draw[anchor=south] (1.75,.9) node {${\vec{\nu}}$};
\draw[anchor=north] (-1,.1) node {${\vec{\lambda}}$};
\end{tikzpicture}$$
Let $X = Q(\Gamma)$ be its based configuration space. We describe two
fibration constructions related to $X$. First, there is a restriction map
$$\pi_{L({\vec{\lambda}}) \sqcup {\mathrm{pt}}}^{L({\vec{\lambda}} \sqcup {\vec{\mu}})}:
\mathrm{Gr}({\vec{\lambda}} \sqcup {\vec{\mu}}) \to \mathrm{Gr}({\vec{\lambda}}) \times \mathrm{Gr}$$
given by restricting to the polyline $L({\vec{\lambda}})$ and the free endpoint
of $L({\vec{\lambda}} \sqcup {\vec{\mu}})$. Then $X$ is the fibred product
$$X = \mathrm{Gr}({\vec{\lambda}} \sqcup {\vec{\mu}})
\times_{\mathrm{Gr}({\vec{\lambda}}) \times \mathrm{Gr}} \mathrm{Gr}({\vec{\lambda}} \sqcup {\vec{\nu}}).$$
Second, there is a projection
$$\pi_{L({\vec{\lambda}})}^\Gamma:X \to \mathrm{Gr}({\vec{\lambda}})$$
given by restricting from $\Gamma$ to $L({\vec{\lambda}})$. The fibres
of this projection are $Z({\vec{\mu}},{\vec{\nu}})$.
Since $\mathrm{Gr}({\vec{\lambda}})$ is simply connected, we get an isomorphism
$$H_\top(X) \cong H_\top(\mathrm{Gr}({\vec{\lambda}})) \tensor H_\top(Z({\vec{\mu}}, {\vec{\nu}}))$$
and thus we obtain an isomorphism
$$H_\top(Z({\vec{\mu}}, {\vec{\nu}})) \stackrel{\cong}{\longrightarrow} H_\top(X)$$
given by $c \mapsto [\mathrm{Gr}({\vec{\lambda}})] \tensor c$.
There is also an inclusion
$$i = \pi_{P({\vec{\lambda}} \sqcup {\vec{\mu}} \sqcup {\vec{\nu}}^* \sqcup {\vec{\lambda}}^*)}^\Gamma:
X \to Z({\vec{\lambda}} \sqcup {\vec{\mu}}, {\vec{\lambda}} \sqcup {\vec{\nu}}),$$
using the polygon which travels twice along the tail of $\Gamma$
and around the loop of $\Gamma$. Combining all this structure, we define
$$1_{{\vec{\lambda}}} \tensor c \stackrel{\mathrm{def}}{=} i_*( [\mathrm{Gr}({\vec{\lambda}})] \tensor c).$$
Tensoring by the identity morphism on the other side is similar and we
leave the construction to the reader.
Finally, to define the cap and cup morphisms for any ${\vec{\lambda}}$, we will
define them for a single minuscule weight $\lambda$. Note that
$$Z(\lambda \sqcup \lambda^*, \emptyset)
= Z(\emptyset, \lambda \sqcup \lambda^*)
= F(\lambda, \lambda^*) \cong \mathrm{Gr}(\lambda).$$
We define the cup $b_\lambda$ and the cap $d_\lambda$ to each be the class
$[\mathrm{Gr}(\lambda)]$ in their respective hom spaces.
\begin{theorem} There is an equivalence of pivotal categories
$$\mathrm{\mathbf{hconv}}(\mathrm{Gr}) \cong \mathrm{\mathbf{perv}}(\mathrm{Gr})_\min \cong \mathrm{\mathbf{rep}}^u(G)_\min.$$
\label{th:convperv} \end{theorem}
Applying \thm{th:convperv} to invariant spaces, we obtain an isomorphism
$$\mathrm{Inv}(V({\vec{\lambda}})) \cong \mathrm{Hom}_{\mathrm{\mathbf{hconv}}(\mathrm{Gr})}(\mathrm{Gr}(\emptyset),\mathrm{Gr}({\vec{\lambda}}))
= H_\top(Z(\emptyset, {\vec{\lambda}})) = H_\top(F({\vec{\lambda}})),$$
which is \thm{th:satbasis}.
\begin{proof} The second equivalence is geometric Satake, so we will just
prove the first equivalence. We begin by showing that it is an equivalence
of monoidal categories.
By the definition, the objects in both categories are parameterized
by sequences ${\vec{\lambda}}$, so the functor on objects is very simple. On
morphisms, the functor is given by the isomorphisms from \thm{th:ginzburg}.
By this theorem, the functor is fully faithful and is compatible with
composition on both sides. (\emph{I.e.}, it is a functor.) To complete the proof
this theorem, we need only to show that the functor is compatible with
the tensor product and with pivotal duality.
To see that it is compatible with the tensor product, we use the same
notation as above. If
$$c \in \mathrm{Hom}((m_{\vec{\mu}})_* \mathbb{C}[\mathrm{Gr}({\vec{\mu}})], (m_{\vec{\nu}})_*\mathbb{C}[\mathrm{Gr}({\vec{\nu}})]),$$
then with respect to the tensor structure in $\mathrm{\mathbf{perv}}(\mathrm{Gr})$,
$I_{(m_{\vec{\lambda}})_*\mathbb{C}[\mathrm{Gr}({\vec{\lambda}})]} \tensor c$ is given by the image of
$c$ under the map
\begin{multline*}
\mathrm{Hom}_{\mathrm{\mathbf{perv}}(\mathrm{Gr})}\left( (m_{\vec{\mu}})_* \mathbb{C}[\mathrm{Gr}({\vec{\mu}})],
(m_{\vec{\nu}})_* \mathbb{C}[\mathrm{Gr}({\vec{\nu}})] \right) \stackrel{\cong}{\longrightarrow} \\
\mathrm{Hom}_{\mathrm{\mathbf{perv}}(\mathrm{Gr}({\vec{\lambda}}) \times \mathrm{Gr})} \left( (\pi_{L({\vec{\lambda}}) \sqcup {\mathrm{pt}}}
^{L({\vec{\lambda}} \sqcup {\vec{\mu}})})_* \mathbb{C}[\mathrm{Gr}({\vec{\lambda}} \sqcup {\vec{\mu}})],
(\pi_{L({\vec{\lambda}}) \sqcup {\mathrm{pt}}}^{L({\vec{\lambda}} \sqcup {\vec{\nu}})})_*\mathbb{C}[\mathrm{Gr}({\vec{\lambda}}
\sqcup {\vec{\nu}})] \right) \stackrel{p_*}{\longrightarrow} \\
\mathrm{Hom}_{\mathrm{\mathbf{perv}}(\mathrm{Gr})} \left( (\pi_{{\vec{\lambda}} \sqcup {\vec{\mu}}})_*
\mathbb{C}[\mathrm{Gr}({\vec{\lambda}} \sqcup {\vec{\mu}})], (m_{{\vec{\lambda}} \sqcup {\vec{\nu}}})_*
\mathbb{C}[\mathrm{Gr}({\vec{\lambda}} \sqcup {\vec{\nu}})] \right).
\end{multline*}
Here $p: \mathrm{Gr}({\vec{\lambda}}) \times \mathrm{Gr} \to \mathrm{Gr}$ is the projection onto the
second factor. This is easily seen to match our above definition.
It remains to check that the pivotal structures match under this equivalence.
Recall from \sec{s:signs} that the pivotal structures on $\mathrm{\mathbf{rep}}(G)$ are
determined by the dimensions $\dim(V(\lambda))$, which are by definition
the values of closed loops. (The discussion there is for pivotal structures
that differ by a sign, but it is true in general.) Moreover, the discrepancy
is multiplicative, so it only needs to be checked for minuscule $\lambda$.
Let $\lambda$ be minuscule. In $\mathrm{\mathbf{hconv}}(\mathrm{Gr})$, the value of a loop labelled
$\lambda$, \emph{i.e.}, the composition
$$d_\lambda \circ b_\lambda \in \mathrm{Hom}(\mathrm{Gr}(\emptyset), \mathrm{Gr}(\emptyset)) = \mathbb{C},$$
is given by the self-intersection of $\mathrm{Gr}(\lambda) \cong F(\lambda,
\lambda^*)$ with itself inside $\mathrm{Gr}(\lambda, \lambda^*)$.
There is a neighbourhood (defined using the pullback of the open big cell)
of $F(\lambda, \lambda^*)$ in $\mathrm{Gr}(\lambda, \lambda^*)$ which is isomorphic
to $T^*\mathrm{Gr}(\lambda)$, under an isomorphism which carries $F(\lambda,
\lambda^*)$ to the zero section $\mathrm{Gr}(\lambda)$.
For any compact, complex $d$-manifold $X$, the self-intersection of $X$
with itself inside $T^*X$ is $(-1)^d \chi(X)$, where $\chi(X)$ is the Euler
characteristic of $X$. (The self-intersection in $TX$ is $\chi(X)$, and
for a complex $d$-manifold the cotangent bundle $T^*X$ has the opposite real
orientation exactly when $d$ is odd.) Applying this to $X = \mathrm{Gr}(\lambda)$,
we conclude that
$$d_\lambda \circ b_\lambda = (-1)^d \chi(\mathrm{Gr}(\lambda))
= (-1)^{\braket{2\lambda, \rho^\vee}} \dim V(\lambda).$$
This is the sign correction that is used to define the pivotal structure
on $\mathrm{\mathbf{rep}}^u(G)$, as desired.
\end{proof}
\subsection{From the free spider to the convolution category}
\sec{s:spiders} describes a pivotal functor
$$\Psi:\mathrm{\mathbf{fsp}}(G) \to \mathrm{\mathbf{rep}}^u(G)_\min.$$
On the other hand, the geometric Satake correspondence and \thm{th:convperv}
yield equivalences
$$\mathrm{\mathbf{rep}}^u(G)_\min \cong \mathrm{\mathbf{perv}}(\mathrm{Gr})_\min \cong \mathrm{\mathbf{hconv}}(\mathrm{Gr}).$$
The composition is a functor $\mathrm{\mathbf{fsp}}(G) \to \mathrm{\mathbf{hconv}}(\mathrm{Gr})$ which we will
also denote by $\Psi$. Our goal now is to describe this functor and in
particular its action on invariant vectors.
Let $\lambda, \mu, \nu$ be a triple of dominant minuscule weights such that
$$\mathrm{Inv}_G(V(\lambda, \mu, \nu)) \ne 0.$$
There is a simple web $w \in
\mathrm{Inv}_{\mathrm{\mathbf{fsp}}(G)}(\lambda, \mu, \nu)$ which contains a single vertex. On the
other hand,
$$\mathrm{Inv}_{\mathrm{\mathbf{hconv}}(\mathrm{Gr})}(\lambda, \mu, \nu) \cong H_\top(F(\lambda, \mu, \nu))$$
is one-dimensional with canonical generator $[F(\lambda, \mu, \nu)]$.
Recall from \sec{s:spiders} that in the construction of the functor
$\mathrm{\mathbf{fsp}}(G) \to \mathrm{\mathbf{rep}}^u(G)_\min$, there was some freedom to choose the image
of the simple web $w$ (it was only defined up to a non-zero scalar). Now,
we fix this choice by setting
$$\Psi(w) \stackrel{\mathrm{def}}{=} [F(\lambda, \mu, \nu)].$$
The functor $\Psi$ is now determined by what it does on vertices and the
fact that it preserves the pivotal structure on both sides.
We are now in a position to prove \thm{th:homclass}, which we will restate
as follows. Recall that
$$d({\vec{\lambda}}) = \dim_\mathbb{C} F({\vec{\lambda}}).$$
\begin{theorem} Let $w$ be a web with boundary ${\vec{\lambda}}$ and dual diskoid
$D = D(w)$. Let
$$\pi:Q(D) \to F({\vec{\lambda}})$$
be the boundary restriction map. There exists a homology class $c(w)
\in H_{2d({\vec{\lambda}})}(Q(D))$ such that $\pi_*(c(w)) = \Psi(w)$. Moreover,
when $Q(D)$ has dimension $d({\vec{\lambda}})$ and is reduced as a scheme, then
$c(w)$ is the fundamental class $[Q(D)]$.
\label{th:homclass2} \end{theorem}
\begin{proof}
We begin by picking a isotopy representative for $w$ such that the height
function is a Morse function and so that the boundary of $w$ is at the
top level. We assume a sequence of horizontal lines $\ell_0,\ldots,\ell_m$
such that in between each pair, $w$ has only a single cap, cup, or a vertex.
We assume further that each vertex is either an ascending Y (it is in the
shape of a Y) or a descending Y (an upside-down Y).
\begin{fullfigure}{f:morseweb}{A web for $\mathrm{SL}(9)$ in Morse position}
\begin{tikzpicture}[scale=0.5]
\begin{scope}[web]
\draw[midto] (0,10) -- (0,8) node[black,left,midway] {$\omega_2$};
\draw[midto] (0,8) -- (0,6);
\draw[midto] (0,6) -- (1,5); \draw[midfrom] (1,5) -- (1,4);
\draw[midfrom] (1,4) -- (1,2);
\draw[midfrom] (3,10) -- (3,9) node[black,left,midway] {$\omega_3$};
\draw[midto] (3,9) -- (2,8);
\draw[midto] (2,8) -- (2,6) node[black,left,midway] {$\omega_3$};
\draw[midto] (2,6) -- (1,5);
\draw[midto] (3,9) -- (4,8) node[black,right,midway] {$\omega_3$};
\draw[midto] (4,8) -- (5,7); \draw[midto] (5,7) -- (5,6);
\draw[midto] (5,6) -- (5,4) node[black,left,midway] {$\omega_8$};
\draw[midto] (5,4) -- (6,3);
\draw[midto] (6,3) -- (6,2);
\draw[midto] (6,10) -- (6,8) node[black,left,midway] {$\omega_5$};
\draw[midto] (6,8) -- (5,7);
\draw[midfrom] (7,10) -- (7,8) node[black,right,midway] {$\omega_4$};
\draw[midfrom] (7,8) -- (7,6); \draw[midfrom] (7,6) -- (7,4);
\draw[midfrom] (7,4) -- (6,3);
\draw[midfrom] (1,2) .. controls (1,0) and (6,0) .. (6,2)
node[black,above,midway] {$\omega_4$};
\end{scope}
\foreach \k in {0,1,...,5}
\draw[dashed] ($(-1,2*\k)$) -- ($(8,2*\k)$) node[right] {$\ell_\k$};
\end{tikzpicture}
\end{fullfigure}
Let ${\vec{\lambda}}^{(k)}$ be the vector of labels of the edges cut by
the horizontal line $\ell_k$. Then ${\vec{\lambda}}^{(0)} = \emptyset$ and
${\vec{\lambda}}^{(m)} = {\vec{\lambda}}$. For example, in \fig{f:morseweb} shows an
$\mathrm{SL}(9)$ web in Morse position, with edges labelled by its minuscule
weights $\omega_k$ with $1 \le k \le 8$. In this example,
$${\vec{\lambda}}^{(1)}=\{\omega_4,\omega_5\} \qquad
{\vec{\lambda}}^{(3)}=\{\omega_7,\omega_6,\omega_1,\omega_4\}.$$
(Note that in $\mathrm{SL}(n)$ in general, $\omega_k^* = \omega_{n-k}$; if an edge
points down as it crosses a line, then we must take the dual weight.)
Let
$$w_k \in \mathrm{Hom}_{\mathrm{\mathbf{fsp}}(G)}({\vec{\lambda}}^{(k-1)}, {\vec{\lambda}}^{(k)})$$
denote the web in the horizontal strip between the lines $\ell_{k-1}$
and $\ell_k$. By examining the above definition, we see that for each
$1 \le k \le m$, there exists a component $X_k \subset Z({\vec{\lambda}}^{(k-1)},
{\vec{\lambda}}^{(k)})$ such that $\Psi(w_k) = [X_k]$. We would like to describe
this component explicitly. For convenience, if
$$\vec{p} = (p_0,p_1,\ldots,p_m) \in \mathrm{Gr}^{m+1}$$
(with $p_0 = t^0$ for us), define $\sigma_i(\vec{p})$ by omitting the term $p_i$.
\begin{enumerate}
\item If $w_k$ is an ascending Y vertex that connects the $i$th point on
$\ell_{k-1}$ to the $i$th and $i+1$st points on $\ell_k$, then
$$X_k = \{(\vec{p},\vec{p}') \in Z({\vec{\lambda}}^{(k-1)}, {\vec{\lambda}}^{(k)})
| \vec{p} = \sigma_i(\vec{p}')\}.$$
\item If $w_k$ is a descending Y vertex that connects the $i$th
and $i+1$st points on $\ell_{k-1}$ to the $i$th point on $\ell_k$, then
$$X_k = \{(\vec{p},\vec{p}') \in Z({\vec{\lambda}}^{(k-1)}, {\vec{\lambda}}^{(k)})
| \vec{p}' = \sigma_i(\vec{p})\}.$$
\item If $w_k$ is a cup that connects the $i$th and $i+1$st points on $\ell_k$,
then
$$X_k = \{(\vec{p},\vec{p}') \in Z({\vec{\lambda}}^{(k-1)}, {\vec{\lambda}}^{(k)})
| \vec{p} = \sigma_i(\sigma_i(\vec{p}')) \}.$$
\item If $w_k$ is a cap that connects the $i$th and $i+1$st points on
$\ell_{k-1}$, then
$$X_k = \{(\vec{p},\vec{p}') \in Z({\vec{\lambda}}^{(k-1)}, {\vec{\lambda}}^{(k)})
| \vec{p}' = \sigma_i(\sigma_i(\vec{p})) \}.$$
\end{enumerate}
Then $w = w_m \circ \cdots \circ w_1$. Since $\Psi$ is a functor,
$$\Psi(w) = \Psi(w_m) * \cdots * \Psi(w_1) = [X_m] * \cdots * [X_1].$$
Now, compositions of convolutions can be computed as a single convolution as
$$[X_m] * \cdots * [X_1]
= (\pi_{0,m})_*(\pi_{0,1}^* [X_1] \cdots \pi_{m-1, m}^*[X_m]),$$
where the intersection products take place in the ambient smooth manifold
$$X = \mathrm{Gr}({\vec{\lambda}}^{(0)}) \times \cdots \times \mathrm{Gr}({\vec{\lambda}}^{(m)}).$$
Here $\pi_{k-1,k}$ denotes the projection from $X$ to
$\mathrm{Gr}({\vec{\lambda}}^{(k-1)}, {\vec{\lambda}}^{(k)})$.
From the definitions, we see that the diskoid configuration spaces $Q(D)$
can be obtained as
$$Q(D) = \pi_{0,1}^{-1}(X_1) \cap \cdots \cap \pi_{m-1, m}^{-1}(X_m).$$
Let
\begin{align*}
c(w) &= \pi_{0,1}^* [X_1] \cap \cdots \cap \pi_{m-1, m}^*[X_m] \\
&= [\pi_{0,1}^{-1}(X_1)] \cap \cdots \cap [\pi_{m-1, m}^{-1}(X_m)].
\end{align*}
Because we are using the intersection product with support, $c(w)$ lives
in $H_{d({\vec{\lambda}})}(Q(D)) $, the homology of the intersection. When $Q(D)$
is reduced of the expected dimension, then the intersection product of the
homology classes corresponds to the fundamental class of the intersection
(see \cite[Sec. 8.2]{Fulton:intersect}), so $c(w) = [Q(D)]$.
Finally, $\pi: Q(D) \to F({\vec{\lambda}})$ is the restriction of $\pi_{0,m}$
to $Q(D)$. Hence we conclude that $\Psi(w) = \pi_*(c(w))$.
\end{proof}
Because $\pi_*(c(w))$ is supported on $\pi(Q(D))$, we immediately obtain
the following.
\begin{corollary} $\Psi(w)$ is a linear combination of the fundamental
classes of the components of $F({\vec{\lambda}})$ which are in the image of $\pi$.
\label{c:support} \end{corollary}
It may not seem clear that $c(w)$ depends only on the web $w$, and not on
the Morse position of $w$ used to construct it. However, a posteriori,
this must be verified by checking that it is invariant under basic isotopy
moves (for example, straightening out a cup/cap pair).
\section{$\mathrm{SL}(3)$ results}
\label{s:sl3}
In this section, we will prove \thm{th:main} and \thm{th:unitriang}.
In preparation for this result, we need to use and extend the geometry of
non-elliptic webs. To review, if $w$ is an $A_2$ web and $D = D(w)$ is
its dual diskoid, then $w$ is non-elliptic if and only if $D$ is $\mathrm{CAT}(0)$.
\subsection{Geodesics in $\mathrm{CAT}(0)$ diskoids}
\label{s:geodesics}
We will be interested in combinatorial (meaning edge-travelling) geodesics
in a type $A_2$ diskoid $D$. These are equivalent to ``minimal cut paths"
of the dual web \cite{Kuperberg:spiders}, when the endpoints of the geodesic
are boundary vertices $D$. Here we will consider geodesics between vertices
that may be in the interior or on the boundary. If both vertices are on
the boundary, then the geodesic is called \emph{complete}.
\begin{fullfigure}{f:diamond}{Two geodesics $\gamma$ and $\gamma'$
connected by a diamond move}
\begin{tikzpicture}[x={(.75cm,0)},y={(.375cm,.659cm)}]
\draw[fill=lightgray] (0,0) -- (1,0) -- (1,1) -- (0,1) -- cycle;
\draw (1,0) -- (0,1); \draw (-2,0) -- (0,0);
\draw (1,1) -- (2,1) -- (2,2);
\draw[darkred,semithick] (-2,.15) -- (-.15,.15)
-- (-.15,1.15) -- (1.85,1.15) -- (1.85,2);
\draw[darkred,semithick] (-2,-.15) -- (1.15,-.15)
-- (1.15,.85) -- (2.15,.85) -- (2.15,2);
\node[anchor=south east] at (-.15,1.15) {$\gamma$};
\node[anchor=north west] at (1.15,-.15) {$\gamma'$};
\end{tikzpicture}
\end{fullfigure}
Geodesics in an $A_2$ diskoid are often not unique. Define a \emph{diamond
move} of a geodesic to be a move in which the geodesic crosses two triangles,
as in \fig{f:diamond}. (This is equivalent to an ``$H$-move" on a cut
path of a non-elliptic web.) We say that two geodesics are \emph{isotopic}
if they are equivalent with respect to diamond moves.
\begin{theorem} Let $p,q$ be two vertices of a $\mathrm{CAT}(0)$, type $A_2$
diskoid $D$. Then the geodesics between $p$ and $q$ subtend a diskoid
which is a skew Young diagram, with each square split into two triangles.
In particular, all geodesics are isotopic, $D$ is geodesically coherent,
and all geodesics lie between two extremal geodesics. Both of the extremal
geodesics are concave on the outside.
\label{th:young} \end{theorem}
Here a \emph{skew Young diagram} is the same as the usual object in
combinatorics with that name, namely the diskoid lying between two geodesic
lattice paths in $\mathbb{Z}^2$. \fig{f:skew} shows an example in which
the squares have been split so that it becomes an $A_2$ diskoid.
\begin{fullfigure}{f:skew}{A skew partition bounded by extremal geodesics
$\gamma$ and $\gamma'$}
\begin{tikzpicture}
\begin{scope}[x={(.75cm,0)},y={(.375cm,.659cm)}]
\foreach \x/\y in {0/0,1/0,2/1,3/1,2/2,3/2,3/3,5/4}
{
\draw[fill=lightgray] (\x,\y) -- ++(1,0) -- ++(0,1) -- ++(-1,0) -- cycle;
\draw (\x,\y) ++(1,0) -- ++(-1,1);
}
\draw (4,4) -- (5,4);
\draw[darkred,semithick] (-.15,-.15) -- ++(0,1.3) -- ++(2,0) -- ++(0,2)
-- ++(1,0) -- ++(0,1) -- ++(2,0) -- ++(0,1) -- ++(1.3,0);
\draw[darkred,semithick] (-.15,-.15) -- ++(2.3,0) -- ++(0,1) -- ++(2,0)
-- ++(0,3) -- ++(2,0) -- ++(0,1.3);
\coordinate (p) at (-.15,-.15); \coordinate (q) at (6.15,5.15);
\coordinate (g) at (1.85,3.15); \coordinate (g1) at (4.15,2.15);
\end{scope}
\fill[darkred] (p) circle (.07); \fill[darkred] (q) circle (.07);
\node[anchor=south east] at (p) {$p$};
\node[anchor=north west] at (q) {$q$};
\node[anchor=south east] at (g) {$\gamma$};
\node[anchor=north west] at (g1) {$\gamma'$};
\end{tikzpicture}
\end{fullfigure}
\thm{th:young} is proven in \cite{Kuperberg:spiders} in the case when $p$
and $q$ are on the boundary. If they are not on the boundary, then we can
reduce to previous case by the removing the simplices of $D$ that do not
lie between two geodesics. The final statement, that an extremal geodesic
$\gamma$ is concave outside of the skew Young diagram, is easy to check:
If $\gamma$ has an angle of $\pi/3$, then it is not a geodesic. If it has
an angle of $2\pi/3$, then an isotopy is available and it is not extremal.
\begin{lemma} If $p$ and $q$ are two vertices of a $\mathrm{CAT}(0)$ diskoid $D$,
then every geodesic $\kappa$ between them extends to a complete geodesic.
\label{l:complete} \end{lemma}
\begin{proof} The argument is based on a geodesic sweep-out construction.
We claim that we can make a sequence of geodesics
$${\vec{\gamma}} = (\gamma_0,\gamma_1,\dots,\gamma_{m-1})$$
from $p$ to the boundary $\partial D$ with certain additional properties.
We require that each consecutive pair $\gamma_k$ and $\gamma_{k+1}$ differ
by either an elementary isotopy or an elementary boundary isotopy (for each
$k \in \mathbb{Z}/m$). The latter consists either of appending an edge to
$\gamma_k$ or removing the last edge, or a \emph{triangle move} as in
\fig{f:triangle}.
\begin{fullfigure}{f:triangle}{A triangle move connecting
geodesics $\gamma_k$ and $\gamma_{k+1}$}
\begin{tikzpicture}
\begin{scope}[x={(.75cm,0)},y={(.375cm,.659cm)}]
\draw[fill=lightgray] (-1,0) -- (-1,-1) -- (0,-1) -- cycle;
\draw (-1,0) -- (0,-1);
\draw (-1,-1) -- (-2,-1) -- (-2,-2);
\draw[darkred,semithick] (.15,-1.15) -- (-1.85,-1.15) -- (-1.85,-2);
\draw[darkred,semithick] (-1.15,.15) -- (-1.15,-.85)
-- (-2.15,-.85) -- (-2.15,-2);
\coordinate (r1) at (.15,-1.15);
\coordinate (r2) at (-1.15,.15);
\coordinate (g1) at (-1,-1.15);
\coordinate (g2) at (-1.15,-.85);
\end{scope}
\fill[darkred] (r1) circle (.07);
\fill[darkred] (r2) circle (.07);
\node[anchor=north west] at (r1) {$r_k$};
\node[anchor=south east] at (r2) {$r_{k+1}$};
\node[anchor=north] at (g1) {$\gamma_k$};
\node[anchor=south east] at (g2) {$\gamma_{k+1}$};
\end{tikzpicture}
\end{fullfigure}
We require that the other endpoint $r_k$ of $\gamma_k$ travel all the way
around $\partial D$ in the counterclockwise direction, as in \fig{f:sweep}.
\begin{fullfigure}{f:sweep}{Making a sequence of geodesics that
sweep out $D$}
\begin{tikzpicture}
\begin{scope}[x={(.75cm,0)},y={(.375cm,.659cm)}]
\draw[darkred,semithick] (0,0) -- ++(0,1) -- ++(-1,1) -- ++(0,1);
\draw (1,3) -- ++(-4,0) -- ++(0,-2) -- ++(3,-3) -- ++(2,0) -- ++(0,1)
-- ++(1,-1) -- ++(0,2) -- ++(-1,1) -- ++(0,1) -- cycle;
\coordinate (r) at (-1,3);
\coordinate (g) at (0,1);
\coordinate (d) at (2,0);
\end{scope}
\fill[darkred] (0,0) circle (.07);
\fill[darkred] (r) circle (.07);
\node[anchor=west] at (0,-.1) {$p$};
\node[anchor=north west] at (r) {$r_k$};
\node[anchor=west] at (g) {$\gamma_k$};
\node at (d) {$D$};
\draw[thick,->] (100:.6) arc (100:300:.6);
\end{tikzpicture}
\end{fullfigure}
If $p$ is on the boundary, then $r_0 = p$, but this is okay. It is easy to
see that if ${\vec{\gamma}}$ exists, then it uses every vertex in $D$. There is
thus a geodesic $\gamma$ from $p$ to $r \in \partial D$ that contains $q$.
We can then repeat the argument with $r$ replacing $p$, to obtain a
geodesic $\gamma'$ from $r$ to some $s \in \partial D$ that contains $p$.
The geodesic $\gamma'$ might not contain $q$, much less all of $\kappa$.
However, because $D$ is geodesically coherent, the path
$$\gamma'' = \gamma'(s,p) \sqcup \kappa \sqcup \gamma(q,r)$$
is a geodesic and satisfies the lemma, as in \fig{f:replace}.
\begin{fullfigure}{f:replace}{A geodesic replacement argument}
\begin{tikzpicture}
\begin{scope}[x={(.75cm,0)},y={(.375cm,.659cm)}]
\draw[semithick,darkred] (0,0) -- ++(-2,2) -- ++(-1,0);
\draw[semithick,darkred] (0,0) -- ++(-1,0) -- ++(-2,2) -- ++(-1,0)
-- ++(0,-1) -- ++ (-1,0);
\draw[semithick,darkred] (-5,1) -- ++(1,-1) -- ++(2,0) -- ++(1,-1) -- ++(1,0)
-- ++(0,1) -- ++(2,0);
\draw[semithick,darkgreen] (-5.2,1.2) -- ++(1,0) -- ++(0,1) -- ++(2.2,0)
-- ++(2,-2) -- ++(2,0);
\draw (-6,2) -- (-5,1) -- (-5,0);
\draw (2,-1) -- (2,1);
\coordinate (p) at (0,0); \coordinate (q) at (-3,2);
\coordinate (r) at (-5,1); \coordinate (s) at (2,0);
\coordinate (k) at (-1.5,1.5); \coordinate (g) at (-2,1);
\coordinate (g1) at (-3,0); \coordinate (g2) at (-1,1.2);
\end{scope}
\fill[darkred] (p) circle (.07); \fill[darkred] (q) circle (.07);
\fill[darkred] (r) circle (.07); \fill[darkred] (s) circle (.07);
\node[anchor=north west] at (p) {$p$};
\node[anchor=north east] at (q) {$q$};
\node[anchor=east] at (r) {$r$\;};
\node[anchor=west] at (s) {\;$s$};
\node[anchor=north east] at (k) {$\kappa$};
\node[anchor=north east] at (g) {$\gamma$};
\node[anchor=north] at (g1) {$\gamma'$};
\node[anchor=south west] at (g2) {$\gamma''$};
\end{tikzpicture}
\end{fullfigure}
To prove the claim, let $\gamma_0$ be the geodesic of length $0$ if $p \in
\partial D$, and otherwise let $\gamma_0$ be the geodesic from $p$ to any
$r_0 \in \partial D$ which is counterclockwise extremal. We construct
${\vec{\gamma}}$ iteratively. Given $\gamma_k$, we apply a diamond move to
make $\gamma_{k+1}$ if such a move is possible. If such a move is not
possible, then let $r_{k+1}$ be the next boundary vertex after $r_k$,
and let $\gamma_{k+1}$ be the clockwise-extremal geodesic from $p$
to $r_{k+1}$, among geodesics that do not cross $\gamma_k$. (In other
words, cut $D$ along $\gamma_k$ to make $D'$, then let $\gamma_{k+1}$
be clockwise-extremal in $D'$.) By geodesic coherence, the region between
$\gamma_k$ and $\gamma_{k+1}$ is either empty or connected; otherwise
we could splice $\gamma_{k+1}$ with $\gamma_k$, so that $\gamma_{k+1}$ would
not be clockwise-extremal.
If the region between $\gamma_k$ and $\gamma_{k+1}$ is empty, then either
$\gamma_k \subseteq \gamma_{k+1}$ or $\gamma_{k+1} \subseteq \gamma_k$.
If it is not empty, then there are two geodesic segments $\gamma_k(s,r_k)$
and $\gamma_{k+1}(s,r_{k+1})$ make a topological triangle $T$ together
with the edge $(r_k,r_{k+1})$, as in \fig{f:toptriangle}.
\begin{fullfigure}{f:toptriangle}{A topological triangle $T$ made
from geodesics}
\begin{tikzpicture}
\draw (30:1) -- (150:1);
\draw[darkred,semithick] (150:1) -- (210:.3)
-- (270:1.5) -- (330:.3) -- (30:1);
\fill[darkred] (30:1) circle (.07); \fill[darkred] (150:1) circle (.07);
\fill[darkred] (270:1.5) circle (.07);
\node[above right] at (30:1) {$r_k$};
\node[above left] at (150:1) {$r_{k+1}$};
\node[below=.5ex] at (270:1.5) {$s$};
\node[below left] at (210:.3) {$\gamma_{k+1}$};
\node[below right] at (330:.3) {$\gamma_k$};
\node at (0,.1) {$T$};
\end{tikzpicture}
\end{fullfigure}
We summarize the properties of the topological triangle $T$: It is
$\mathrm{CAT}(0)$, all three sides are concave, and its angles at the corners are
at least $\pi/3$. Thus $T$ is flat, all three sides are flat (unlike in
the figure), and all three angles equal $\pi/3$. Thus, $T$ is a face of
$D$ and $\gamma_k$ and $\gamma_{k+1}$ differ by a triangle move.
As $k$ increases, eventually $r_k = r_0$. Once the diamond moves are
exhausted for this choice of $r_k$ (there are none if $p$ is on the
boundary), the sequence of geodesics returns to the beginning.
\end{proof}
The sweep-out construction in the proof of \lem{l:complete} also
yields this lemma.
\begin{lemma} Let $D$ be a $\mathrm{CAT}(0)$ diskoid with a boundary vertex $p$.
Then every edge of $D$ either lies on a complete geodesic from $p$ to
some $q \in \partial D$, or it lies in a diamond move or a triangle move
between two geodesics from $p$.
\label{l:everyedge} \end{lemma}
Finally, there is a relation between fans as described in \sec{s:paths}
and non-elliptic webs. Given a diskoid $D$ with boundary ${\vec{\lambda}}$, let
${\vec{\mu}}(D)$ be the sequence of distances $d(p,q_k)$, where $p$ is the base
point of $D$ and $q_k$ is the sequence of boundary vertices of $D$. Then:
\begin{theorem} \cite{Kuperberg:spiders} Given a sequence of $A_2$
minuscule weights ${\vec{\lambda}}$, the map $D \mapsto {\vec{\mu}}(D)$ is a bijection
between $\mathrm{CAT}(0)$ diskoids and minuscule paths of type ${\vec{\lambda}}$.
\end{theorem}
So we can write $D({\vec{\lambda}},{\vec{\mu}})$ as the non-elliptic web with boundary
${\vec{\lambda}}$ and minuscule path ${\vec{\mu}}$.
\subsection{Unitriangularity}
\label{s:unitriang}
We apply \sec{s:geodesics} to prove the following result. It is a bridge
result, based on the geometry of affine buildings, that we will use to
relate web bases to the geometric Satake correspondence; in particular,
to prove \thm{th:unitriang}.
\begin{theorem} Let ${\vec{\lambda}}$ be a minuscule sequence of type $A_2$
and let ${\vec{\mu}}$ be a minuscule path of type ${\vec{\lambda}}$. If $f \in
Q(A({\vec{\lambda}},{\vec{\mu}}))$ is a fan configuration, then it extends
uniquely to a diskoid configuration $f \in Q(D({\vec{\lambda}},{\vec{\mu}}))$.
\label{th:extend} \end{theorem}
\begin{proof} The construction derives from the constraints that make
the extension unique. Let $p$ be the base vertex of $D$, so that $f(p)
= 0 \in \mathrm{Gr}$. Suppose that $q$ is the $k$th boundary vertex of $D$, and
that $\gamma$ is a geodesic from $p$ to $q$. Then $d(f(p),f(q)) = \mu_k$,
and by definition $\mu_k$ is also the length of $\gamma$. If $\Sigma$ is
an apartment containing $f(p)$ and $f(q)$, then $f(q) = \mu_k$ in suitable
coordinates in $\Sigma$. It follows that there is a unique geodesic in
$\Sigma$ with the same sequence of edge weights as $\gamma$, and which
connects $f(p)$ with $f(q)$. Thus $f$ extends uniquely to $\gamma$.
We claim that this extension of $f$ is consistent for vertices of $D$. First,
every vertex of $D$ is contained in some complete geodesic from $p$ since
by \lem{l:complete} any geodesic from $p$ to a vertex extends to a
complete geodesic. Suppose that $\gamma$ and $\gamma'$ are two geodesics
from $p$ to $q \in \partial D$ and $q' \in \partial D$, respectively.
Suppose further that $r \in \gamma \cap \gamma'$. Then every apartment
that contains $p$ and $r$ contains both geodesics $\gamma(p,r)$ and
$\gamma'(p,r)$. In particular, each apartment $\Sigma \supseteq \gamma$
and $\Sigma' \supseteq \gamma'$ does. It follows that the choices for
$f(r)$ induced by $\gamma$ and $\gamma'$ are the same.
We claim that if $(r,s)$ is an edge in $D$, then
\eq{e:edge}{d(r,s) = d(f(r),f(s)).}
By \lem{l:everyedge}, there are three cases: Either $(r,s)$ occurs
in a complete geodesic from $p$ to some $q$, or it occurs in a diamond
move between two such geodesics $\gamma$ and $\gamma'$, or $r$ and $s$
are both on the boundary and $(r,s)$ occurs in a triangle move between
two geodesics $\gamma$ and $\gamma'$. In the first case, \eqref{e:edge}
is true by construction. In the second case, $f(\gamma)$ and $f(\gamma')$
are contained in a single apartment, because every apartment contains all
geodesics from $f(p)$ to $f(q)$. In the third case, there is an apartment
containing $p$ and $(r,s)$ by the axioms for a building, since they are
both simplices. In both cases, the existence of this common apartment
implies \eqref{e:edge}.
\end{proof}
Now let ${\vec{\lambda}}$ be a minuscule dominant sequence, and let ${\vec{\mu}}$ be
a minuscule path of type ${\vec{\lambda}}$. Then there is a corresponding
non-elliptic web $w({\vec{\lambda}}, {\vec{\mu}})$ with dual diskoid $D({\vec{\lambda}}, {\vec{\mu}})$.
There is also a corresponding component $\bar{Q(A({\vec{\lambda}}, {\vec{\mu}}))}$
of $F({\vec{\lambda}})$.
We have two bases for $H_\top(F({\vec{\lambda}}))$, one given by
$[\bar{Q(A({\vec{\lambda}}, {\vec{\mu}}))}]$ and the other given by $\Psi(w({\vec{\lambda}},
{\vec{\mu}}))$, and both bases are indexed by the minuscule path ${\vec{\mu}}$.
Under the isomorphism
$$H_\top(F({\vec{\lambda}})) \cong \mathrm{Inv}(V({\vec{\lambda}})),$$
these become the Satake and web bases, respectively, the first by definition
and the second by \thm{th:homclass2}. Our purpose in this section is to
prove that the transition matrix between these two basis is unitriangular.
Define a partial order on minuscule paths by the rule that
${\vec{\nu}}\leq{\vec{\mu}}$ when $\nu_i\leq \mu_i$ for all $i$.
\begin{theorem} The transition matrix between the Satake and web bases is
unitriangular with respect to the partial order $\leq$.
\label{th:unitriangweak} \end{theorem}
In the next section, we will use this result to deduce \thm{th:unitriang},
which concerns a weaker partial order and is thus a stronger statement.
We divide the proof of \thm{th:unitriangweak} into the following two lemmas.
\begin{lemma} Suppose that ${\vec{\nu}} \not\le {\vec{\mu}}$. Then the coefficient of
$[\bar{Q(A({\vec{\lambda}}, {\vec{\nu}}))}]$ in $\Psi(w({\vec{\lambda}}, {\vec{\mu}}))$ is 0.
\label{l:coeff2} \end{lemma}
\begin{proof}
By \cor{c:support}, it suffices to show that if $Q(A({\vec{\lambda}},
{\vec{\nu}}))$ is contained in $\pi(Q(D({\vec{\lambda}}, {\vec{\mu}})))$, then ${\vec{\nu}} \le {\vec{\mu}}$.
Let $f\in Q(D({\vec{\lambda}}, {\vec{\mu}}))$. If $q_i$ is the $i$th boundary vertex of
the diskoid $D({\vec{\lambda}}, {\vec{\mu}})$, then $f(q_i) \in \bar{\mathrm{Gr}(\mu_i)}$.
On the other hand, if $\pi(f) \in Q(A({\vec{\lambda}}, {\vec{\nu}}))$, then $f(q_i)
\in \mathrm{Gr}(\nu_i)$ . Thus $\nu_i \le \mu_i$ for all $i$ as desired.
\end{proof}
\begin{lemma} The coefficient of $[\bar{Q(A({\vec{\lambda}}, {\vec{\mu}}))}]$
in $\Psi(w({\vec{\lambda}}, {\vec{\mu}}))$ is 1.
\label{l:coeff1} \end{lemma}
\begin{proof}
Let $Z = \bar{\pi^{-1}(Q(A({\vec{\lambda}}, {\vec{\mu}}))}$. Then $Z$ is a component
of $Q(D({\vec{\lambda}}, {\vec{\mu}}))$, and it has dimension $d({\vec{\lambda}})$ by
\thm{th:extend}. Recall that from \thm{th:homclass2}, that we have a
homology class $c(w) \in H_{d({\vec{\lambda}})}(Q(D))$ such that $\pi_*(c(w))
= \Psi(w)$. Using the notation of the proof of \thm{th:homclass2},
$$Q(D({\vec{\lambda}}, {\vec{\mu}})) =
\pi_{0,1}^{-1}(X_1) \cap \cdots \cap \pi_{n-1, n}^{-1}(X_n)$$
and
$$c(w) = [\pi_{0,1}^{-1}(X_1)] \cap \cdots \cap [\pi_{n-1,n}^{-1}(X_n)]$$
Since $Z$ is a component of the expected dimension, we see that
the coefficient of $[Z]$ in $c(w)$ is the length of the local ring
of $Q(D({\vec{\lambda}}, {\vec{\mu}}))$ along $Z$ (by \cite{Fulton:intersect},
Proposition 8.2). This length equals 1 since following lemma shows that
the scheme $\pi^{-1}(Q(A({\vec{\lambda}}, {\vec{\mu}})))$ is isomorphic to the reduced
scheme $Q(A({\vec{\lambda}}, {\vec{\mu}}))$.
The degree $\pi|_Z$ is 1, so $\pi_*([Z]) = [\bar{Q(A({\vec{\lambda}},
{\vec{\mu}}))}]$. Moreover, $Z$ is the only component of $Q(D({\vec{\lambda}}, {\vec{\mu}}))$
which maps onto $\bar{Q(A({\vec{\lambda}}, {\vec{\mu}}))}$, so we conclude that the
coefficient of $[\bar{Q(A({\vec{\lambda}}, {\vec{\mu}}))}]$ in $\pi_*c(w)$ is also 1,
as desired.
\end{proof}
\begin{lemma}
The restriction of the map $\pi: Q(D({\vec{\lambda}}, {\vec{\mu}})) \to F({\vec{\lambda}})$
to $\pi^{-1}(Q(A({\vec{\lambda}}, {\vec{\mu}}))$ is an isomorphism of schemes onto the
reduced scheme $Q(A({\vec{\lambda}}, {\vec{\mu}}))$.
\end{lemma}
\begin{proof}
First note that $Q(A({\vec{\lambda}}, {\vec{\mu}}))$ is reduced since it is isomorphic to a
iterated fibred product of varieties by the proof of \thm{th:haines}.
Let $X = \pi^{-1}(Q(A({\vec{\lambda}}, {\vec{\mu}})), Y = Q(A({\vec{\lambda}}, {\vec{\mu}}))$. We have
already shown in \thm{th:extend} that the map $\pi: X \to
Y$ gives a bijection at the $\mathbb{C}$-points. Now, let $S$ be any scheme of
finite-type over $\mathbb{C}$. The proof of \thm{th:extend} uses some
building-theoretic arguments which don't obviously work for $S$-points.
However, the argument in the first paragraph of the proof does work
for any $S$, as follows. Following the notation in that paragraph, let
$\gamma$ be a geodesic in $\Gamma$ from the base point $p$ of $D({\vec{\lambda}},
{\vec{\mu}})$ to the $k$-th boundary vertex $q$ and let ${\vec{\nu}}$ be the lengths
along this geodesic (by definition $\sum \nu_i = \mu_k$). Let $f \in
X(S)$. Then the restriction of the map $m: \mathrm{Gr}({\vec{\nu}}) \to \mathrm{Gr}$
to $m^{-1}(\mathrm{Gr}(\mu_k))$ is an isomorphism of schemes, and in particular
is an injection on $S$-points. Hence we see that $f(r)$ is determined by
$f(q)$ for all $r$ along the geodesic. Since every internal vertex of the
diskoid lies on some geodesic, $f \in X(S)$ is determined by its restriction
to the boundary. Thus, the map $X(S) \to Y(S)$ is injective.
So we have a map from a scheme to a smooth variety which is a bijection
on $\mathbb{C}$-points and is an injection on $S$-points. By the following lemma,
the map is an isomorphism.
\end{proof}
\begin{lemma} Let $X, Y$ be finite-type schemes over $\mathbb{C}$. Assume that $Y$
is reduced and normal. Let $\phi: X \to Y$ be a morphism which
induces a bijection on $\mathbb{C}$-points and an injection on $S$-points for all
finite-type $\mathbb{C}$-schemes $S$. Then $\phi$ is an isomorphism.
\end{lemma}
\begin{proof} Consider the maps
$$X_\mathrm{red} \to X \to Y.$$
The composition $X_\mathrm{red} \to Y$ is a bijection on $\mathbb{C}$-points and
hence it is an isomorphism \cite[Thm. A.11]{Kumar:kacmoody}. This allows
us to construct a map $\psi:Y \to X$ such that $\phi \psi = id_Y$.
The fact that $\phi$ induces an injection on $S$-points means that the map
$$\mathrm{Hom}_{\mathrm{Sch}}(X, X) \xrightarrow{\phi \circ} \mathrm{Hom}_{\mathrm{Sch}}(X, Y)$$
is injective. Consider what happens to $id_X$ and $\psi\phi$ under
this map. They are sent to $\phi$ and $\phi\psi\phi$ respectively.
But since $\phi\psi = \mathrm{id}_Y$, these two elements of $\mathrm{Hom}_{\mathrm{Sch}}(X,
Y)$ are equal. Hence by the injectivity, $\mathrm{id}_X = \psi \phi$
and hence $\phi$ is an isomorphism.
\end{proof}
\subsection{Consequences of the cyclic action}
\label{s:cyclic}
The goal of this section is to prove \thm{th:main} and then derive some
corollaries. The proof is based on \thm{th:extend}. However, we first
need to understand the cyclic action on webs and Satake fibres, \emph{i.e.}, the
action that results from changing the base point of a polygon or a diskoid.
Fix a minuscule sequence ${\vec{\lambda}} =(\lambda_1, \dots, \lambda_n)$ and
consider the corresponding Satake fibre $F({\vec{\lambda}})$. Also regard the
indices of the sequence ${\vec{\lambda}}$ as lying in $\mathbb{Z}/n$. For each $i \in
\mathbb{Z}/n$, we define
$${\vec{\lambda}}^{(i)} = (\lambda_{i+1}, \lambda_{i+2}, \dots, \lambda_{n-1},
\lambda_0, \lambda_1, \dots, \lambda_i)$$
to be the $i$th cyclic permutation of ${\vec{\lambda}}$ (so that ${\vec{\lambda}}^{(0)}
= {\vec{\lambda}}$).
Let $Z$ be an irreducible component of $F({\vec{\lambda}})$. Since $G(\mathcal{O})$ is
connected and acts on $F({\vec{\lambda}})$, we see that $Z$ is $G(\mathcal{O})$-invariant.
Define
$$Z_1 = \{([g_1^{-1} g_2], \dots, [g_1^{-1} g_n], t^0)
| ([g_1], \dots, [g_n]) \in Z \subset F({\vec{\lambda}}^{(1)}) \},$$
and by iteration define $Z_i \subseteq F({\vec{\lambda}}^{(i)})$ for all $i \in \mathbb{Z}/n$.
This yields a bijection
$$\mathrm{Irr}(F({\vec{\lambda}})) \cong \mathrm{Irr}(F({\vec{\lambda}}^{(1)}))$$
which we call geometric rotation of components.
Another way to think about $Z_i$ is to think about the unbased configuration
space of $P({\vec{\lambda}})$ and to note that it fibres over $\mathrm{Gr}$ in $n$
different ways, by choosing each of the $n$ vertices of $P({\vec{\lambda}})$
as the base point. (But the geometry of these fibrations is subtle,
because the fibres do not have to be isomorphic algebraic varieties.)
A straightforward calculation in convolution algebras (in which all
intersections are transverse), shows that geometric rotation matches
pivotal rotation in $\mathrm{\mathbf{hconv}}(\mathrm{Gr})$. At the same time, \thm{th:satbasis}
tells us that the diagram
\begin{equation} \label{e:rotation}
\begin{tikzpicture}[description/.style={fill=white,inner sep=2pt},baseline]
\matrix (m) [matrix of math nodes, row sep=3em,
column sep=2.5em, text height=1.5ex, text depth=0.25ex]
{ H_\top(F({\vec{\lambda}})) & \mathrm{Inv}(V({\vec{\lambda}})) \\
H_\top(F({\vec{\lambda}}^{(1)})) & \mathrm{Inv}(V({\vec{\lambda}}^{(1)})) \\ };
\path[->] (m-1-1) edge node[auto] {$\Phi$} (m-1-2)
(m-1-1) edge node[auto] {$R$} (m-2-1)
(m-2-1) edge node[auto] {$\Phi$} (m-2-2)
(m-1-2) edge node[auto] {$R$} (m-2-2);
\end{tikzpicture}
\end{equation}
commutes, where the invariant spaces on the right are in $\mathrm{\mathbf{rep}}(G)^u_\min$.
By \thm{th:haines}, $Z = \bar{Q(A({\vec{\lambda}},{\vec{\mu}}))}$ for some minuscule path
${\vec{\mu}}$ of type ${\vec{\lambda}}$. From $({\vec{\lambda}},{\vec{\mu}})$, we obtain a diskoid $D
= D({\vec{\lambda}}, {\vec{\mu}})$. In $D$, the distances from the base point to the
other boundary vertices are given by ${\vec{\mu}}$. Now for each $i \in \mathbb{Z}/n$,
let ${\vec{\mu}}^{(i)}$ denote the sequence of distances from the $i$th boundary
vertex to the rest of the boundary. Since a rotated $\mathrm{CAT}(0)$ diskoid is
still a $\mathrm{CAT}(0)$ diskoid, we see that $D = D({\vec{\lambda}}^{(i)}, {\vec{\mu}}^{(i)})$
as well.
\begin{lemma}
For each $i$, $Z_i = \bar{Q(A({\vec{\lambda}}^{(i)}, {\vec{\mu}}^{(i)}))}$.
\label{l:cyclic} \end{lemma}
Although this lemma may look purely formal, it is (as far as we know)
a non-trivial identification of two different cyclic actions. The cyclic
action used to define $Z_i$ is defined directly from the geometric Satake
correspondence; it comes from the fact that the unbased configuration
space of $P({\vec{\lambda}})$ fibres over $\mathrm{Gr}$ in more than one way. The cyclic
action on the right, in particular the definition of ${\vec{\mu}}^{(i)}$, comes
instead from rotating webs. The two cyclic actions ``should be" the same
because the diagram analogous to \eqref{e:rotation} for webs commutes
(since $\mathrm{\mathbf{spd}}(\mathrm{SL}(3))$ is equivalent to $\mathrm{\mathbf{rep}}^u(\mathrm{SL}(3))$). However, the
lemma is non-trivial because it is not true that the invariant vector
$\Psi(w({\vec{\lambda}}, {\vec{\mu}}))$ coming from the web equals the fundamental class
of the corresponding component.
\begin{proof} Our proof uses \thm{th:unitriangweak}, the unitriangularity
theorem. Let $M$ be the unitriangular change of basis matrix; the rows
of $M$ are labelled by the web basis, while the columns are indexed by the
geometric Satake basis. Since both bases are cyclically invariant as in
the diagram \eqref{e:rotation}, there is a combinatorial cyclic action on
the rows and columns of $M$ that takes $M$ to itself.
Suppose for the moment that $M$ is an abstractly unitriangular matrix whose
rows and columns are labelled by two sets $A$ and $B$. In other words,
there exists an unspecified bijection $A \cong B$, and a linear or partial
order of $A$ that makes $M$ unitriangular. Then the partial order may
not be unique, but the bijection is. If we choose any compatible linear
order, then it is easy to see that the expansion of $\det M$ has only
one non-zero term. This term selects the unique compatible bijection.
Since it is unique, it intertwines the two cyclic actions in our case.
\end{proof}
Say that ${\vec{\nu}} \le_S {\vec{\mu}}$ when ${\vec{\nu}}^{(i)} \le {\vec{\mu}}^{(i)}$ for all $i \in
\mathbb{Z}/n$. If $D$ and $E$ are the diskoids of $w({\vec{\nu}})$ and $w({\vec{\mu}})$, then
this condition says that $d_D(p,q) \le d_E(p,q)$ for every two vertices
on their common boundary. \thm{th:unitriang} follows by combining
\thm{th:unitriangweak} with \lem{l:cyclic}.
We define a subset $U \subseteq Z$ as follows:
$$U = \{ (L_i)_{i \in \mathbb{Z}/n} \in F({\vec{\lambda}}) | d(L_i, L_j) = \mu^{(i)}_j \}.$$
\lem{l:cyclic} shows that $U$ is dense in $Z$. The following proposition
then completes the proof of \thm{th:main}.
\begin{proposition} Restricting the configuration to the boundary gives
an isomorphism
$$\pi: Q_g(D) \stackrel{\cong}{\longrightarrow} U.$$
\end{proposition}
\begin{proof} By definition, $U$ consists of those configurations of $D$
that preserve all distances between boundary vertices. By \lem{l:complete},
these are exactly the configurations that preserve all distances in $D$.
\end{proof}
If $f \in Q_g(D)$ is a global isometry, then in particular it is an embedding
of $D$ into the affine building $\Delta$. This has an interesting area
consequence.
\begin{lemma} Let $K$ be a 2-dimensional simplicial complex with trivial
homology, $H_*(K,\mathbb{Z}) = H_*({\mathrm{pt}})$. Then every simplicial 1-cycle $\alpha$
in $K$ is the homology boundary of a unique 2-chain $\beta$.
\label{l:homol} \end{lemma}
\begin{proof} If $\beta_1$ and $\beta_2$ are two such 2-chains, then
$\beta_1-\beta_2$ is closed and therefore null-homologous. Since $K$
has no 3-simplices, the only way for $\beta_1$ and $\beta_2$ to be
homologous is if they are equal.
\end{proof}
\begin{theorem} If a $\mathrm{CAT}(0)$, type $A_2$ diskoid $D$ is embedded in an
affine building $\Delta$, then it is the unique least area diskoid that
extends the embedding of its boundary $P$.
\end{theorem}
\begin{proof} Let $f$ be the embedding. Then $f_*([D])$ is a 2-chain whose
1-norm is the area of $D$. If $f':D' \to \Delta$ is another extension of
$P$, then $f'_*([D']) = f_*([D])$ and the area of $D'$ cannot be smaller
than the area of $D$. Moreover, if they have equal area, then $f^{-1}
\circ f$ is a bijection between the faces of $D'$ and the faces of $D$.
The faces of $D'$ must be connected in the same way as those of $D$, and
attached to $P$ in the same way, because each edge in $\Delta$ has at most
two faces of $f(D)$.
\end{proof}
By contrast, the $A_2$ spider relations \eqref{e:a2spider} reduce the
area of a diskoid. The following proposition is easy to check, as well
as inevitable given \prop{l:homol} and \thm{th:main}:
\begin{proposition} If $w$ is a web with a face with 2 or 4 sides, so
that the dual diskoid $D$ has a vertex with 2 or 4 triangles, then in
any configuration $f:D \to \mathrm{Gr}$ these triangles land on top of each other
in pairs.
\label{p:ontop} \end{proposition}
\prop{p:ontop} thus motivates the relations \eqref{e:a2spider}
as moves that locally remove area from a configuration $f$.
\subsection{Web bases are not Satake}
\label{s:notsatake}
In \sec{s:unitriang}, we showed that the transformation between the
web basis and the Satake basis is unitriangular with respect to the
given order. Thus it is reasonable to ask if this transformation is
the identity. As with Lustzig's dual canonical basis, there is an early
agreement between the two. For any web with no internal faces, that is,
whose dual diskoid has no internal vertices, the image of the map $\pi$
is $\bar{Q(A({\vec{\lambda}}, {\vec{\mu}}))}$ by \thm{th:smooth}, and $\pi$ is injective.
It follows from \cor{c:support} and \lem{l:coeff1} that $[\bar{Q(A({\vec{\lambda}},
{\vec{\mu}}))}]$ is the web vector.
Now consider the following web $w({\vec{\mu}})$, with the indicated base point:
$$w({\vec{\mu}}) = \begin{tikzpicture}[scale=0.4,baseline]
\foreach \angle in {0,120,240}
{
\begin{scope}[rotate=\angle,web]
\foreach \x/\y in {-3.5/-.866,-3.5/.866,-2/-1.732,-2/0,-2/1.732,
-.5/-2.598,-.5/-.866,-.5/.866,-.5/2.598,1/-1.732,1/0,1/1.732,
2.5/-.866,2.5/.866}
\draw[midto] (\x,\y) -- ++(1,0);
\end{scope}
}
\draw (0,0) circle (3.605);
\fill (240:3.605) circle (.1);
\end{tikzpicture}$$
In \cite{Kuperberg:notdual}, it was shown that this is the first web whose
invariant vector is not dual canonical. This is the web associated with
the minuscule path
\begin{multline*}
{\vec{\mu}} = (0, \omega_1, \omega_1+\omega_2, \omega_1+2\omega_2, 3\omega_2,
\omega_1+3\omega_2, \\
2\omega_1+2\omega_2, 3\omega_1+\omega_2, 3\omega_1,
2\omega_1+\omega_2, \omega_1+\omega_2, \omega_2, 0)
\end{multline*}
of type
$$ {\vec{\lambda}} = (\omega_1, \omega_2, \omega_2, \omega_1, \omega_1, \omega_2,
\omega_2, \omega_1, \omega_1, \omega_2, \omega_2, \omega_1).
$$
Let
$${\vec{\nu}}=(0,\omega_1,0,\omega_2,0,\omega_1,0,\omega_2,0,\omega_1,0,\omega_2,0).$$
This is another minuscule path also of type ${\vec{\lambda}}$; the
corresponding web $w({\vec{\nu}})$ is much simpler and is both a Satake vector
and a dual canonical vector:
$$w({\vec{\nu}}) = \begin{tikzpicture}[scale=0.4,baseline]
\foreach \angle in {0,120,240}
{
\begin{scope}[rotate=\angle,web]
\draw[midto] (1,3.464) .. controls (.5,2.598)
and (-.5,2.598) .. (-1,3.464);
\draw[midto] (1,-3.464) .. controls (.5,-2.598)
and (-.5,-2.598) .. (-1,-3.464);
\end{scope}
}
\draw (0,0) circle (3.605);
\fill (240:3.605) circle (.1);
\end{tikzpicture}$$
In \cite{Kuperberg:notdual}, it was shown that
$$\Psi(w({\vec{\mu}})) = b({\vec{\mu}}) + b({\vec{\nu}}),$$
where $b({\vec{\mu}})$ denotes the dual canonical basis vector
indexed by ${\vec{\mu}}$.
\begin{theorem} Let $w({\vec{\mu}})$, ${\vec{\lambda}}$, ${\vec{\mu}}$, and ${\vec{\nu}}$ be as
above. Then the invariant vector $\Psi(w({\vec{\mu}}))$ is not in the Satake
basis. More precisely, it has a coefficient of 2 for the basis vector
$[\bar{Q(A({\vec{\lambda}}, {\vec{\nu}}))}]$.
\end{theorem}
\begin{proof}
We will show that the general fibre of $\pi$ over $Q(A({\vec{\lambda}}, {\vec{\nu}}))$
is of size 2. We give the faces of the web the following labels:
$$\begin{tikzpicture}[scale=0.6]
\foreach \angle in {0,120,240}
{
\begin{scope}[rotate=\angle,web]
\foreach \x/\y in {-3.5/-.866,-3.5/.866,-2/-1.732,-2/0,-2/1.732,
-.5/-2.598,-.5/-.866,-.5/.866,-.5/2.598,1/-1.732,1/0,1/1.732,
2.5/-.866,2.5/.866}
\draw[midto] (\x,\y) -- ++(1,0);
\end{scope}
}
\node at ($(240:2)+(180:2)$) {$p_1$};
\node at ($(180:2)+(120:2)$) {$\ell_1$};
\node at ($(120:2)+(60:2)$) {$p_2$};
\node at ($(60:2)+(0:2)$) {$\ell_2$};
\node at ($(0:2)+(300:2)$) {$p_3$};
\node at ($(300:2)+(240:2)$) {$\ell_3$};
\node at ($(240:1)+(180:1)$) {$\ell_1'$};
\node at ($(180:1)+(120:1)$) {$p_1'$};
\node at ($(120:1)+(60:1)$) {$\ell_2'$};
\node at ($(60:1)+(0:1)$) {$p_2'$};
\node at ($(0:1)+(300:1)$) {$\ell_3'$};
\node at ($(300:1)+(240:1)$) {$p_3'$};
\node at (0,0) {$c$};
\end{tikzpicture}$$
If $f\in Q(D({\vec{\lambda}}, {\vec{\mu}}))$ then $\pi(f)\in Q(A({\vec{\lambda}},{\vec{\nu}}))$ if
and only if $f$ assigns $p_i\in \mathrm{Gr}(\omega_1)$ and $\ell_i\in \mathrm{Gr}(\omega_2)$
on those faces and assigns $t^0 \in \mathrm{Gr}(0)$ to all empty faces. In order
to determine the fibre of $\pi$ over a point in $Q(A({\vec{\lambda}},{\vec{\nu}}))$
we must calculate the possible choices for $p_i'$, $\ell_i'$ and $c$
satisfying the appropriate conditions. Since $p_i\in \mathrm{Gr}(\omega_1)$ and
$\ell_i\in \mathrm{Gr}(\omega_2)$, this forces $p_i'\in \mathrm{Gr}(\omega_1)$ and $\ell_i'\in
Gr(\omega_2)$ and $c\in\bar{\mathrm{Gr}(\omega_1+\omega_2)}$. We can think of the
points of $\mathrm{Gr}(\omega_1)$ and $\mathrm{Gr}(\omega_2)$ as, respectively, the points
and lines in $\mathbb{CP}^2$. Then the conditions given by the edges of
the web are as following: $p_i'$ is a point on the line $\ell_i$ and $\ell_i'$
is a line containing the points $p_i$, $p_{i-1}'$ and $p_i'$.
\begin{fullfigure*}{f:twosolutions}
{The two solutions to the problem for the given $\ell_i$ and $p_i$}
\subfloat[]{\begin{tikzpicture}[scale=0.55]
\useasboundingbox (-6,-5) rectangle (6,5);
\clip (-5,-5) rectangle (5,5);
\draw[domain=-5:5] plot(\x,{(-1-1.81*\x)/0.99});
\draw[domain=-5:5] plot(\x,{(--3.51--0.22*\x)/2.34});
\draw[domain=-5:5] plot(\x,{(-0.06-2.03*\x)/-1.35});
\draw (-2.56,4.5) node {$\ell_1$};
\draw (-4.5,1.55) node {$\ell_2$};
\draw (2.50,4.5) node {$\ell_3$};
\draw[blue,domain=-5:5] plot(\x,{(--2.34--0.37*\x)/3.3});
\draw[blue,domain=-5:5] plot(\x,{(--2.23-3.24*\x)/0.89});
\draw[blue,domain=-5:5] plot(\x,{(-1.51-0.91*\x)/-1.16});
\draw (4.5,0.75) node {$\ell'_1$};
\draw (-4.5,-1.65) node {$\ell'_2$};
\draw (-0.2,4.5) node {$\ell'_3$};
\fill[blue] (-0.89,0.61) circle (.1); \draw (-0.98,0.1) node {$p'_1$};
\fill[blue] (0.27,1.52) circle (.1); \draw (0.51,2.2) node {$p'_2$};
\fill[blue] (0.48,0.76) circle (.1); \draw (1.1,0.45) node {$p'_3$};
\fill[red] (-1.31,1.38) circle (.1); \draw (-1.00,1.7) node {$e_1$};
\fill[red] (1.04,1.59) circle (.1); \draw (1.75,2) node {$e_2$};
\fill[red] (-0.32,-0.43) circle (.1); \draw (0.2,-0.5) node {$e_3$};
\fill[darkgreen] (-4.19,0.24) circle (.1); \draw (-4.1,-0.24) node {$p_1$};
\fill[darkgreen] (-1.89,-0.16) circle (.1); \draw (-1.87,-0.7) node {$p_2$};
\fill[darkgreen] (1.37,-2.48) circle (.1); \draw (1.9,-2.5) node {$p_3$};
\end{tikzpicture}}
\subfloat[]{\begin{tikzpicture}[scale=0.55]
\useasboundingbox (-6,-5) rectangle (6,5);
\clip (-5,-5) rectangle (5,5);
\draw[domain=-5:5] plot(\x,{(-1-1.81*\x)/0.99});
\draw[domain=-5:5] plot(\x,{(--3.51--0.22*\x)/2.34});
\draw[domain=-5:5] plot(\x,{(-0.06-2.03*\x)/-1.35});
\draw (-2.56,4.5) node {$\ell_1$};
\draw (4.5,2.4) node {$\ell_2$};
\draw (2.50,4.5) node {$\ell_3$};
\draw[blue,domain=-5:5] plot(\x,{(-6.07-1.7*\x)/4.43});
\draw[blue,domain=-5:5] plot(\x,{(-3.34-1.4*\x)/2.12});
\draw[blue,domain=-5:5] plot(\x,{(-5.7-2.58*\x)/4.33});
\draw (4.5,-2.67) node {$\ell'_1$};
\draw (4.5,-3.6) node {$\ell'_2$};
\draw (3.7,-4.5) node {$\ell'_3$};
\fill[blue] (0.24,-1.46) circle (3pt); \draw (0.58,-1) node {$p'_1$};
\fill[blue] (-4.09,1.12) circle (3pt); \draw (-3.85,1.6) node {$p'_2$};
\fill[blue] (-0.75,-1.08) circle (3pt); \draw (-0.7,-1.7) node {$p'_3$};
\fill[red] (-1.31,1.38) circle (3pt); \draw (-1.00,1.7) node {$e_1$};
\fill[red] (1.04,1.59) circle (3pt); \draw (1.75,2) node {$e_2$};
\fill[red] (-0.32,-0.43) circle (3pt); \draw (0.2,-0.5) node {$e_3$};
\fill[darkgreen] (-4.19,0.24) circle (3pt); \draw (-4.1,-0.24) node {$p_1$};
\fill[darkgreen] (-1.89,-0.16) circle (3pt); \draw (-1.87,0.3) node {$p_2$};
\fill[darkgreen] (1.37,-2.48) circle (3pt); \draw (1.55,-3) node {$p_3$};
\end{tikzpicture}}
\end{fullfigure*}
Suppose that either the $p_i$ are not collinear and or the $\ell_i$ are not
concurrent. Then by the duality of points and lines, we may assume that
the $\ell_i$ are not concurrent. Let $e_i$ be the intersection of $\ell_i$ and
$\ell_{i+1}$. Then we can express the points $p_i'$ in barycentric coordinates
given by $e_i$:
\begin{align*}
p_1'=(t_1,0,1-t_1) \\ p_2'=(1-t_2,t_2,0) \\ p_3'=(0,1-t_3,t_3).
\end{align*}
Note that by doing this we restrict ourselves to an affine subspace of
$\P^2$, so we may lose, but we don't gain solutions. The collinearity
condition results in the equations
$$p_i=(1-s_i)p_i'+s_ip_{i-1}.$$
Solving this problem amounts to solving
\begin{align*}
(1-s_1)t_1 &= p_{11} & s_1(1-t_3) &= p_{12} \\
(1-s_2)t_2 &= p_{22} & s_2(1-t_1) &= p_{23} \\
(1-s_3)t_3 &= p_{33} & s_3(1-t_2) &= p_{31},
\end{align*}
where $p_{ij}$ are the barycentric coordinates of the $p_i$. If none of
these coordinates are $0$, then we can eliminate all but one variable to get
the relation
$$t_1=\frac{p_{11}}{1-\frac{p_{12}}{1-\frac{p_{33}}
{1-\frac{p_{31}}{1-\frac{p_{22}}{1-\frac{p_{23}}{1-t_1}}}}}}.$$
The right side of this equation is a composition of fractional
linear transformations that condenses to a single fractional linear
transformation
$$t_1 = \frac{\alpha_{11} t_1 + \alpha_{12}}{\alpha_{21}t_1 + \alpha_{22}}$$
with generic coefficients. Thus, generically,
we obtain a quadratic equation for $t_1$ with 2 solutions.
It remains to determine the face $c$, which lies in
$\mathrm{Gr}(\omega_1+\omega_2)$. If $c\not\in \mathrm{Gr}(0)$, then the conditions given
by the edges of the web would be $p_i'=p_j'$ and $\ell_i'=\ell_j'$ for all
$i,j$ which cannot happen since either $p_i$ are not collinear or $\ell_i$
are not concurrent. Thus for any solution of the above equations, we get
exactly one element in $Q(D({\vec{\lambda}}, {\vec{\mu}}))$. And for any generic point
$p \in Q(A({\vec{\lambda}}, {\vec{\nu}}))$, the fibre $\pi^{-1}(p)$ has 2 points.
Let $X$ denote the closure in $Q(D({\vec{\lambda}}, {\vec{\mu}}))$ of the union of all
fibres $\pi^{-1}(p)$ with 2 points. Then $X$ is either a component of
$Q(D({\vec{\lambda}}, {\vec{\mu}}))$ or a union of two components. Moreover, $X$ contains
all components of $Q(D({\vec{\lambda}}, {\vec{\mu}}))$ which map onto $Q(A({\vec{\lambda}},
{\vec{\nu}}))$. Since the above argument shows that the scheme-theoretic fibre of
$\pi$ over a general point of $Q(A({\vec{\lambda}}, {\vec{\nu}}))$ is two reduced points,
we also know that $X$ is generically reduced. Hence the coefficient of
$[X]$ in the homology class $c(w)$ from \thm{th:homclass2} is 1. Since the
map $\pi: X \to Q(A({\vec{\lambda}}, {\vec{\nu}}))$ is of degree 2 and since $X$
contains all components mapping to $Q(A({\vec{\lambda}}, {\vec{\nu}}))$, the coefficient of
$[Q(A({\vec{\lambda}}, {\vec{\nu}}))]$ in $\pi_*(c(w))$ is 2. In particular, $\pi_*(c(w))$
differs from $[\bar{Q(A({\vec{\lambda}}, {\vec{\mu}}))}]$, as desired.
\end{proof}
In fact, we suspect that $Q(D({\vec{\lambda}}, {\vec{\mu}}))$ only has two components,
which would imply that
$$\Psi(w({\vec{\mu}})) = [\bar{Q(A({\vec{\lambda}}, {\vec{\mu}}))}]
+ 2 [\bar{Q(A({\vec{\lambda}},{\vec{\nu}}))}].$$
Otherwise, $\Psi(w({\vec{\mu}}))$ has these two terms and perhaps others.
Either way, the coefficient of 2 is different from what arises in the dual
canonical basis \cite{Kuperberg:notdual}:
$$\Psi(w({\vec{\mu}})) = b({\vec{\mu}}) + b({\vec{\nu}}).$$
Thus,
\begin{theorem} The geometric Satake bases for invariants of $G = \mathrm{SL}(3)$
are eventually not dual canonical.
\label{th:notdual} \end{theorem}
This is not such a surprising statement in light of the well-known fact that
the canonical and semicanonical basis do not coincide (as a consequence of
the work of Kashiwara-Saito \cite{KS:crystal}). In both \thm{th:notdual}
and in the canonical/semicanonical situation, a homology basis does not
coincide with a basis defined using a bar-involution. The analogy between
these two results could perhaps be made precise using skew Howe duality
($\mathrm{SL}(3)$, $\mathrm{SL}(n)$-duality).
It is known that $\Psi(w({\vec{\mu}}))$ is the first basis web that is not
dual canonical, \emph{i.e.}, the only basis web up to rotation with 12 or fewer
minuscule tensor factors. We conjecture that it is also the first basis
web for $\mathrm{SL}(3)$ that is not geometric Satake. Equivalently, we conjecture
that all three bases first diverge at the same position.
\begin{question} For arbitrary $G$,
is the dual canonical basis of an invariant space $\mathrm{Inv}_G(V(\lambda))$
positive unitriangular in the geometric Satake basis?
\end{question}
\section{Euler convolution of constructible functions}
\label{s:euler}
In this section, we switch from convolution in homology to convolution
in constructible functions. The idea of defining convolution algebras
using constructible functions is common in geometric representation theory
(see for example \cite{Lusztig:constructible}).
More specifically, we will define a new category $\mathrm{\mathbf{econv}}(\mathrm{Gr})_0$ which
conjecturally is equivalent to $\mathrm{\mathbf{rep}}^u_{-1}(G)_\min$, and we will prove
this conjecture for $G = \mathrm{SL}(2)$ and $G = \mathrm{SL}(3)$. When computing invariant
vectors from webs, the construction is a state model as in \sec{s:hstate},
where the counting is done using Euler characteristic.
\subsection{Generalities on constructible functions}
If $X$ is a proper complex algebraic variety over $\mathbb{C}$ and $f:X \to \mathbb{C}$
is a constructible function, then we define the \emph{Euler characteristic
integral} (see \cite{MacPherson:chern} or \cite{Joyce:constructible})
$$\int_X f d\chi \in \mathbb{C}$$
by linear extension starting with the characteristic functions of closed
subvarieties. Namely, if $f = f_Y$ is the characteristic function of a
closed subvariety $Y \subseteq X$, then we define
$$\int_X f_Y d\chi \stackrel{\mathrm{def}}{=} \chi(Y).$$
If $\pi:X \to Y$ is a proper morphism between algebraic varieties and $f:X
\to \mathbb{C}$ is a constructible function on $X$, then we define the push-forward
of $f$ under $\pi$ by integration along fibres:
$$(\pi_*f)(p) \stackrel{\mathrm{def}}{=} \int_{\pi^{-1}(p)} f d\chi.$$
If $\C_c(X)$ denotes the vector space of constructible functions on $X$,
this pushforward is then a linear map:
$$\pi_*:\C_c(X) \to \C_c(Y).$$
The following result is well-known --- see for example
\cite[Prop. 1]{MacPherson:chern} and \cite[Thm. 3.8]{Joyce:constructible}.
\begin{theorem} The Euler characteristic integral push-forward of
constructible functions is a well-defined covariant functor from the
category of proper morphisms between algebraic varieties over $\mathbb{C}$, to
the category of complex vector spaces.
\label{th:efunctor} \end{theorem}
\subsection{Construction of the categories}
Given $G$ simple and simply-connected as before, we can define a pivotal
category $\mathrm{\mathbf{econv}}(\mathrm{Gr})$ in a similar fashion to $\mathrm{\mathbf{hconv}}(\mathrm{Gr})$, except that
we will replace homology with constructible functions throughout.
The objects of $\mathrm{\mathbf{econv}}(\mathrm{Gr})$ are the $\mathrm{Gr}({\vec{\lambda}})$, where ${\vec{\lambda}}$ is
a sequence of minuscule weights. As in $\mathrm{\mathbf{hconv}}(\mathrm{Gr})$, the tensor product
is defined by convolution.
We define the invariant space of $\mathrm{Gr}({\vec{\lambda}})$ to be the vector space of
constructible functions on the Satake fibre:
$$\mathrm{Inv}_{\mathrm{\mathbf{econv}}(\mathrm{Gr})}(\mathrm{Gr}({\vec{\lambda}})) \stackrel{\mathrm{def}}{=} \C_c(F({\vec{\lambda}})).$$
The hom spaces are defined in an equivalent way:
$$\mathrm{Hom}_{\mathrm{\mathbf{econv}}(\mathrm{Gr})}(\mathrm{Gr}({\vec{\lambda}}),\mathrm{Gr}({\vec{\mu}}))
\stackrel{\mathrm{def}}{=} \C_c(Z({\vec{\lambda}},{\vec{\mu}})).$$
We define the convolution of two hom spaces by convolution as in
$\mathrm{\mathbf{hconv}}(\mathrm{Gr})$. We could proceed exactly as in $\mathrm{\mathbf{hconv}}(\mathrm{Gr})$, but the
``local'' nature of constructible functions allows us a simpler definition.
Fix three minuscule sequences ${\vec{\lambda}}, {\vec{\mu}}, {\vec{\nu}}$. Let $\Gamma$ be a graph
homeomorphic to a theta ($\theta$) with three arcs that are polylines of
type ${\vec{\lambda}}$,${\vec{\mu}}$, and ${\vec{\nu}}$ with a common base point:
$$\Gamma \;\;=\;\; \begin{tikzpicture}[baseline]
\draw[darkred,semithick] (0,0) -- (1,1) -- (2,.5) -- (3,1) -- (3.5,0)
-- (3,-1) -- (1,-1) -- cycle;
\draw[darkred,semithick] (0,0) -- (1,0) -- (2.5,-.5) -- (3.5,0);
\draw[darkred,semithick,midto] (1,0) -- (2.5,-.5);
\draw[darkred,semithick,midto] (1,1) -- (2,.5);
\draw[darkred,semithick,midto] (1,-1) -- (3,-1);
\fill[darkred] (0,0) circle (.07);
\path[draw=darkred,fill=white] (3.5,0) circle (.07);
\draw[anchor=north] (1.75,-1.1) node {${\vec{\lambda}}$};
\draw[anchor=north] (1,-.1) node {${\vec{\mu}}$};
\draw[anchor=south] (1.75,.9) node {${\vec{\nu}}$};
\end{tikzpicture}$$
Then there are projections
\begin{align*}
\pi_{{\vec{\lambda}},{\vec{\mu}}}:Q(\Gamma) &\to Z({\vec{\lambda}},{\vec{\mu}}) \\
\pi_{{\vec{\lambda}},{\vec{\nu}}}:Q(\Gamma) &\to Z({\vec{\lambda}},{\vec{\nu}}) \\
\pi_{{\vec{\mu}},{\vec{\nu}}}:Q(\Gamma) &\to Z({\vec{\mu}},{\vec{\nu}})
\end{align*}
Given
\begin{align*}
f &\in \mathrm{Hom}_{\mathrm{\mathbf{econv}}(\mathrm{Gr})}(\mathrm{Gr}({\vec{\lambda}}),\mathrm{Gr}({\vec{\mu}})) \\
g &\in \mathrm{Hom}_{\mathrm{\mathbf{econv}}(\mathrm{Gr})}(\mathrm{Gr}({\vec{\mu}}),\mathrm{Gr}({\vec{\nu}})),
\end{align*}
we can define their composition by Euler characteristic integration over
configurations of the middle polyline $L({\vec{\mu}})$, and using the fact that
constructible functions pull back and multiply as well as push forward:
$$g \circ f \stackrel{\mathrm{def}}{=} (\pi_{{\vec{\lambda}},{\vec{\nu}}})_*(\pi_{{\vec{\lambda}},{\vec{\nu}}}^*(f)
\pi_{{\vec{\mu}},{\vec{\nu}}}^*(g)).$$
It is routine to check that these structures define a pivotal category.
The hom spaces in the category $\mathrm{\mathbf{econv}}(\mathrm{Gr})$ are too large for our purposes.
We will restrict them by just looking at those constructible functions
generated by the constant functions on the Satake fibres corresponding to
trivalent vertices. More precisely, define a pivotal functor
$$E:\mathrm{\mathbf{fsp}}(G) \to \mathrm{\mathbf{econv}}(\mathrm{Gr})$$
which takes the generating vertex in
$$\mathrm{Inv}_{\mathrm{\mathbf{fsp}}(G)}(\lambda,\mu,\nu),$$
to the identity function on $F(\lambda,\mu,\nu)$. Again, $\lambda$, $\mu$,
and $\nu$ are all minuscule and we are assuming that there is a vertex, so
$$\mathrm{Inv}_G(V(\lambda, \mu, \nu)) \ne 0.$$
Let $\mathrm{\mathbf{econv}}(\mathrm{Gr})_0$ denote the image of the functor $E$; it has
the same objects as $\mathrm{\mathbf{econv}}(\mathrm{Gr})$, but smaller hom spaces.
\subsection{Equivalence with the representation category}
Before stating the main conjecture and result, we can describe more
explicitly how the functor $E$ expresses an Euler characteristic state model.
The following result can be seen by chasing through the definitions.
\begin{proposition} Given a web $w \in \mathrm{\mathbf{fsp}}(G)$ with boundary ${\vec{\lambda}}$
and dual diskoid $D$, $E(w)$ is the function on the Satake fibre
$F({\vec{\lambda}})$ whose value at $p \in F({\vec{\lambda}})$ is $\chi(\pi^{-1}(p))$.
(Here $\pi:Q(D) \to F({\vec{\lambda}})$ is the map which restricts a diskoid
configuration to its boundary.)
\end{proposition}
So we are indeed producing a function which counts (using Euler
characteristic) ways to extend the boundary configuration to a diskoid
configuration.
We are now ready to formulate our alternate version of the geometric
Satake correspondence.
\begin{conjecture} There is an equivalence of pivotal categories:
$$\mathrm{\mathbf{econv}}(\mathrm{Gr}({G^\vee}))_0 \cong \mathrm{\mathbf{rep}}^u_{-1}(G)_\min.$$
\label{c:eulerrep} \end{conjecture}
Recall from Sections~\ref{s:signs} and \ref{s:examples} that
$\mathrm{\mathbf{rep}}^u_{-1}(G)_\min$ and $\mathrm{\mathbf{rep}}^u(G)_\min$ have the same information except
for a sign correction. We offer the following corollary of \conj{c:eulerrep}
as a stand-along conjecture.
\begin{conjecture} Let $w$ be any closed web with dual diskoid $D$. Then
$$\Psi(w) = \pm \chi(Q(D)).$$
\label{c:eulerclosed} \end{conjecture}
Here $\Psi(w)$ denotes the value of $w$ in the pivotal category
$\mathrm{\mathbf{rep}}^u(G)_\min$ and the sign comes as a result of the sign correction
between $\mathrm{\mathbf{rep}}^u_{-1}(G)_\min$ and $\mathrm{\mathbf{rep}}^u(G)_\min$.
\begin{theorem} \conj{c:eulerrep} holds when $G = \mathrm{SL}(2)$ and $G = \mathrm{SL}(3)$.
\end{theorem}
\begin{proof} We will first argue the more difficult case $G = \mathrm{SL}(3)$.
We argue by checking the skein relations of $\mathrm{\mathbf{spd}}_{-1}(\mathrm{SL}(3))$. The first
two skein relations,
\begin{align*}
\begin{tikzpicture}[baseline=-.5ex,web,allow upside down]
\draw[midto,rotate=180] (0,0) circle (.4);
\end{tikzpicture}\;\; &= \;\; 3 \\
\begin{tikzpicture}[baseline=-.5ex,web]
\draw[midto] (-1.25,0) -- (-.5,0);
\draw[midto] (.5,0) -- (1.25,0);
\draw (.5,0) arc (30:150:.577); \draw[midto] (0,.288) -- (-.01,.288);
\draw (.5,0) arc (330:210:.577); \draw[midto] (0,-.288) -- (-.01,-.288);
\end{tikzpicture}\;\; &= \;\; 2 \;\;
\begin{tikzpicture}[baseline=-.5ex,web]
\draw[midto] (-.5,0) -- (.5,0);
\end{tikzpicture}
\end{align*}
are straightforward, because the relevant fibres are always $\P^2$ and
$\P^1$, respectively. The third skein relation,
$$\begin{tikzpicture}[baseline=-.5ex,web]
\draw (.35,.35) arc (60:120:.7) arc (150:210:.7)
arc (240:300:.7) arc (-30:30:.7);
\draw[midfrom] (0,.444) -- (.01,.444);
\draw[midfrom] (-.444,0) -- (-.444,-.01);
\draw[midfrom] (0,-.444) -- (-.01,-.444);
\draw[midfrom] (.444,0) -- (.444,.01);
\draw[midto] (.35,.35) -- (.707,.707);
\draw[midfrom] (-.35,.35) -- (-.707,.707);
\draw[midto] (-.35,-.35) -- (-.707,-.707);
\draw[midfrom] (.35,-.35) -- (.707,-.707);
\end{tikzpicture}\;\; =\;\;
\begin{tikzpicture}[baseline=-.5ex,web]
\draw (.5,.5) arc (315:225:.707);
\draw[midto] (0,.293) -- (.01,.293);
\draw (-.5,-.5) arc (135:45:.707);
\draw[midto] (0,-.293) -- (-.01,-.293);
\end{tikzpicture}\;\; + \;\
\begin{tikzpicture}[baseline=-.5ex,web]
\draw (.5,-.5) arc (225:135:.707);
\draw[midto] (.293,0) -- (.293,.01);
\draw (-.5,.5) arc (45:-45:.707);
\draw[midto] (-.293,0) -- (-.293,-.01);
\end{tikzpicture},$$
is a little bit more work. The diskoid dual to the left side consists
of four triangles. The configuration space of the quadrilateral
$P(\omega_1,\omega_2,\omega_1,\omega_2)$ has two components, corresponding
to the two ways to collapse the quadrilateral to two edges. In each case,
there is a unique extension to the diskoid in which the diskoid collapses to
two triangles. The remaining case that should be checked is the intersection
of the two components in which the quadrilateral collapses to a single edge.
In this case the fibre is $\P^1$, because there is a $\P^1$ of ways to
extend the edge to a triangle, and the diskoid can collapse onto this
triangle. Thus the local Euler characteristic at the intersection is 2,
which matches the sum on the right side of the skein relation.
Thus, the image of $E$ is either equivalent to $\mathrm{\mathbf{rep}}^u_{-1}(\mathrm{SL}(3))_\min$,
or it is a quotient. However, $\mathrm{\mathbf{rep}}^u_{-1}(\mathrm{SL}(3))_\min$ is simple
as a linear-additive, pivotal category, because the pairing of dual
invariant spaces is non-degenerate. (Or, \thm{th:unitriang} also
implies that basis webs are linearly independent after applying $E$
because of unitriangularity.) Therefore the image of $E$ is equivalent
to $\mathrm{\mathbf{rep}}^u_{-1}(\mathrm{SL}(3))_\min$ itself.
In the case $G = \mathrm{SL}(2)$, we only need to check the skein relation
\eqref{e:tl} with $q=-1$:
$$\begin{tikzpicture}[baseline=-.5ex,web]
\draw (0,0) circle (.4);
\end{tikzpicture} \;\;= \;\;2.$$
In this case the diskoid of the left side is a based edge, the diskoid of
the right side is a point, the fibre is $\P^1$, and its Euler characteristic
is 2 as desired.
\end{proof}
It should also be possible to prove \conj{c:eulerrep} when $G = \mathrm{SL}(m)$.
The idea is to use the geometric skew Howe duality of Mirkovi\'c and
Vybornov \cite{MV:bdg} and the ideas in \cite[Sec. 6]{Kamnitzer:lectures}
to express this conjecture in terms of constructible functions on quiver
varieties for the Howe dual $\mathrm{SL}(n)$. Then we are in a position to apply
Nakajima's work from \cite[Sec. 10]{Nakajima:kacmoody}. Note that this
approach does not make use of the geometric Satake correspondence.
\subsection{Relationship with homological convolutions}
A constructible function is constant on a dense open subset of any
irreducible variety. If $X$ is an irreducible variety, we write
$f(X)$ for the value of $f$ on this dense open subset.
We can define a non-functor $\Xi$ from $\mathrm{\mathbf{econv}}(\mathrm{Gr})$ to $\mathrm{\mathbf{hconv}}(\mathrm{Gr})$
as follows. On objects, $\Xi$ is the identity, while on morphisms we define
$$\Xi:\C_c(Z({\vec{\lambda}}, {\vec{\mu}})) \longrightarrow H_\top(Z({\vec{\lambda}}, {\vec{\mu}}))$$
by the formula
$$\Xi:f \mapsto \sum_{X \in \mathrm{Irr}(Z({\vec{\lambda}}, {\vec{\mu}}))} f(X)[X].$$
The map $\Xi$ is not a functor because it does not respect convolution
(as some simple examples show). However, we offer the following tentative
conjecture.
\begin{conjecture} The map $\Xi$ between hom spaces restricts to an
equivalence of pivotal categories from $\mathrm{\mathbf{econv}}(\mathrm{Gr})_0$ to $\mathrm{\mathbf{hconv}}(\mathrm{Gr})$
up to a sign correction of the tensor and pivotal structures.
\label{c:eulerhom} \end{conjecture}
This conjecture implies \conj{c:eulerrep} one because this conjectured
equivalence is compatible with the functors from $\mathrm{\mathbf{fsp}}(G)$. This conjecture
would also imply the following simple formula for the expansion of the
invariant vectors coming from webs in the Satake basis which generalizes
\conj{c:eulerclosed}.
\begin{conjecture}[Corollary of \conj{c:eulerhom}]
Let $w$ be a minuscule web with boundary ${\vec{\lambda}}$ and dual diskoid
$D$. Then we can expand $\Psi(w)$ in the Satake basis as
$$\Psi(w) = \pm \sum_{X \in \mathrm{Irr}(F({\vec{\lambda}}))} \chi(\pi^{-1}(x))[X],$$
where $x$ is a generic point of each $X$, and $\pi:Q(D) \to F({\vec{\lambda}})$
is the restriction map from a diskoid configuration to its boundary.
\end{conjecture}
As partial evidence for \conj{c:eulerhom}, we note that a similar result
has been conjectured in the quiver variety setting.
\section{Future work}
This article is hopefully only the beginning of an investigation into
configuration spaces of diskoids and their relations to presented pivotal
categories, or spiders.
\subsection{Basis webs for $\mathrm{SL}(n)$}
In future work, the first author will establish the following
generalization of \thm{th:sl3basis} and \thm{th:unitriang} to $\mathrm{SL}(n)$.
\begin{theorem} Given a sequence of minuscule weights ${\vec{\lambda}}$ of $\mathrm{SL}(n)$,
there is a map $w({\vec{\mu}})$ from minuscule path ${\vec{\mu}}$ of type ${\vec{\lambda}}$
to webs. The image of these webs in $\mathrm{Inv}(V({\vec{\lambda}}))$ are a basis,
and the change of basis to the Satake basis is upper unitriangular with
respect to the partial order on minuscule paths.
\label{th:westbury} \end{theorem}
The geometric results of the current article are used to establish that
the webs $w({\vec{\mu}})$ form a basis and as far as we are aware no elementary
proof is available. This is in sharp contrast to the $\mathrm{SL}(3)$ case
where the basis webs were originally established by elementary means.
The webs $w({\vec{\mu}})$ themselves are constructed combinatorially using the
idea of Westbury triangles \cite{Westbury:so7} and \cite{Westbury:general}.
Recently, Westbury has combinatorially obtained \thm{th:westbury} for
the case of a tensor product of standard representation and their duals.
Kim \cite{Kim:thesis} (for $n=4$) and Morrison \cite{Morrison:diagram}
(for general $n$) conjecture a set of generating relations for kernel of
$\mathrm{\mathbf{fsp}}(\mathrm{SL}(n)) \to \mathrm{\mathbf{rep}}(\mathrm{SL}(n))_\min$. Using \thm{th:westbury}, we hope
to establish Kim's and Morrison's conjectures.
\subsection{Other rank 2 groups}
Since there are established definitions of spiders for $B_2$ and $G_2$,
it seems quite possible that the results in this paper could be generalized
to these two cases, but there are two important problems to resolve. First,
the vertex set of the corresponding affine buildings are no longer simply the
points of the affine Grassmannian. Second, since we want to study $\mathrm{\mathbf{rep}}(G)$
rather than $\mathrm{\mathbf{rep}}(G)_\min$, it is necessary to look at webs labelled not just
by minuscule weights but by fundamental weights. When $G$ is not $\mathrm{SL}(n)$,
it is no longer the case that all fundamental weights are minuscule;
thus the results of this paper would need to be extended to cover this case.
\subsection{Other discrete valuation rings}
Our results in this article apply only to the affine Grassmannians of the
discrete valuation ring $\mathcal{O} = \mathbb{C}[[t]]$. In fact, the affine Grassmannian
$\mathrm{Gr}$ exists (as a set) and Bruhat-Tits building $\Delta$ exists and is
$\mathrm{CAT}(0)$ for any complete discrete valuation ring $\mathcal{O}$. It is a well-known
open problem to state and prove a geometric Satake correspondence in this
setting, it is only known in the equal characteristic case $\mathcal{O} = k[[t]]$
for a field $k$. Since the building geometry is so similar for all choices
of $\mathcal{O}$, our results could be interpreted as (further) evidence that a
geometric Satake correspondence exists for all $\mathcal{O}$.
\subsection{Webs in surfaces}
Another possible generalization is from webs in disks to webs in surfaces.
If $\Sigma$ is a closed surface and $G$ has rank 1 or 2, there is an
analogous basis of non-elliptic webs on $\Sigma$ \cite{SK:confluence}, which
are equivalent to $\mathrm{CAT}(0)$ triangulations. (Or, $\Sigma$ can have boundary
circles with marked points, but the closed case is especially interesting.)
This web basis is a basis of the skein module of $\Sigma \times [0,1]$, which
is also the coordinate ring of the variety of representations $\pi_1(\Sigma)
\to G$. Our results suggest an interpretation of this coordinate ring in
terms of certain simplicial maps from the universal cover of $\Sigma$ to
the affine building $\Delta$. This should be related to the conjectures
of Fock-Goncharov \cite{FG:Teichmuller}.
\subsection{Categorification}
We would also like to apply our results to categorification and knot
homology. According to the philosophy of \cite{CK:coherent2}, to each web
$w$ with dual diskoid $D$ and boundary ${\vec{\lambda}}$, we should associate an
object $A(w)$ in the derived category of coherent sheaves on $\mathrm{Gr}({\vec{\lambda}})$.
When the configuration space $Q(D)$ has the expected dimension, then $A(w)$
should be the pushforward $\pi_*(\mathcal{O}_{Q(D)})$ of the structure sheaf
of $Q(D)$. It would also be nice to understand foams (as introduced by
Khovanov \cite{Khovanov:sl3}) in this language. In particular, it would
be interesting to consider the configuration spaces of duals of foams.
Some ideas in this direction have been pursued by Frohman.
\subsection{Quantum groups}
Finally, developing a $q$-analogue of our theory is also an open problem.
As mentioned earlier there is a functor from the free spider $\mathrm{\mathbf{fsp}}(G)$ to
$\mathrm{\mathbf{rep}}_q(G)$, the representation category of the quantum group for any $q$
not a root of unity. However, our geometric Satake machinery only applies
in the case when $q = 1$. Hopefully, we can extend to general $q$ using
the quantum geometric Satake developed by Gaitsgory \cite{Gaitsgory:twisted}.
\providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace}
\providecommand{\MR}{\relax\ifhmode\unskip\space\fi MR }
\providecommand{\MRhref}[2]{%
\href{http://www.ams.org/mathscinet-getitem?mr=#1}{#2}
}
\providecommand{\href}[2]{#2}
|
\section{Introduction}
\label{sec:intro}
\begin{defn}\label{defn:Lambdafreegp} Let $\Lambda$ be an ordered abelian group. By a $\Lambda$-\emph{free} group we mean a finitely generated group $G$ equipped with a free Lyndon $\Lambda$-length function. We call $G$
a \emph{regular} $\Lambda$-\emph{free} group if the length function is free and regular (see section~\ref{sec:prelim} for the definitions).
\end{defn}
Length functions were introduced by Lyndon~\cite{Lyndon} (see also~\cite{LS77}) to generalize Nielsen methods for free groups. They generalize the notion of a word metric on a group and remain a valuable tool in the combinatorial group theory. The Chiswell construction, introduced by Chiswell~\cite{ChiswellConstruction} for groups with $\mathbb{R}$-valued length functions and generalized by Morgan and Shalen~\cite{MS} to $\Lambda$-\emph{free} group for arbitrary ordered group $\Lambda$, relates length functions with group actions on $\Lambda$-trees. The construction allows one to use free group actions and free Lyndon length functions as two equivalent languages describing the same objects. We refer to the book \cite{ChiswellBook} by Chiswell for a detailed discussion on the subject.
Thus, $\Lambda$-\emph{free} groups can be thought of as groups acting freely on $\Lambda$-trees, and regular $\Lambda$-\emph{free} groups are precisely those acting with unique orbit of branch points~\cite{KharlampovichMyasnikov10}.
Bass-Serre theory of groups acting freely on simplicial trees~\cite{Serre} implies in particular that free groups are precisely $\mathbb{Z}$-free groups in our terminology. According to Rips' theorem, a finitely generated group is $\mathbb{R}$-free if and only if it is a free product of free abelian groups and surface groups (with few exceptions)~\cite{GLP}. Free actions on $\mathbb{R}$-trees cover all Archimedean actions, since every group acting freely on a $\Lambda$-tree for an Archimedean ordered abelian group $\Lambda$ acts freely also on an $\mathbb{R}$-tree.
Group actions on $\Lambda$-trees for an arbitrary ordered abelian group $\Lambda$ were introduced by Morgan and Shalen~\cite{MS} and studied by Alperin and Bass~\cite{AlperinBass}. Bass~\cite{Bass} proved a version of combination theorem for finitely generated groups acting freely on ($\Lambda \oplus \mathbb{Z}$)-trees with respect to the lexicographic order on $\Lambda \oplus \mathbb{Z}$; this was generalized by Martino and Rourke~\cite{MartinoRourke}. A structure theorem for finitely generated groups acting freely on an ${\mathbb R}^n$-tree (with the lexicographic order) was proved by Guirardel~\cite{Guirardel}. The theorem states that a finitely generated $\mathbb{R}^n$-free group is the fundamental group of a graph of groups with cyclic edge subgroups and vertex groups that are finitely generated $\mathbb{R}^{n-1}$-free groups. The structure theorem, along with the combination theorem for relatively hyperbolic groups proved by Dahmani~\cite{Dahmani}, implies that finitely generated $\mathbb{R}^n$-free groups are hyperbolic relative to maximal abelian subgroups. Note that every $\mathbb{Z}^n$-free group is in particular $\mathbb{R}^n$-free, so that finitely generated $\mathbb{Z}^n$-free groups are relatively hyperbolic; we use this result below.
Kharlampovich, Myasnikov, Remeslennikov and Serbin proved necessary and sufficient conditions for a finitely generated group to be regular $\mathbb{Z}^n$-free~\cite{KharlampovichMyasnikov10}. In the present paper we use the 'only if' part - that is, a structure theorem~\cite[Theorem 7]{KharlampovichMyasnikov10} (see also theorems~\ref{thm:main2} and~\ref{thm:singleHNN} in the present paper). The structure theorem describes a finitely generated regular $\mathbb{Z}^n$-free group as a group obtained from a finitely generated free group by a sequence of finitely many HNN-extensions of a particular type. The same authors showed in~\cite{KharlampovichMyasnikov11} that every finitely generated $\mathbb{Z}^n$-free group embeds by a length-preserving monomorphism into a finitely generated regular $\mathbb{Z}^m$-free group, for some $m$. (Notice that by \cite{ChiswellMuller}, every $\Lambda$-free group embeds by a length-preserving monomorphism into a regular $\Lambda$-free group.) We use these theorems to prove our main result (cf. Theorems~\ref{thm:regularZnfreecoherent} and~\ref{thm:ZnfreeCAT0}):
\begin{thm}\label{thm:ZnfreeCAT0intro} A finitely generated $\mathbb{Z}^n$-free group $G$ acts properly discontinuously and cocompactly on a CAT(0) space. If the length function $l\colon G\rightarrow\mathbb{Z}^n$ is also regular then $G$ is the fundamental group of a geometrically coherent space; in particular, $G$ is coherent.
\end{thm}
We refer the reader to the book~\cite{BridsonHaefliger} for definitions and basic properties of CAT(0) spaces and groups acting on them.
Recall that a group $G$ is called \emph{coherent} if every finitely generated subgroup of $G$ is finitely presented. Definition~\ref{defn:coherentspace} of a geometrically coherent space can be found in section~\ref{sec:gluing}. The coherence of regular $\mathbb{Z}^n$-free groups can also be deduced from the structure theorem~\cite[Theorem 7]{KharlampovichMyasnikov10}, using the Bass-Serre theory.
Since $\mathbb{Z}^n$-free groups are relatively hyperbolic, the criterion~\cite[Theorem 1.2.1]{HruskaKleiner} proved by Hruska and Kleiner applies, and we conclude that the CAT(0) space in theorem~\ref{thm:ZnfreeCAT0intro} has isolated flats in the sense of Hruska~\cite{HruskaThesis}:
\begin{thm}\label{thm:CAT0isolflats} A finitely generated $\mathbb{Z}^n$-free group $G$ acts properly discontinuously and cocompactly on a CAT(0) space with isolated flats.
\end{thm}
It is unknown, whether all hyperbolic groups are CAT(0). Consequently, the question is open for groups hyperbolic relative to abelian parabolic subgroups. Even the case of $\mathbb{R}^n$-free groups with $n\geq 2$ is open, still.
Kharlampovich and Myasnikov proved in~\cite{KharlampovichMyasnikov08} that every finitely generated fully residually free (or limit) group acts freely on a $\mathbb{Z}^n$-tree for some $n$. These groups form a proper subclass of the class of $\mathbb{Z}^n$-free groups, as the following example shows.
\begin{exmp} $G=\langle x_1,x_2,x_3\mid x_1^2x_2^2x_3^2 = 1\rangle$ is an example of a $\mathbb{Z}^2$-free group which is not fully residually free. Indeed, since $G$ splits as the amalgamated product of 2-generated free groups:
\[
G=\langle x_1,x_1^2x_2\rangle\ast_{\langle x_1^2x_2=(x_2x_3^2)^{-1}\rangle}\langle x_2x_3^2,x_3\rangle,
\]
by Bass' Theorem~\cite{Bass} it is $\mathbb{Z}^2$-free. To show that $G$ is not (fully) residually free, note that in a free group, $x_1^2x_2^2$ is not a proper square unless $[x_1,x_2]=1$.
\end{exmp}
For limit groups theorem~\ref{thm:CAT0isolflats} was proven by Alibegovic and Bestvina~\cite{AlibegovicBestvina}. We adapt their approach to the more general situation.
Notice that ${\mathbb Z}^n$-free groups possess better algorithmic properties than relatively hyperbolic groups with abelian parabolics do in general. For example, by Nikolaev's result \cite{Nikol}, ${\mathbb Z}^n$-free groups have decidable membership problem for finitely generated subgroups while there are examples of hyperbolic groups where this problem is undecidable.
\medskip
The paper is organized as follows. Section~\ref{sec:prelim} contains definitions as well as basic properties of free actions on $\mathbb{Z}^n$-trees. The results presented in this section can be found in~\cite{KharlampovichMyasnikov10}, where the authors use the language of infinite words. In Section~\ref{sec:structure} we construct a weighted word metric on a given regular $\mathbb{Z}^n$-free group $G$, to be used in the proof of theorem~\ref{thm:ZnfreeCAT0intro} in Section~\ref{sec:gluing}.
\section{Length functions and actions}
\label{sec:prelim}
\subsection{Lyndon length functions and free actions on trees}
\label{subsec:Lyndon}
Let $G$ be a group and $\Lambda$ an ordered
abelian group. Then a function $l: G \rightarrow \Lambda$ is called a {\it
(Lyndon) length function} on $G$ if the following conditions hold:
\begin{enumerate}
\item [(L1)] $\forall\ g \in G:\ l(g) \geqslant 0$ and $l(1)=0$;
\item [(L2)] $\forall\ g \in G:\ l(g) = l(g^{-1})$;
\item [(L3)] $\forall\ g, f, h \in G:\ c(g,f) > c(g,h)
\rightarrow c(g,h) = c(f,h)$,
\noindent where $c(g,f) = \frac{1}{2}(l(g)+l(f)-l(g^{-1}f))$.
\end{enumerate}
Sometimes we refer to length functions with values in $\Lambda$ as to $\Lambda$-length functions.
While $c(g,f)$ may not belong to $\Lambda$ in general, we are interested in the case when it does.
\begin{defn} \textbf{(Free Length Function)}
A length function $l:G \rightarrow A$ is called {\em free} if it satisfies the following two axioms.
\begin{enumerate}
\item [(L4)] $\forall\ g, f \in G:\ c(g,f) \in \Lambda.$
\item [(L5)] $\forall\ g \in G:\ g \neq 1 \rightarrow l(g^2) > l(g).$
\end{enumerate}
\end{defn}
Moreover, the following particular case is of crucial importance.
\begin{defn} \textbf{(Regular Length Function)} A length function $l: G \rightarrow \Lambda$ is called {\it regular} if it satisfies the {\it regularity} axiom:
\begin{enumerate}
\item [(L6)] $\forall\ g, f \in G,\ \exists\ u, g_1, f_1 \in G:$
$$g = u g_1,\ l(g)=l(u)+l(g_1);\ f = u f_1,\ l(f)=l(u)+l(f_1); \ l(u) = c(g,f).$$
\end{enumerate}
\end{defn}
Many examples of groups with regular free length functions are given in \cite{KharlampovichMyasnikov10}.
\subsection{$\Lambda$-trees}
Group actions on $\Lambda$-trees provide a geometric counterpart to $\Lambda$-length functions. To explain we need the following definitions, which we also use later in the text.
Let $X$ be a non-empty set, $\Lambda$ an ordered abelian group. An $\Lambda$-{\em metric on} $X$ is a mapping
$d\colon X \times X\longrightarrow A$ that satisfies the usual axioms of a metric.
The pair $(X,d)$ is an {\em $\Lambda$-metric space}. If $(X,d)$ and
$(X',d')$ are $\Lambda$-metric spaces, an {\it isometry} from
$(X,d)$ to $(X',d')$ is a mapping $f: X \rightarrow X'$ such that
$d(x,y) = d'(f(x),f(y))$ for all $x,y \in X$.
For elements $a,b \in \Lambda$ the {\it closed segment} $[a,b]$ is
defined by
$$[a,b] = \{c \in \Lambda \mid a \leqslant c \leqslant b \}.$$
More generally, a {\it segment} in an $\Lambda$-metric space $X$ is the image of an
isometry $$\alpha: [a,b] \rightarrow X$$ for some $a,b \in
\Lambda$. We denote the segment by $[\alpha(a), \alpha(b)]$; $\alpha(a)$ and $\alpha(b)$ are its endpoints. The length of the segment is $p(\alpha(a),\alpha(b))=b-a$. A $\Lambda$-metric space $(X,d)$ is
{\it geodesic} if for all $x,y \in X$, there is a segment of length $d(x,y)$ in $X$ with the endpoints $x,y$.
A {\em $\Lambda$-tree} is a $\Lambda$-metric space $(X,d)$ such that:
\begin{enumerate}
\item[(T1)] The space $(X,d)$ is geodesic,
\item[(T2)] If two segments in $X$ intersect in a single point, which
is an endpoint of both, then their union is a segment,
\item[(T3)] The intersection of two segments with a common endpoint is also a segment.
\end{enumerate}
\subsection{Actions on trees}
Let a groups $G$ act on a $\Lambda$-tree $T$ by isometries. Recall that an element $g\in G$ is called \emph{elliptic} if $g$ fixes a point in $T$, $g$ is an \emph{inversion} if $g$ does not fix a point in $T$ but $g^2$ does, and $g$ is \emph{hyperbolic} if $g$ is neither an elliptic element, nor an inversion.
An action of $G$ on $X$ is termed {\em free} if for every $1\neq g \in G$ neither $g$ nor $g^2$ has a fixed point in $X$; in other words, an action of $G$ on $T$ is free if and only if every nontrivial element of $G$ is hyperbolic.
Let a group $G$ act on a $\Lambda$-tree $T$ with a base point $x$, and let the length function $l_x\colon G\rightarrow \Lambda$ based at $x$ be defined as follows:
$l_x(g) = d(x,gx),\forall\ g \in G$. If $l_x$ is a free regular Lyndon length function on $G$ then the action is free, and $\forall g,f\in G$ $\exists w\in G$ so that $[x,gx]\cap[x,fx]=[x,wx]$. The element $w$ is called the \emph{common beginning} of $f$ and $g$, denoted by $com(f,g)=w$.
Along with the notion of a length function $l\colon G\rightarrow \Lambda$ we use the \emph{translation length function }$l_T\colon G\rightarrow \Lambda$, defined as follows:
\[
l_T(g)=\min_{t\in T}d(t,gt).
\]
Clearly, $g$ is elliptic iff $l_T(g)=0$. Every hyperbolic element $g\in G$ has an \emph{axis} which we denote by $A_g$.
The axis of $g$ is the set of all the points in $T$ that $g$ moves by the shortest possible distance:
\[
A_g=\{t\in T\mid d(t,gt)=l_T(g)\}.
\]
Note that while the values $l(g)$ of the length function depend on the choice of a base point $x\in T$, the values of the translation length function $l_T(g)$ do not. For every hyperbolic element $g\in G$, $l(g)\geq l_T(g)$, and $l(g)=l_T(g)$ iff $x\in A_g$.
Unlike $l(g)$, the translation length is a conjugacy class invariant, so that
\[
l_T(a^{-1}ga)=l_T(g), \ \forall a\in G.
\]
\begin{lem} \label{lem:axesthroughx} Let $G$ act freely on a $\Lambda$-tree $T$, and let us fix a base point $x\in T$. If the associated length function $l\colon G\rightarrow \Lambda$ is regular then every $g\neq 1$ in $G$ is conjugate to an element $f\in G$ so that $x\in A_f$.
\end{lem}
\begin{proof} Assume that $x\notin A_g$. Let $y\in A_g$ be the point on $A_g$ closest to $x$. Note that $[x,y]$=$[x,gx]\cap[x,g^{-1}x]$, so that $d(x,y)=c(g,g^{-1})$. By the regularity of $l$, there is $u\in G$ so that $y=ux$. Then $x=u^{-1}y$, and hence $x\in u^{-1}A_g = A_{ugu^{-1}}$. We set $f=ugu^{-1}$ to obtain the claim.
\end{proof}
In what follows, we will only deal with those representatives of conjugacy classes that are shortest with respect to $l$. The axes of these elements necessarily contain the base point $x$. If $F$ is a finitely generated free group acting freely on its Cayley graph $T$, the associated length function is clearly regular. Without loss of generality, in this case we can assume that the base point of $T$ corresponds to the identity of $F$.
\subsection{Basic properties of $\mathbb{Z}^n$-free groups}
If the length $l$ is Archimedean, that is $\Lambda=\mathbb{Z}$ or $\Lambda\subseteq\mathbb{R}$, then two hyperbolic elements $g$ and $f$ commute in $G$ iff $A_f=A_g$. If $\Lambda=\mathbb{Z}^n$, so that the length is not Archimedean, following~\cite{KharlampovichMyasnikov10}, we introduce the notion of height, as follows.
\begin{defn}\label{defn:height} \textbf{(Height of element)} Let $l\colon G\rightarrow\mathbb{Z}^n$ be a length function. By the \emph{height} of an element $g\in G$, denoted $ht(g)$, we mean the number of the most right non-zero component of $l(g)$. If $l(g)=0$ then we say that $ht(g)=0$. By the height of a set $S\subset G$ we mean the maximum height of its elements: $$ht(S)=\max_{g\in S}\{ht(g)\}. $$
\end{defn}
For a general ordered abelian group $\Lambda$, we write $l(g)\gg l(f)$ (or $l_T(g)\gg l_T(f)$) whenever $l(g)$ is infinitely larger than $l(f)$ (or $l_T(g)$ is infinitely larger than $l_T(f)$). For $\Lambda=\mathbb{Z}^n$, $l(g)\gg l(f)$ is equivalent to $ht(g)>ht(f)$. We call $l(g)$ and $l(f)$ (or $l_T(g)$ and $l_T(f)$) \emph{comparable} if none of them is infinitely larger than the other. For $\Lambda=\mathbb{Z}^n$, $l(g)$ and $l(f)$ are comparable iff $ht(g)=ht(f)$.
Let $F$ be a free group - that is, $F$ is $\mathbb{Z}$-free. If $u\in F$ is cyclically reduced and $sus^{-1}=v$ for some $v,s\in F$, then $l(v)>l(u)$, unless $s=\hat{s}u^k$, where $\hat{s}^{-1}$ is an initial subword of $u$ and $k$ is an integer, hence $v$ is a cyclic permutation of $u$ and $l(v)=l(u)$. In Lemma~\ref{lem:axisofstableletter} below we study generalizations of this.
\begin{lem}\label{lem:axisofstableletter} Let $G$ be a free $\mathbb{Z}^n$-group, and let $u$ and $v$ be cyclically reduced elements of $G$. Suppose that $sus^{-1}=v$ for some nontrivial element $s\in G$. Let $\hat{s}\in s \langle u\rangle$ be a short representative of the coset $s \langle u\rangle$, so that $l(\hat{s})\leq l(u)$. Then the following hold:
\begin{enumerate}
\item \label{e:shorts} If $ht(\hat{s})\leq ht(u)$ then $[x,\hat{s}^{-1}x]\subset[x,ux]$. If $s$ and $u$ commute then also $[x,\hat{s}x]\subset[x,u^{-1}x]$; otherwise, $[x,\hat{s}x]\subset[x,vx]$.
\item \label{e:longs} If $ht(\hat{s}) > ht(u)$ then $[x,u^px]\subset [x,sx]$ either for all non-negative integers $p$, or for all non-positive integers $p$.
If $s$ and $u$ do not commute, we also have that $[x,v^px]\subset [x,s^{-1}x]$, with $p$ as above.
\end{enumerate}
\end{lem}
\begin{proof} The assumption that both $u$ and $v$ are cyclically reduced and conjugate in $G$ implies that the axes $A_u$ of $u$ and $A_v$ of $v$ contain the base point $x$, and
$$l(u)=l_T(u)=l_T(v)=l(v).$$ In particular, $A_v=A_{sus^{-1}}=sA_u$ contains both $x$ and $sx$.
It is instructive to look first at the case when $ht(\hat{s}) = ht(u)$, so that $l(s)$ and $l(u)$ are comparable. In this case, for some integer $k$,
\begin{equation}\label{e:shortsinclusion}
x\in[su^kx,su^{m}x]=[su^ks^{-1}(sx),su^{m}s^{-1}(sx)],
\end{equation}
where $m=k+1$ if $k\geq 0$, and $m=k-1$ if $k<0$.
Equivalently, $s^{-1}x\in[u^kx,u^{m}x]$. We can choose $\hat{s}$ so as to write this latter inclusion as either $s^{-1}=u^k\hat{s}^{-1}$ if $k\geq 0$, or $s^{-1}=u^m\hat{s}^{-1}$ if $k< 0$. In any case, $[x,\hat{s}^{-1}x]\subset[x,ux]$. Note that in a free group this means that $\hat{s}^{-1}$ is an initial subword of $u$, and hence $v$ is a cyclic permutation of $u$.
If $s$ and $u$ commute then $v=u$ and $sA_u=A_u=s^{-1}A_u$, in particular $s^{-1}x\in A_u$. Since $d(x,sx)=d(x,s^{-1}x)$, $sx\neq s^{-1}x$, and so the tripod with the vertices $x,sx,s^{-1}x$ is actually the segment $[s^{-1}x,sx]\subset A_u$, we conclude that $[x,sx]\cap[x,s^{-1}x]=\{x\}$. It follows that similar to~(\ref{e:shortsinclusion}), there is the following inclusion:
\begin{equation}\label{e:commutativeinclusion}
x\in[s^{-1}u^{-k}x,s^{-1}u^{-m}x]=[u^{-k}(s^{-1}x),u^{-m}(s^{-1}x)]\ \Leftrightarrow\ sx\in [u^{-k}x,u^{-m}x],
\end{equation}
and for the same choice of $\hat{s}$ as above, $[x,\hat{s}x]\subset[x,u^{-1}x]$.
Let now $ht(\hat{s}) < ht(u)$, which amounts to $l(\hat{s})\ll l(u)=l(v)=l(sus^{-1})$. It follows that in~(\ref{e:shortsinclusion}) $k=0$ and $\hat{s}=s$, hence $[x,s^{-1}x]=[x,\hat{s}^{-1}x]\subset[x,ux]$.
The conclusion for $v$ follows from the observation that $s^{-1}vs=u$.
To prove the assertion~(\ref{e:longs}), note that for any integer $p$, $d(x,u^px)\ll d(x,sx)$. If $ux\in[x,s^{-1}x]$ then $u^{-1}x\in[x,sx]$, and so $[x,u^px]\subset[x,sx]$ for all $p\leq 0$, while $ux\notin[x,s^{-1}x]$ implies that $[x,ux]\in[x,sx]$, and so $[x,u^px]\subset[x,sx]$ for all $p\geq 0$.
Since $s^{-1}vs=u$, similar arguments show that $[x,vx]\subset[x,s^{\epsilon}x]$ where $\epsilon$ is either 1, or $-1$. W.l.o.g., assume that $[x,u^px]\subset[x,sx]$ for all $p\geq 0$. If also $[x,vx]\subset[x,sx]$ then necessarily $[x,vx] = [x,ux]$, as $l(u)=l(v)$. It follows that $v^{-1}ux=x$, and since the action is free, this implies $u=v$; that is, $s$ and $u$ commute. If $s$ and $u$ do not commute, and so $u\neq v$, we must have $[x,vx]\subset[x,s^{-1}x]$.
\end{proof}
\subsection{Abelian Subgroups.}\label{sec:abeliansbgps}
\begin{lem}\label{lem:reducedcentralizer} Let $G$ be a $\mathbb{Z}^n$-free group, $u,s\in G$ be nontrivial. If $u$ and $s$ commute then $x\in A_u\Leftrightarrow x\in A_s$.
\end{lem}
\begin{proof} Let $ht(s)\leq ht(u)$. Assume that $x\in A_u$; by the proof of Lemma~\ref{lem:axisofstableletter}(\ref{e:shorts}), $x,sx,s^{-1}x\in A_u$, and $[x,sx]\cap[x,s^{-1}x]=\{x\}$. It follows that $x\in A_s$.
Now, let $x\notin A_u$. By the way of contradiction, assume that $x\in A_s$. Let $y\in A_u$ be arbitrary. $sA_u=A_u$, hence $s^ky\in A_u$ for all $k$. If $y\in A_s\cap A_u$ then $x\in[s^ky,s^my]\subset A_u$ for some $k$ and $m$, and we get $x\in A_u$, a contradiction. In particular, $A_s\cap A_u=\emptyset$. Therefore, assume that $y\notin A_s$, and let $z\in A_s$ be the closest point to $y$. Then for some $k,m$, $$x\in[s^kz,s^mz]\subset[s^ky,s^my]\subset A_u,$$ a contradiction. Therefore, $x\notin A_u$ implies $x\notin A_s$.
\end{proof}
\begin{cor}\label{cor:reducedcentralizer} let $G$ be a regular $\mathbb{Z}^n$-free group, and let $\tilde{A}$ be an abelian subgroup of $G$. Then $\tilde{A}$ is conjugate to an abelian subgroup $A$ of $G$ such that every element of $A$ is cyclically reduced.
\end{cor}
\begin{proof} This follows from Lemmas~\ref{lem:axesthroughx} and ~\ref{lem:reducedcentralizer}.
\end{proof}
In the proof of Theorem~\ref{thm:rankabelian} below we use the following definition.
\begin{defn}\label{defn:Rbounded} Let $R=\{r_1,r_2,\dots,r_n\}$ be a set of elements of $G$ so that $ht(r_j) = j$, for all $j=1,2,\dots,n$. Let $C=(L_1,L_2,\dots,L_n)$ be the vector of the greatest non-zero components of the elements of $R$. More precisely, if $l(r_j)=(c^{(j)}_1,\dots,c^{(j)}_{j},0,\dots,0)$ then $L_j=c^{(j)}_j$.
A set
$P=\{p_1,p_2,\dots,p_k,\dots\}\subset G$ is called $R$-\emph{bounded} if for every $p\in P$ with the length $l(p)=(a_1,a_2,\dots,a_n)$ we have that $\forall j\leq n,\ a_j\leq L_j$.
\end{defn}
\begin{rem} With this definition, the set $R$ may not be $R$-bounded; for instance, the set $R=\{(1,0),(2,1)\}$ is not $R$-bounded. The set $\{(0,0),(1,0),(0,1),(1,1)\}$ is the maximal $R$-bounded subset of $\mathbb{Z}^2$.
\end{rem}
\begin{lem}\label{lem:Rboundedabelian} Let $G$ be a group with a regular free $\mathbb{Z}^n$-length function, and let $R=\{r_1,r_2,\dots,r_n\}$ be a set of elements of $G$ so that $ht(r_j)=j$, for all $j=1,2,\dots,n$. If $A$ is an abelian subgroup of $G$ then the number of $R$-bounded subsets of $A$ of a fixed finite cardinality is finite.
\end{lem}
\begin{proof} Since the Lemma holds for $\mathbb{Z}^n$, the assertion in general follows from the observation that for arbitrary $s,t\in A$, $l(s)=l(t)$ implies $s=t^{\pm 1}$.
\end{proof}
\begin{thm}\label{thm:rankabelian} Let $G$ be a free $\mathbb{Z}^n$-group, and let $A$ be an abelian subgroup of $G$. Then $A$ has a generating set $\{t_1,t_2,\dots,t_k\}$ so that $ht(t_i)<ht(t_{i+1})$ for all $i=1,\dots, k-1$. In particular, every abelian subgroup of $G$ can be generated by at most $n$ elements.
\end{thm}
\begin{proof} By Corollary~\ref{cor:reducedcentralizer}, we can assume that every element of $A$ is cyclically reduced.
Let us be given a generating set $S$ for $A$. First, assume that $S$ has elements of height 1. Let $S_1\subset S$ be the subset of all these elements. All the elements of height 1 in $G$ form a $\mathbb{Z}$-free subgroup $H$; hence, $\langle S_1\rangle\subset H$ is an abelian subgroup of a free group. Therefore, $\langle S_1\rangle=\langle s_0\rangle$ is cyclic, and we can assume that $S$ has only one element of height 1.
Next, we claim that if $ht(s)=ht(t)$ then $\langle s,t\rangle=\langle p,\tilde{s},\tilde{t}\ \rangle$ for some $p$,$\tilde{s}$ and $\tilde{t}$ with $ht(\tilde{s}),ht(\tilde{t})<ht(s)=ht(p)$. Indeed, let $m_s>0$ and $m_t>0$ be the largest non-zero components of $l(s)$ and $l(t)$, respectively. We set
\begin{equation}\label{e:gcd}
m=\sigma m_s+\tau m_t=gcd(m_s,m_t),\quad p=s^{\sigma }t^{\tau}.
\end{equation}
Then we obtain the claim for $$\tilde{s}=sp^{-\frac{m_s}{m}},\quad \tilde{t}=tp^{-\frac{m_t}{m}}.$$
Note that the largest nonzero component of $p$ equals $m$, so $l(p)\leq\min\{l(s),l(t)\}$. If $m=m_s$ then $p=s$, and we set $\tilde{s}=1$, so that $\langle s,t\rangle=\langle p, \tilde{t}\ \rangle$.
Now, we show that if $N\leq n$ is the maximum height of the elements of $S$:
\[
N=\max_{s\in S}\{ht(s)\},
\]
and $S$ has $N$ elements, all of different heights, then $A$ is generated by exactly $N$ elements, all of different heights.
Indeed, let $P=\{s_1,s_2,\dots,s_N\}$ be the $N$ given elements of $S$, so that $ht(s_i)=i$, for all $i=1,2,\dots,N$. Using (a version of) the Gaussian elimination process, we can replace $s_1,s_2,\dots,s_N$ by a $P$-bounded set $R=\{r_1,r_2,\dots,r_N\}$ so that $\langle s_1,s_2,\dots,s_N\rangle=\langle r_1,r_2,\dots,r_N\rangle$. Given another element $s\in S$, by iterating the procedure described above we can replace $R$ by another $P$-bounded set $R'$ of elements with different heights. If $\langle R\rangle=\langle R,s\rangle = \langle R'\rangle$ then $s\in\langle R\rangle$, and so it could be deleted from $S$. Otherwise, $\langle R\rangle\subsetneqq\langle R'\rangle$ which implies that at least one of the components of the vector $C_{R'}$ associated with the set $R'$ (see Definition~\ref{defn:Rbounded}) is strictly smaller than the corresponding component of $C_R$. In particular, $R'$ is a $P$-bounded set. Note that $R'$ has the same cardinality as $R$ does, because two elements of $A$ of height one always belong to a cyclic subgroup, as we noted above. However, elements of higher heights may ``disappear" in a similar way. Since the cardinality of our $P$-bounded generating sets is fixed, by Lemma~\ref{lem:Rboundedabelian}, the ascending chain of subgroups of $A$ generated by these sets is finite. In other words, we only need to add finitely many elements from $S$ so as to obtain a $P$-bounded generating set $\{t_1,\dots,t_N\}$ for $A$. The condition $ht(t_i)<ht(t_{i+1})$ for all $i$ holds by construction.
The same proof works in the general case. We only note that the generating sets we construct do not have to have an element for every height between $1$ and $n$. For instance, if $P$ has no element of height $j$ then we set the $j$th component in $C_R$ to be $L_j=0$. If an element of height $j$ is added at a later step then we call the new set $P$, make it $P$-reduced and start the process anew. This can only happen finitely many times, and the cardinality of any generating set never exceeds $n$.
\end{proof}
\begin{lem} Let $G$ be a group equipped with a free regular length function. Let $u$ and $v$ be cyclically reduced elements of $G$ with $l(u)=l(v)$, and suppose that $sus^{-1}=v$ for some nontrivial element $s\in G$. Then for some cyclic permutation $\hat{u}$ of $u$ and some cyclic permutation $\hat{v}$ of $v$ there is a cyclically reduced element $z\in G$ so that $z\hat{u}z^{-1}=\hat{v}$.
\end{lem}
\begin{proof} If $x\notin[s^{-1}x,sx]$ then $[x,s^{-1}x]\cap[x,sx]=[x,wx]$ for some $w\in G$, as the length function is regular. We have that $s=wzw^{-1}$ for some $z\in G$ (cf. Lemma~\ref{lem:axesthroughx}).
If $ht(s) > ht(u)$ and $l(w) \geq l(u)$ then $s$ and $u$ must commute. For if not then by Lemma~\ref{lem:axisofstableletter}, $[x,ux]\subset[x,wx]$ and $[x,vx]\subset [x,wx]$. Since $l(u)=l(v)$, it follows that $u=v$, hence $s$ and $u$ commute, a contradiction. The claim now follows from Lemma~\ref{lem:reducedcentralizer}.
So, we can assume that $l(w)\leq l(u)$. Then either case of Lemma~\ref{lem:axisofstableletter} implies $u=wu_1$ and $v=wv_1$. Therefore,
\[
wzw^{-1}(wu_1)wz^{-1}w^{-1}=wv_1\ \Rightarrow\ zu_1wz^{-1}=v_1w,
\]
where $z$ is cyclically reduced.
\end{proof}
\section{Structure of regular $\mathbb{Z}^n$-free groups} \label{sec:structure}
\begin{defn}\label{defn:coherentdirections} Let $G$ act on a $\Lambda$-tree $T$ with a base point $x$. Let $f$ and $g$ be hyperbolic elements with overlapping axes so that $A_g\cap A_f$ contains more than one point. We say that $f$ and $g$ have \emph{coherent directions}, or simply that $f$ and $g$ are \emph{coherent} if they there is a point $p\in A_f\cap A_g$ so that $p\notin [gp,gf]$.
\end{defn}
\begin{lem} \label{lem:sumoflettersthroughbase}
Let $G$ be a finitely generated group acting freely on a $\mathbb{Z}^n$-tree with a base point $x$. Let $g_1,\dots,g_m$ be nontrivial cyclically reduced (not necessarily distinct) elements of $G$, and let $u=g_1\dots g_{m}$. Assume that whenever for two consecutive factors $g=g_{i}$ and $f=g_{i+1}$ in the word $u$ the intersection $A_g\cap A_f$ contains more than one point, $g$ and $f$ are coherent.
Then the axis of $u$ contains $x$, and
$$l(u)=\sum_{i=1}^m l(g_i).$$
\end{lem}
\begin{proof} We prove the assertion by induction on the length $m$ of $u$. Note that every $g\neq 1$ in $G$ is hyperbolic. For $m=1$ the claim is trivial, and for $m=2$ it is proved in~\cite{ChiswellBook}. The proof shows that under our assumptions, $l_T(fg)=l_T(f)+l_T(g)$, and the axis of $fg$ contains the intersection $A_f\cap A_g\subseteq A_{fg}$. Moreover, both intersections $A_f\cap A_{fg}$ and $A_g\cap A_{fg}$ contain more than one point, and $f$ and $fg$, as well as $g$ and $fg$ are coherent. In particular, if we denote by $$u_1=f_1g_3\dots,g_{m},$$ where $f_1=g_1g_2$, both assumptions will hold for $u_1$ as well. The claim now follows by induction.
\end{proof}
\begin{defn} An abelian subgroup of a regular $\mathbb{Z}^n$-free group is termed in~\cite{KharlampovichMyasnikov10} a \emph{cyclically reduced centralizer} if every element of $A$ is cyclically reduced (cf. Corollary~\ref{cor:reducedcentralizer}).
\end{defn}
\begin{thm} \cite[Theorem 7]{KharlampovichMyasnikov10}
\label{thm:main2}
Let $G$ be a finitely generated group with a regular free Lyndon
length function in $\mathbb{Z}^n$. Then $G$ can be represented as
a union of a finite series of groups
$$G_1 < G_2 < \cdots < G_r = G,$$
where $G_1$ is a free group of finite rank, and
$$G_{i+1} = \langle G_i, s_{i,1},\ \ldots,\ s_{i,k_i} \mid s_{i,j}^{-1}\ C_{i,j}\ s_{i,j} = \phi_{i,j}(C_{i,j}) \rangle,$$
where for each $j \in [1,k_i],\ C_{i,j}$ and $\phi_{i,j}(C_{i,j})$ are cyclically reduced centralizers of $G_i$,
$\phi_{i,j}$ is an isomorphism, and the following conditions are satisfied:
\begin{enumerate}
\item[(1)] $C_{i,j} = \langle c^{(i,j)}_1, \ldots, c^{(i,j)}_{m_{i,j}} \rangle,\
\phi_{i,j}(C_{i,j}) = \langle d^{(i,j)}_1, \ldots, d^{(i,j)}_{m_{i,j}} \rangle$,
where $\phi_{i,j}(c^{(i,j)}_k) = d^{(i,j)}_k,\ k \in [1,m_{i,j}]$ and
$$ht(c^{(i,j)}_k) = ht(d^{(i,j)}_k) < ht(d^{(i,j)}_{k+1}) = ht(c^{(i,j)}_{k+1}),\
k \in [1,m_{i,j}-1],$$
$ht(s_{i,j})>ht(c^{(i,j)}_k)$,
\item[(2)] $\ell(\phi_{i,j}(w)) =\ell (w)$ for any $w \in C_{i,j}$,
\item[(3)] $w$ is not conjugate to $\phi_{i,j}(w)^{-1}$ in $G_i$ for any $w \in C_{i,j}$,
\item[(4)] if $A, B \in \{C_{i,1}, \phi_{i,1}(C_{i,1}), \ldots, C_{i,k_i},
\phi_{i,k_i}(C_{i,k_i})\}$ then either $A = B$, or $A$ and $B$ are not conjugate
in $G_i$,
\item[(5)] $C_{i,j}$ can appear in the list
$$\{ C_{i,k}, \phi_{i,k}(C_{i,k}) \mid k \neq j \}$$
not more than twice.
\item[(6)] $ht(s_{i,j})=ht(s_{i,k})$, for all $1\leq k,j\leq k_i$.
\end{enumerate}
\end{thm}
\begin{cor}\label{cor:coherenceinabeliansbgp} In every associated abelian subgroup $C$, $\phi(C)$ of the HNN extensions from Theorem~\ref{thm:main2} generators can be chosen so that their directions are all coherent in the meaning of Definition~\ref{defn:coherentdirections}, and the choice of the directions of the generators in $C$ and $\phi(C)$ is compatible with $\phi$.
\end{cor}
\begin{proof} In $C_{i,j} = \langle c^{(i,j)}_1, \ldots, c^{(i,j)}_{m_{i,j}} \rangle$ choose all $c^{(i,j)}_k$ be coherent with $c^{(i,j)}_1$ by replacing $c^{(i,j)}_k$ with its inverse if necessary. In $\phi_{i,j}(C_{i,j})= \langle d^{(i,j)}_1, \ldots, d^{(i,j)}_{m_{i,j}} \rangle$ choose all the generators be coherent with $d^{(i,j)}_1=\phi(c^{(i,j)}_1)$. Then $d^{(i,j)}_k=\phi(c^{(i,j)}_k)$, by the assertions (1) and (2) of Theorem~\ref{thm:main2}. Indeed, if $c$ and $a$ were two coherent generators of $C_{i,j}$, while $\phi_{i,j}(c),\phi_{i,j}(a)\in\phi_{i,j}(C_{i,j})$ were not coherent then $l(ac)<l(\phi_{i,j}(ac))$ would contradict (2).
\end{proof}
Observe, that if $l:G \to \Lambda$ is a Lyndon length function with values
in some ordered abelian group $\Lambda$ and $\mu:\Lambda \to \Lambda^\prime$
is a homomorphism of ordered abelian groups then the composition
$l^\prime = \mu \circ l$ gives a Lyndon length function
$l^\prime : G \to \Lambda ^\prime$. In particular, since ${\mathbb Z}^{n-1}$ is a
convex subgroup of ${\mathbb Z}^n$ then the canonical projection $\pi_n:{\mathbb Z}^n \to {\mathbb Z}$
such that $\pi_n(x_1,x_2,\ldots,x_n) = x_n$ is an ordered homomorphism, so
the composition $\pi_n \circ |\cdot|$ gives a Lyndon length function
$\lambda: G \to {\mathbb Z}$ such that $\lambda(g) = \pi_n(|g|)$. Notice also that
if $u = g \circ h$ then $\lambda(u) = \lambda(g) + \lambda(h)$ for any
$g, h, u \in G$.
If $Y$ is a finite set of elements in the group $G$ acting freely and regularly on ${\mathbb Z}^n$-tree, we denote by $Y_+$ the set of elements with non-zero length $\lambda$ and by $Y_0$ the rest of elements.
\begin{defn} \textbf{(Reduced generating set)}\label{defn:reducedset}\cite[Proposition 2]{KharlampovichMyasnikov10}.
A finite set $Y$ of elements of the group $G$ acting freely and regularly on ${\mathbb Z}^n$-tree is called reduced if it satisfies the following conditions:\begin{itemize}
\item[(a)] all elements of the subset $Y_+$ are cyclically reduced;
\item[(b)] if $f,g \in Y_+^{\pm 1},\ f \neq g$ then $\lambda(com(f,hg)) = 0$
for any $h \in \langle Y_0 \rangle$;
\item[(c)] if $f \in Y_+^{\pm 1}$ and $\lambda(com(f,h f)) > 0$ for some
$h \in \langle Y_0 \rangle$ then $\lambda(com(f,hf)) = \lambda(f)$.
\item[(d)] There are finitely many elements $\{h_1,h_2,\dots,h_m\}$ so that $\lambda(h_i)=0$ for all $i=1,2,\dots,m$,
$T=Y\cup\{h_1,h_2,\dots,h_m\}$ is reduced, and the following condition holds:
if $f \in Y_+^{\pm 1}$ and $\lambda(com(f,hf)) > 0$ for some
$h \in \langle T_0 \rangle$ then $\lambda(com(f,hf)) = \lambda(f)$ and
$f^{-1}hf \in \langle T_0 \rangle$.
\end{itemize}
\end{defn}
\begin{lem} \label{lem:reductioninnormalforms} Let $G$ be a $\mathbb{Z}^n$-free complete group obtained by a finite number of HNN extensions from a finitely generated free group $F(X)$, where $X$ is a free basis. Fix a reduced generating set $X\cup\{t_1,\dots,t_l\}$ for $G$. Let $w_1,\dots,w_r$ be a finite number of words in normal forms. If for some of these words $w=u_1t_{i_1}u_2t_{i_2}\dots u_m t_{i_m}$ $(u_j\in F(X))$ its length $l(w)$ is strictly smaller than the sum of the lengths of the letters then we can replace $t_1,\dots,t_l$ by new generators $\bar{t}_1,\dots,\bar{t}_l$ so that $w=\bar{u}_1\bar{t}_{i_1}\bar{u}_2\bar{t}_{i_2}\dots \bar{u}_m \bar{t}_{i_m}$ is a normal form and
$$l(w)=\sum_{j=1}^m (l(\bar{u}_j)+l(\bar{t}_{i_j})).$$ Moreover, the new generating set is also reduced.
\end{lem}
\begin{proof} W.l.o.g., we can assume that $l(w_j)$ is strictly smaller than the sum of the lengths of the letters in its normal form, for all $j$. Let $S=\{s_1,\dots,s_m\}$ be the set of all the stable letters of maximal height $N\leq n$ that occur in the words $w_j$, $j=1,\dots,r$; we order the elements of $S$ arbitrarily. Let $s=s_1$ conjugate $C=\langle v,t_1,\dots,t_k\rangle$ to $\phi(C)$. Then by the choice of $S$ and by Theorem~\ref{thm:main2},
$$ht(s)>ht(t_k)>\dots>ht(t_1)>ht(v).$$
We fix $j$ and in the words $w_j$ and $w_j^{-1}$ find all the occurrences of subwords of the form $d^\alpha s f^\beta$ with $d\in C$, $f\in\phi(C)$, $\alpha\neq 0$ and $\beta\neq 0$, so that
$l(d^\alpha s)<l(d^\alpha) + l(s)$ and $l(s f^\beta) < l(s) + l(f^\beta)$.
Note that for these inequalities to hold, $d^\alpha$ and $s$, as well as $s$ and $f^\beta$ should not be coherent. If $s$ and $t_k$ are coherent, this implies that $d^\alpha$ and $t_k$, as well as $t_k$ and $f^\beta$ should not be coherent, either. If there is no such $\alpha$ (or no such $\beta$) then we set $\alpha=0$ (or $\beta=0$).
Recall that by Lemma~\ref{lem:axisofstableletter}, $[x,d^px]\subset [x,sx]$ and $[x,f^px]\subset [x,s^{-1}x]$ either for every positive, or for every negative integer $p$, and that $[x,t_k^qx]\subset [x,sx]$ and $[x,\phi(t_k)^qx]\subset [x,s^{-1}x]$ either for every positive, or for every negative integer $q$. Since $t_k$ is an element of the largest height in $C$, there exist $\gamma,\delta$ so that
\begin{equation}\label{e:subsegment}
[x,d^{\alpha} x]\subseteq [x,t_k^\gamma x]'\quad [x,f^{\beta} x]\subseteq [x,(\phi(t_k))^\delta x].
\end{equation}
For all the occurrences of subwords $d^\alpha s f^\beta$ in $w_j^{\pm 1}$, we choose $\gamma_j$ and $\delta_j$ minimal in the absolute value with respect to the inclusions~(\ref{e:subsegment}).
If all $\alpha=0$ (or all $\beta=0$) then we set $\gamma_j=0$ (or $\delta_j=0$). Let $$K=\max_j\{|\gamma_j|\}, \quad M=\max_j\{|\delta_j|\}.$$
If $K=0$ then we seek a subword $hs$ with $h\notin C$ so that $l(hs)<l(h)+l(s)$. If there is such a subword in some $w_j^{\pm 1}$ then we change the value of $K$ to $K=1$. Similarly, if $M=0$ we seek subwords $sh$ with $h\notin C$ so that $l(sh)<l(h)+l(s)$ and set $M=1$ if we find one. If at least one of $K,$ $M$ is not 0, in the generating set we replace $s$ by the new letter $\bar{s}$, where $s=t_k^K\bar{s}\phi(t_k)^M$. Since in this latter product both pairs of adjacent letters translate in the same direction and $l(t_k)=l(\phi(t_k))$, it follows that
$$l(s)=l(t_k^K)+ l(\bar{s}) + l(\phi(t_k)^M)\ \Rightarrow\ l(\bar{s})= l(s)-(K+M)l(t_k).$$
We express $s$ in terms of $\bar{s}$ in every word $w_j^{\pm 1}$ and freely reduce the words whenever possible. Note that since the generating set is reduced, the free reductions only affect infinitesimal elements and never affect $\bar{s}$; in particular, the words all remain in normal forms. We also note that $ht(t_k)\geq ht(d),ht(f)$, and whenever $ht(t_k)= ht(d)$ (or $ht(t_k)=ht(f)$), both $t_k^K$ and $d^\alpha$ (or both $\phi(t_k)^M$ and $f^\beta$) cancel in the free reduction.
We proceed with $s=s_2$, then $s=s_3,\dots,s_m$, in the same way as above. Now, we consider all the stable letters of height $N-1$ and apply the same procedure (to the modified words $w_j$) to replace them by new generators, whenever needed. Note that if a stable letter $s$ in question belongs to an abelian subgroup then we can still apply the same procedure, the only difference is that $\phi(C)=C$ and $\phi$ is the trivial automorphism. The letters $t_k$ in the abelian subgroups are replaced by $t_{k-1}^K\bar{t}_k t_{k-1}^M$ and $t_1$ is replaced by $v^K\bar{t}_1v^M$, whenever applicable.
Having done with all the letters of height 2, we continue with $s=\bar{s}_1$, $s=\bar{s}_2$, etc., and keep going until free cancellations are no longer possible. The process will be finite because in every new round, when we come back to $\bar{s}_i$, the cancellations only affect infinitesimal factors of the height, which is higher than before. That is, if $l(d s_i)<l(d) + l(s_i)$ first, and then we have $l(\bar{d} \bar{s}_i)<l(\bar{d}) + l(\bar{s}_i)$, it follows that $ht(d)<ht(\bar{d})$.
All the words remain in normal forms. The length of the product of two elements is smaller than the sum of the lengths of the factors only in the case when their axes overlap and the elements translate in the opposite directions on the common segments. In the process above we remove all such occasions while keeping track of the changes of the lengths of the letters involved. The equality of the lengths from the statement of the lemma follows. The generating set remains reduced because every change of a stable letter only affects an infinitesimal part of it.
\end{proof}
The following statement is used in section ~\ref{sec:gluing} below.
\begin{thm} \textbf{(A single HNN-extension)}
\label{thm:singleHNN}
Let $G$ be a finitely generated group with a regular free Lyndon
length function in $\mathbb{Z}^n$. Then $G$ can be represented as
a union of a finite series of groups
$$G_1 < G_2 < \cdots < G_r = G,$$
where $G_1$ is a free group of finite rank, and $\forall i=1,\dots,r-1$
$$G_{i+1} = \langle G_i, s_{i} \mid s_{i}^{-1}\ C_{i}\ s_{i} = \phi_{i}(C_{i}) \rangle$$
with $C_{i}$ and $\phi_{i}(C_{i})$ maximal abelian subgroups of $G_i$,
$\phi_{i}$ an isomorphism, and so that the following conditions are satisfied:
\begin{enumerate}
\item[(1)] $\ell(\phi_{i}(w)) = \ell(w)$ $\forall w \in C_{i}$,
\item[(2)] $\phi_{i}(w)$ and $w^{-1}$ are not conjugate in $G_i$, for any $w \in C_{i}$,
\item[(3)] $C_{i} = \langle c^{(i)}_1, \ldots, c^{(i)}_{m_{i}} \rangle,$ with
$$ht(c^{(i)}_k) < ht(c^{(i)}_{k+1}),\ 1\leq k\leq m_i-1, \quad
ht(\phi_{i}(c^{(i)}_j)) = ht(c^{(i)}_j),\ 1\leq j\leq m_i;$$
\item[(4)] Generators of each $C_i$ can be chosen as follows: $c_1^{(i)}$ is a cyclically reduced element of $G_{j_1-1}$, and
$c_2^{(i)}=s_{j_1},\ldots ,c_m^{(i)}=s_{j_{(m-1)}},$ where \newline $2\leq j_1,\ldots, j_{(m-1)}<i$.
\item[(5)] Note that $C_{1}, \dots, C_i, \phi_{1}(C_{1}), \dots, \phi_{i}(C_{i})$ are subgroups of $G_i$, $\forall i$. If $A, B \in \{C_{j}, \phi_{j}(C_{j}), \mid\ 1\leq j\leq i\}$ then either $A = B$, or $A$ and $B$ are not conjugate in $G_i$;
\item[(6)] There is a reduced generating set of $G$, considered the alphabet of $G$, so that the length in ${\mathbb Z}^n$ of every generator of each centralizer $C_{i}$ is the sum of lengths of these new letters.
\end{enumerate}
\end{thm}
\begin{proof} Statements (1),(2),(3) and (5) follow from theorem~\ref{thm:main2}.
To prove (4) and (6), we start with a reduced generating set $X\cup\{t_1,\dots,t_l\}$ for $G$, where $X$ is a free basis of the free group $G_1$. We assume that the generating set of $G$ satisfies the conclusions of Lemma~\ref{lem:reductioninnormalforms}.
Proof of (4). We distinguish the following two types of the HNN-extensions:
\begin{enumerate}
\item Extensions of centralizers, when $\phi$ is the identity automorphism of the associated subgroup.
\item Extension of conjugacy classes, when $C_i$ and $\phi(C_i)$ are not conjugate in the base group $G_i$.
\end{enumerate}
Let $U=\{u_1,\dots,u_b\}\subset G_1$ be shortest representatives of those conjugacy classes of elements of the free group whose centralizers are extended in $G$. Here ``shortest" means shortest with respect to the length function; note that in the free group $G_1$ $l(u)$ equals the word length $|u|$ of $u$. So,
$$C_{G_1}(u_i)\subsetneqq C_{G}(u_i)\quad i=1,2,\dots,b,$$ and our choice of the elements of $U$ ensures that the centralizer $C_{G_1}(u_i)=\langle u_i\rangle$ is cyclically reduced. By a series of HNN-extensions we extend the centralizer $C_{G_1}(u)$ to $C_{G}(u)=\langle u,t_1,\dots,t_m\rangle$, for each $u\in U$; the sets of stable letters for distinct elements of $U$ are disjoint, and their number $m$ depends on $u$. By Theorem~\ref{thm:rankabelian}, we can assume that the inequality $ht(t_j)<ht(t_{j+1})$ holds for all $j=1,2,\dots,m-1$. Furthermore, consider all those $u\in U$ whose conjugacy classes $[u]_{G_1}$ in $G_1$ are extended in $G$, so that $[u]_{G_1}\subsetneqq[u]_{G}$. For every such $u$, we consider in $G_i$ the elements of $G_i$ that are conjugate to $u$ in $G$ but not in $G_i$. These are the elements $v$ that satisfy both $[u]_{G_i}\neq[v]_{G_i}$ and $[u]_G=[v]_G$. Let $V_i=\{v^{(i)}_1,\dots,v^{(i)}_{q_i}\}$ be the set of shortest representatives of those conjugacy classes $[v]_{G_i}$ in $G_i$. We call $u$ and the elements of $V_i$ \emph{principal elements}. Note that $V_r=\varnothing$ and $$V_1\supseteqq V_2\supseteqq\dots\supseteqq V_{r-1}.$$
Also note that every centralizer in $G$ is an abelian subgroup of $G$, and every cyclically reduced centralizer is the centralizer of a cyclically reduced element, according to Lemma~\ref{lem:reducedcentralizer}. It follows that in every associated subgroup of the HNN-extensions in the statement of Theorem~\ref{thm:main2} one can choose a principal element as above.
Theorem~\ref{thm:main2} describes how $G$ is obtained from $G_1$ by a sequence of HNN-extensions. We slightly deviate from the procedure in that at every step $i$ we choose the associated subgroups of HNN-extensions to be \emph{all} those, and only those, that are the centralizers of \emph{principal} elements. More precisely, whenever we extend the centralizer of $u$ at a step $j$, we extend the centralizer of $v$ for every $v\in V_{j+1}$, so that $C_{G_j}(u)\simeq C_{G_j}(v)$, which we do not necessarily do when following the process of Theorem~\ref{thm:main2}. (Note that if $v\in[u]_G$ then $C_{G}(v)\simeq C_{G}(u)=\langle u,t_1,\dots,t_m\rangle$.)
Eventually, we extend the centralizer $C_{G_1}(v)$ to $C_{G}(v)=\langle v,s_1,\dots,s_m\rangle$, for each $v\in V_1$; the sets of stable letters are all disjoint. We shall always assume that $[\tilde{u} t_j]_G=[\tilde{v} s_j]_G$, for some $\tilde{u}\in\langle u,t_1,\dots,t_{j-1}\rangle$ and $\tilde{v}\in\langle v,s_1,\dots,s_{j-1}\rangle$, for all $j$.
Having extended the conjugacy classes of elements from $G_1$, we apply the same procedure to $G_2$, then to $G_3$, etc. For each one of these $G_i$, we look for elements whose centralizers do not contain elements of height less than $i$; shortest representatives of their conjugacy classes will have height $i$.
By changing the process, we might have added more stable letters than appear in Theorem~\ref{thm:main2}. We need to assign a value of the length function $l$ to each one of the new letters. The new generators, that we might have introduced, correspond to existing elements of $G$ written as products in the old generators, and we could assign the lengths of the new generators correspondingly. However, for the assertion (6) to be true, we need to achieve a little more. In some cases, we will replace the generators of $C_{i,j}$ by our new stable letters whose length will be adjusted.
We proceed as follows. Let $C_{i,j}$ from the statement of theorem~\ref{thm:main2} be conjugate to $C_{G}(u)$ in $G$. Fix a generator $c_k^{(i,j)}$ of $C_{i,j}$. It follows from the construction that $c_k^{(i,j)}=wp$ for some stable letter $p$ and $w\in C_{i,j}$ so that $l(w)\ll l(p)$. We collect all the stable letters $p_1,p_2,\dots$ that appear as factors in the conjugates of $c^{i,j}$ in all the HNN-extensions with $C_{i,j}$ as an associated subgroup. In the finite set of elements, that we obtain in this way, we find an element of the shortest length $L_t$. In $\langle u,t_1,\dots,t_{k}\rangle$, there is a unique element $z$ of length $l(z)=L_t$; we choose this element to be a new generator: $t_k:=z$. We do this for all the generators in every centralizer $C_G(u)$ with $u\in U$. Then we express all the stable letters in the generating set of $G_i$ in terms of the new stable letters, and adjust the elements of $G$ accordingly. Note
that since we have chosen $t_k$ to have the minimum length, every stable letter $p$ in the conjugates of $C_{i,j}$ with $ht(p)=ht(t_k)$ will be replaced by the product $\tilde{u}t_k$ where for all $l$, $[x,\tilde{u}^lx]\subset A_{t_k}$ and $\tilde{u}$ translates points of those common segments to the same direction as $t_k$ does.
The generating set $Z$ that we obtain is different from the generating set $Y$ obtained in theorem~\ref{thm:main2}; for instance, $Z_0$ contains more elements than $Y_0$. However, the elements of $Z_+$ and those of $Y_+$ can only differ by infinitesimal factors, hence $Z$ is also reduced.
(6) By Lemma~\ref{lem:reductioninnormalforms}, we can assume that for $w_1,\dots,w_r$, whose centralizers are extended in $G$, the length $l(w_i)$ is the sum of the lengths of the old letters. The choice of the new generators for the abelian subgroups $C_{i,j}$ of $G$ does not lead to any length reduction in the words $w_1,\dots,w_r$. Therefore $l(w_j)$ remains equal to the sum of the lengths of the new letters.
\end{proof}
\subsection{Weighted word metric}
In the proof of Key Lemma below we use the following result.
\begin{prop}\cite{GK} \label{GK} Any two nontrivial ordered abelian groups
satisfy the same existential sentences in the language
$L=\{0,+,-,<\}.$
\end{prop}
\begin{lem} \textbf{(Key Lemma)} Let $G$ be a group equipped with a free regular length function with values in $\mathbb{Z}^n$. One can assign positive integer weights to the generators of $G$ so that in every HNN-extension $t^{-1}Ct=\phi(C)$ in the statement of theorem~\ref{thm:singleHNN}, for every element $g\in C$ we have that the weights of $g$ and of $\phi(g)$ are equal. By the weight of a product we mean the sum of the weights of the factors.
\end{lem}
\begin{proof} According to theorem~\ref{thm:singleHNN}, $G$ is obtained from the finitely generated free group $G_1=F(X)$ by finitely many HNN-extensions with finitely generated associated subgroups and stable letters $t_1,\dots,t_k$. We fix a generating set $Y=X\cup\{t_1,\dots,t_k\}$ for $G$ as in theorem~\ref{thm:singleHNN}(7). Since the associated subgroups are free abelian groups and the generatoring set of each one of them is a free basis, it suffices to arrange for the weight of every generator be equal the weight of its image. Let $U$ denote the set of all the generators of the associated subgroups of all the HNN extensions; we think of them as pairs $u$ and $v=\phi(u)$. $U$ is a finite set. For every such generator $u=g_1\dots g_m\in U$, $(g_i\in Y^{\pm 1})$ and its image $$\phi(u)=v=f_1\dots f_{m_v}$$ we have $l(u)=l(v)$ by the assertion~\ref{thm:singleHNN}(1), and
\[
l(u)= \sum_{i=1}^{m_u} l(g_i),\quad l(v) = \sum_{i=1}^{m_v} l(f_i),
\]
theorem~\ref{thm:singleHNN}(7). Therefore, we can write the following finite system of equations and inequalities that are true in $\mathbb{Z}^n$:
\begin{eqnarray*}
\sum_{i=1}^{m_u} l(g_i) = \sum_{i=1}^{m_v} l(f_i) \\
l(g)>0,\ \forall g\in Y,
\end{eqnarray*}
where $u$ and $v$ run over all the (pairs of) elements of $U$. Note that we use no constants when writing these. Therefore, we have a finite set of formulas in the language $\{0,+,-,<\}$ satisfiable in the nontrivial ordered abelian group $\mathbb{Z}^n$. Proposition~\ref{GK} implies that these formulas are satisfiable in $\mathbb{Z}$. This allows us to define a weighted word metric $wm\colon G\rightarrow\mathbb{Z}^+\cup\{0\}$ so as to have the system of equations and inequalities in $\mathbb{Z}$, as follows:
\begin{eqnarray}\label{e:equation}
\sum_{i=1}^{m_u} wm(g_i) = \sum_{i=1}^{m_v} wm(f_i).
wm(g)>0,\ \forall g\in Y.
\end{eqnarray}
There is a weighted Cayley graph of a group $G$, associated with the weighted metric $wm$. In the weighted graph the length of an edge labeled by a generator $g\in Y$ is $wm(g)$. The values of $wm$ on the generating set $Y$ are defined by the equations~(\ref{e:equation}), $wm(1)=0$ by definition, and for an arbitrary element $h\in G$, its weighted length
\[
wm(h)=\min_{g_1\dots g_{m_h}=h}\{\sum_{i=1}^{m_h}wm(g_i)\}
\]
is the shortest (weighted) distance from 1 to $h$ in the weighted Cayley graph of $G$. Since the normal forms that we use for the HNN extensions and the normal forms of elements in the abelian subgroups of $G$ are invariant with respect to the change of word metric, $wm(u)= wm(v)$, for all the (pairs of) elements $u,v\in U$.
\end{proof}
\section{Gluing CAT(0) spaces} \label{sec:gluing}
In this section, following the strategy introduced in~\cite{AlibegovicBestvina}, we prove that $\mathbb{Z}^n$-free groups are CAT(0).
The first ingredient of the proof is the following statement.
\begin{prop} \cite[Proposition II.11.13]{BridsonHaefliger} \label{prop:gluing} Let $X$ and $A$ be locally CAT(0) metric spaces. If $A$ is compact and $\varphi,\phi\colon A\rightarrow X$ are local isometries, then the quotient of $X\coprod(A\times[0,1])$ by the equivalence relation generated by $$[(a,0)\sim\varphi(a);
(a,1)\sim\phi(a)],$$ for all $a\in A$, is locally CAT(0).
\end{prop}
To state the following theorem, we should remind the definition of a geometrically coherent space, inspired by the work of Wise~\cite{WiseSectionalCurv}, and introduced in \cite{AlibegovicBestvina}.
\begin{defn}\label{defn:coherentspace} \textbf{(Geometrically coherent space)} Let $X$ be a connected locally CAT(0) space, and let $C$ be a connected subspace of $X$. $C$ is a \emph{core} of $X$ if $C$ is compact, locally CAT(0), and the inclusion $C\hookrightarrow X$ induces a $\pi_1$-isomorphism.
Let $Y$ be a connected locally CAT(0) space. $Y$ is called \emph{geometrically coherent} if every covering space $X\rightarrow Y$ with $X$ connected and $\pi_1(X)$ finitely generated has the following property. For every compact subset $K\subset X$ there is a core $C$ of $X$ containing $K$.
\end{defn}
\begin{defn} \textbf{(Maximal separated torus)}
Let $U$ be a connected geodesic metric space and $A\subset U$ be a $k$-torus, so that $A=c_1\times c_2\times\dots\times c_k$ for some non-homotopic simple closed curves $c_1, c_2,\dots, c_k\subset U$. We say that $A$ is a \emph{maximal torus in} $U$ if $A$ is not contained in any bigger torus $T\subset U$. A maximal torus $A$ is \emph{separated in} $U$ if for any torus $\hat A\subset U$ so that $A\cap\hat A\neq\emptyset$ it follows that $\hat A\subseteq A$.
\end{defn}
We are interested in the case when every torus of $U$ is contained in a maximal torus, and all tori maximal in $U$ are separated in $U$. In the language of fundamental groups, this means that every abelian subgroup of $\pi_1(U)$ is contained in a maximal abelian subgroup, and maximal abelian subgroups are conjugacy separated; in this case $\pi_1(U)$ is called a CSA group. Being relatively hyperbolic, $\mathbb{Z}^n$-free groups are CSA.
In the statement below, by a 1-torus we mean a simple closed curve.
\begin{thm}\label{thm:finitecore} Let $U$ be a geometrically coherent locally CAT(0) space, let $T$ be a $k$-torus for some integer $k\geq 1$, and let $\varphi\colon T\rightarrow U$ and $\phi\colon T\rightarrow U$ be local isometries. Furthermore, assume that both images $\varphi(T)$ and $\phi(T)$ are locally convex, maximal and separated in $U$. We also assume that if $\varphi(T)\neq\phi(T)$ then the $\pi_1(U)$-orbits of $\varphi(T)$ and $\phi(T)$ are distinct.
Let $Y$ be the quotient of
\[
U\coprod T\times[0,1]
\]
by the equivalence relation generated by $$[(a,0)\sim\varphi(a),\ (a,1)\sim\phi(a),\forall a\in T].$$
Then $Y$ is geometrically coherent.
\end{thm}
\begin{proof} Let us be given a covering space $p\colon X\rightarrow Y$ with $X$ connected and $\pi_1(X)$ finitely generated, and a compact subset $K\subset X$. Then $X$ can be viewed as a graph of spaces with vertex spaces the components of $p^{-1}(U)$ and edge spaces the components of $p^{-1}(T\times{\frac 12})$.
Let $G=\pi_1(X)$ be the fundamental group of $X$, and let $\Gamma$ be the associated graph of groups. Then $\pi_1(\Gamma)=\pi_1(X)=G$; note that $G$ is finitely generated. Every edge group in $\Gamma$ is either free abelian of rank less or equal to $k$, or trivial. Every boundary component $A$ of each edge space is the product of lines and circles: $$A=c_1\times c_2\times\dots\times c_p\times\mathbb{R}^q,$$ where $p+q=k\geq 0$ and each $c_i$ is a circle. In the vertex space, circles are identified with simple closed curves, and lines are identified with lines, so that each boundary component of every vertex space is identified with either a torus, the product of a torus with lines, or the product of several lines. It may happen that both boundary components of an edge space are glued to the same vertex space.
Note that our assumptions on the images $\varphi(T)$ and $\phi(T)$ of $T$ imply that if $\varphi(T)\neq\phi(T)$ then the boundary components of distinct edge spaces in $X$ are identified with disjoint subsets of the vertex spaces. This is because maximal tori of $U$ are separated in $U$ and the $\pi_1$-orbits of $\varphi(T)$ and $\phi(T)$ are disjoint.
First, assume that the graph of spaces $X$ is finite; in particular, $X$ contains finitely many vertex and edge spaces. In this case, the graph of groups $\Gamma$ is also finite. It follows that every vertex group of $\Gamma$ is finitely generated. Therefore, in any vertex space of $X$, every compact subset is contained in a core. Given the compact set $K$ in $X$, to find a core $C\supset K$ in $X$, we first select the product of circles $c_j$ and segments $[a_l,b_l]$ with the unit interval $I=[0,1]$
$$B_E=c_1\times\dots\times c_p\times[a_1,b_1]\times\dots\times[a_q,b_q]\times[0,1]$$
inside each product $E=c_1\times\dots\times c_p\times\mathbb{R}^q\times[0,1]$ of $p$ circles and $q$ lines with $I$, so that $K\cap E\subset B_E$.
For every vertex space $V$ of $X$ we choose a core $C(V)$ so that it contains the intersection $V\cap K$ with $K$; $C(V)$ contains the intersection $V\cap B_E$ for every edge $E=A\times I$ in $X$; and finally, $C(V)$ contains the intersection $V\cap T_{GA}$ for every generalized annulus $GA=T_{GA}\times I$ in $X$.
The core $C$ is the union of the following spaces:
\begin{itemize}
\item the cores $C(V)$ of all vertex spaces,
\item all the generalized annuli $GA$ in $X$, and
\item all the products $B_E$ in $X$.
\end{itemize}
$C$ is clearly compact as the finite union of compact spaces. $C$ is locally CAT(0), according to Proposition~\ref{prop:gluing}. $C$ is connected by construction. Furthermore, the graph of groups $\Gamma_C$ corresponding to $C$ has the vertex groups isomorphic to those of $\Gamma$, and the same edge groups as $\Gamma$, with the same monomorphisms, so $$\pi_1(C)=\pi_1(\Gamma_C)=\pi_1(\Gamma)=\pi_1(X).$$
Now, consider the case when $X$ is an infinite graph of spaces. In this case the corresponding graph $\Gamma$ is also infinite. We need a preliminary step. If $K$ intersects vertex spaces, pick an arbitrary vertex space $V$ with $V\cap K\neq \emptyset$; otherwise, choose any vertex space $V$ in $X$. Starting with $V$, build an ascending chain of finite connected subgraphs $X_0\subset X_1\subset\dots\subset X$ of $X$ which exhaust $X$, and consider the corresponding subgraphs $\Gamma_0\subset\Gamma_1\subset\dots\subset \Gamma$ of $\Gamma$. The fundamental groups $G_0\subset G_1\subset\dots$ of these subgraphs form an ascending chain of subgroups of $G=\pi_1(\Gamma)$. The subgraphs $\Gamma_i$ exhaust $\Gamma$; therefore, the subgroups $G_i$ exhaust $G$. Since $G$ is finitely generated, the chain of subgroups must stabilize after finitely many steps: $$G_0\subset G_1\subset\dots G_n=G_{n+1}=\dots=G.$$ Therefore, the finite subgraph $\Gamma_n$ has the same fundamental group as $\Gamma$, and so for the corresponding subspace $X_n$ we have that $\pi_1(X_n)=\pi_1(X)$. Given a compact subset $K$, we may need to replace $X_n$ by a larger subgraph $X_m$ so as to include $K$. Clearly, a core in $X_m$ is a core in $X$, so we can apply to $X_m$ the procedure described above to find a core.
\end{proof}
\begin{thm}\label{thm:regularZnfreecoherent} Let $G$ be a finitely generated group with a free regular $\mathbb{Z}^n$-length function. Then $G$ is the fundamental group of a geometrically coherent space. In particular, every finitely generated subgroup of $G$ is CAT(0).
\end{thm}
\begin{proof} We build a sequence of spaces corresponding to the chain of HNN-extensions in theorem~\ref{thm:singleHNN}.
$U_0$ is the wedge of circles corresponding to the free group $F=G_0$. Note that every maximal abelian subgroup $C$ of $G_0$ is cyclic; let $C=\langle c^{(0)}\rangle$. Let $\alpha$ and $\beta$ be the reduced paths in $U_0$ corresponding to $c^{(0)}$ and $\phi(c^{(0)})$. Note that the lengths $l(\alpha)$ and $l(\beta)$ are equal. We rescale the metric on $S^1$ so as to make the total length of the circle equal $l(\alpha)$, and let $\varphi\colon S^1\rightarrow \alpha$ and $\phi\colon S^1\rightarrow \beta$ be local isometries. $U_1$ is the quotient of
\[
U_0\coprod S^1\times[0,1]
\]
by the equivalence relation generated by $[(a,0)\sim\varphi(a),\ (a,1)\sim\phi(a),\forall a\in S^1]$.
If $c^{(0)}\neq \phi(c^{(0)})$, $c^{(0)}$ and $\phi(c^{(0)})$ are not conjugate in $G_0$; therefore, in this latter case the $\pi_1(U_0)$-orbits of $\alpha$ and $\beta$ are disjoint. It can be readily seen that all the conditions of Theorem~\ref{thm:finitecore} are satisfied. It follows that $U_1$ is geometrically coherent.
To proceed, assume that $G_i=\pi_1(U_i)$, where $U_i$ is a geometrically coherent space. Let $c^{(i)}_1, \ldots, c^{(i)}_{m_{i}}$ be generators of a maximal abelian subgroup $C_i$ of $G_i$, chosen so that the conditions of theorem~\ref{thm:singleHNN} are satisfied.
Let $\alpha_j$ and $\beta_j$ be the reduced paths in $U_i$ corresponding to $c_j^{(i)}$ and $\phi(c_j^{(i)})$. Note that the lengths $l(\alpha_j)$ and $l(\beta_j)$ are equal. We define $$T=\prod_{j=1}^{m_i} S^1_j$$ and rescale the metric on each $S^1_j$ so as to make the total length of the circle equal $l(\alpha_j)$. This defines the metric on $T$. Let $\varphi\colon S^1_j\rightarrow \alpha_j$ and $\phi\colon S^1_j\rightarrow \beta_j$ be local isometries. They extend to local isometries $T\rightarrow \varphi(T)$ and $T\rightarrow \phi(T)$. $U_{i+1}$ is the quotient of
\[
U_i\coprod T\times[0,1]
\]
by the equivalence relation generated by $$[(a,0)\sim\varphi(a),\ (a,1)\sim\phi(a),\forall a\in T].$$
All the conditions of Theorem~\ref{thm:finitecore} are satisfied. It follows that $U_{i+1}$ is geometrically coherent.
By induction on the number $n$ of HNN-extensions in theorem~\ref{thm:singleHNN}, we obtain the claim.
\end{proof}
By~\cite{KharlampovichMyasnikov11}, every group acting freely on a $\mathbb{Z}^n$-tree embeds by a length preserving monomorphism into a regular $\mathbb{Z}^n$-free group. Therefore, we have the following result.
\begin{thm}\label{thm:ZnfreeCAT0} Let $G$ be a finitely generated group acting freely on a $\mathbb{Z}^n$-tree. Then $G$ acts properly and cocompactly on a CAT(0) space with isolated flats.
\end{thm}
\subsection*{Acknowledgement} The authors wish to thank Denis Serbin for helpful conversations.
\bibliographystyle{plain}
|
\part{Requirements}
\label{part:requirements}
This document has been written to help in the choice of a linear algebra
library to be included in Verdandi, a scientific library for data
assimilation. The main requirements are
\begin{enumerate}
\item Portability: Verdandi should compile on BSD systems, Linux, MacOS,
Unix and Windows. Beyond the portability itself, this often ensures that
most compilers will accept Verdandi. An obvious consequence is that all
dependencies of Verdandi must be portable, especially the linear algebra
library.
\item High-level interface: the dependencies should be compatible with the
building of the high-level interface (e. g. with SWIG, this implies that
headers (.hxx) have to be separated from sources (.cxx)).
\item License: any dependency must have a license compatible with Verdandi
licenses (GPL and LGPL).
\item C++ templates, sparse matrices and sparse vectors have to be supported.
\end{enumerate}
Information reported here was collected from December 2008 to March 2009.
\newpage
\part{CPPLapack}
CPPLapack is a C++ class wrapper for Blas and Lapack.
\libtab {CPPLapack} {beta} {Mar. 2005} {Mar. 2005}
{http://cpplapack.sourceforge.net/} {in} {4} {Apr. 2004} {GPL (}
\libdesc {-- Real double-precision and complex double-precision vectors and
matrices \\
-- Real double-precision and complex double-precision band, symmetric and
sparse matrices}
{-- Eigenvalues computation \\
-- Linear systems solving \\
-- SVD decomposition}
{-- Blas \\
-- Lapack}
{Almost the same as the performance of original Blas and Lapack}
{Platform independent}
{Requires Blas and Lapack}
{-- A few bugs and unsupported Blas and Lapack functions}
{-- Templates not supported, float not supported \\
-- Alpha version of sparse matrix classes with bugs (the authors advise
developers not to use these classes in their code) \\
-- No sparse vectors \\
-- No separation between headers and sources}
\newpage
\part{Eigen}
Eigen is a C++ template library for linear algebra, part of the KDE project.
\libtab {Eigen} {2.0-beta6} {soon released} {Jan. 2009}
{http://eigen.tuxfamily.org/} {in} {7 contributors, 2} {Dec. 2006} {LGPL and
GPL (}
\libdesc {-- Dense and sparse matrices and vectors \\
-- Plain matrices/vectors and abstract expressions \\
-- Triangular and diagonal matrices \\
-- Column-major (the default) and row-major matrix storage}
{-- Triangular, SVD, Cholesky, QR and LU solvers \\
-- Eigen values/vectors solver for non-selfadjoint matrices \\
-- Hessemberg decomposition \\
-- Tridiagonal decomposition of a selfadjoint matrix}
{For sparse matrices: TAUCS, umfpack, cholmod and SuperLU}
{Very efficient, see benchmark:
\url{http://eigen.tuxfamily.org/index.php?title=Benchmark}}
{Standard C++ 98, compatible with any compliant compiler such as \\
-- GCC, version 3.3 and newer \\
-- MSVC (Visual Studio), 2005 and newer \\
-- ICC, recent versions \\
-- MinGW, recent versions}
{No dependency}
{-- Templates supported \\
-- Todo includes: interface to Lapack and eigensolver in non-selfadjoint case
\\
-- Examples of users: \\
KDE related projects such as screensavers, kgllib, kglengine2d,
solidkreator, painting and image editing \\
Avogadro, an opensource advanced molecular editor \\
VcgLib, C++ template library for the manipulation and processing of
triangular and tetrahedral meshes \\
MeshLab, for the processing and editing of unstructured 3D triangular meshes
and point cloud \\
The Yujin Robot company uses Eigen for the navigation and arm control of
their next generation robots (switched from blitz, ublas and tvmet)}
{-- Sparse matrices and vectors still experimental \\
-- Eigen 2 is a beta version (Eigen 1 is the old stable version)}
\newpage
\part{Flens}
Flens (Flexible Library for Efficient Numerical Solutions) is a C++ library
for scientific computing providing interface for Blas and Lapack. Flens
intends to be the building block of choice for the creation of serious
scientific software in C++.
\libtab {Flens} {RC1} {Jul. 2007} {Feb. 2008}
{http://flens.sourceforge.net} {in} {9} {2004} {BSD License
(}
\libdesc {-- General, triangular and symmetric matrix types \\
-- Storage formats: full storage (store all elements), band
storage (store only diagonals of a banded matrix), packed storage (store
only the upper or lower triangular part) \\
-- Sparse matrix types: general and symmetric, compressed row storage;
random access for initialization}
{-- Linear systems solving using QR factorization \\
-- Cg and Pcg methods}
{Blas and Lapack}
{-- Natural mathematical notation: e. g. \code{y += 2 * transpose(A) * x + 1.5
* b + c} without sacrificing performances (see section
\ref{sec:flens_overloaded} in Appendix)\\
-- Very efficient, see benchmarks:
\url{http://flens.sourceforge.net/session2/tut4.html},
\url{http://grh.mur.at/misc/sparselib_benchmark/report.html},
\url{http://flens.sourceforge.net/session1/tut9.html} and section
\ref{sec:benchmark_seldon}}
{-- Tested on Mac OS X, Ubuntu Linux and a SUSE Opteron cluster \\
-- GCC: version 4 or higher \\
-- Intel C++ compiler (icc): version 9.1 \\
-- Pathscale (pathCC): GNU gcc version 3.3.1, PathScale 2.3.1 driver}
{Requires Blas, Lapack and CBlas}
{-- Extensible: e. g. easy integration of user-defined matrix/vector types \\
-- Flexible: e. g. generic programming of numerical algorithms \\
-- Flens implements a view concept for vectors and dense matrices: a
vector can reference a part of a vector, or also a row, column or diagonal of a
matrix; you can apply to these views the same operations as for regular
vectors and matrices. \\
-- Templated matrices and vectors with several storage formats}
{-- Lack of portability: not recently tested on Windows (once compiled with
Microsoft Visual Studio Express compiler with minor modifications in the
Flens code) \\
-- No eigenvalues computation \\
-- No sparse vectors \\
-- No hermitian matrices}
\newpage
\part{Gmm++}
Gmm++ is a generic matrix template library inspired by MTL and ITL.
\libtab {Gmm++} {3.1} {Sep. 2008} {Sep. 2008}
{http://home.gna.org/getfem/gmm_intro.html} {in} {2 contributors and 2}
{Jun. 2002} {LGPL (}
\libdesc {Sparse, dense and skyline vectors and matrices}
{-- Triangular solver, iterative generic solvers (Cg, BiCgStag, Qmr, Gmres)
with preconditioners for sparse matrices (diagonal, based on MR iterations,
ILU, ILUT, ILUTP, ILDLT, ILDLTT) \\
-- Reference to sub-matrices (with sub-interval, sub-slice or sub-index) for
any sparse dense or skyline matrix for read or write operations \\
-- LU and QR factorizations for dense matrices \\
-- Eigenvalues computation for dense matrices}
{-- Blas, Lapack or Atlas for better performance \\
-- SuperLU 3.0 (sparse matrix direct solver) for sparse matrices}
{Very efficient, see benchmarks:
\url{http://grh.mur.at/misc/sparselib_benchmark/report.html} and
\url{http://eigen.tuxfamily.org/index.php?title=Benchmark}}
{-- Linux/x86 with g++ 3.x and g++ 4.x \\
-- Intel C++ Compiler 8.0 \\
-- Linux/Itanium with g++ \\
-- MacOS X Tiger (with the python and matlab interface) \\
-- Windows with MinGW and MSys (without the Python and Matlab interface)}
{No special requirement}
{-- Templates supported \\
-- Examples of users: IceTools, an open source model for glaciers; EChem++:
a problem solving environment for electrochemistry \\
-- Gmm++ is included in Getfem++, a generic and efficient C++ library for
finite element methods, awarded by the second price at the "Trophées du
Libre 2007" in the category of scientific softwares \\
-- Provides a high-level interface to Python and Matlab via Mex-Files for
Getfem++, covering some functionalities of Gmm++
}
{-- No separation between headers and sources (only header files) \\
-- No eigenvalues computation for sparse matrices \\
-- Gmm++ primary aim is not to be a standalone linear algebra library, but
is more aimed at interoperability between several linear algebra packages.}
\newpage
\part{GNU Scientific Library (GSL)}
GSL is a numerical library for C and C++ programmers. The library provides a
wide range of mathematical routines covering subject areas such as linear
algebra, Blas support and eigensystems.
\libtab {GSL} {GSL-1.12} {Dec. 2008} {Dec. 2008}
{http://www.gnu.org/software/gsl/} {in} {18} {1996} {GPL (}
\libdesc {General vectors and matrices}
{-- Eigenvalues and eigenvectors computation \\
-- Functions for linear systems solving: LU Decomposition, QR Decomposition,
SVD Decomposition, Cholesky Decomposition, Tridiagonal Decomposition,
Hessenberg Decomposition, Bidiagonalization, Householder Transformations,
Householder solver for linear systems, Tridiagonal Systems, Balancing }
{-- Blas (level 1, 2 and 3)\\
-- CBlas or Atlas \\
-- Many extensions such as Marray, NEMO, LUSH (with full interfaces to GSL,
Lapack, and Blas) and PyGSL}
{Not evaluated}
{-- GNU/Linux with gcc
-- SunOS 4.1.3 and Solaris 2.x (Sparc)
-- Alpha GNU/Linux, gcc
-- HP-UX 9/10/11, PA-RISC, gcc/cc
-- IRIX 6.5, gcc
-- m68k NeXTSTEP, gcc
-- Compaq Alpha Tru64 Unix, gcc
-- FreeBSD, OpenBSD and NetBSD, gcc
-- Cygwin
-- Apple Darwin 5.4
-- Hitachi SR8000 Super Technical Server, cc}
{Easy to compile without any dependencies on other packages}
{}
{-- Can be called from C++ but written in C \\
-- No sparse matrices and vectors \\
-- Templates not supported}
\newpage
\part{IT++}
IT++ is a C++ library of mathematical, signal processing and communication
routines. Templated vector and matrix classes are the core of the IT++
library, making its functionality similar to that of MATLAB and GNU
Octave. IT++ makes an extensive use of existing open source libraries
(e. g. Blas, Lapack, ATLAS and FFTW).
\libtab {IT++} {4.0.6} {Oct. 2008} {Oct. 2008} {http://itpp.sourceforge.net/}
{in} {19 contributors and 11} {2001} {GPL (not L}
\libdesc {-- Diagonal, Jacobsthal, Hadamard and conference matrices \\
-- Templated vectors and matrices \\
-- Sparse vectors and matrices}
{-- Matrix decompositions such as eigenvalue, Cholesky, LU, Schur, SVD and
QR \\
-- Linear systems solving: over- and underdetermined, LU factorization and
Cholesky factorization}
{-- Blas, Lapack and FFTW\\
-- Optionally Atlas, MKL and ACML}
{Not evaluated}
{GNU/Linux, Sun Solaris, Microsoft Windows (with Cygwin, MinGW/MSYS or
Microsoft Visual C++) and Mac OS X}
{Packages available (Fedora RPM, Debian GNU/Linux and openSUSE)}
{-- Templates supported \\
-- Separation between headers and sources \\
}
{-- Its main use is in simulation of communication systems and for performing
research in the area of communications. It is also used in areas such as
machine learning and pattern recognition. \\
-- One important high-performance feature missing in IT++ is the ability to
create "submatrix-views" and "shallow copies" of matrices, i.e. pass
submatrices by reference -instead of by value (compared to Lapack++). }
\newpage
\part{Lapack++}
Lapack++ is a library for high performance linear algebra computations.
\libtab {Lapack++} {beta 2.5.2} {Jul. 2007} {Jul. 2007}
{http://lapackpp.sourceforge.net/} {in} {7} {1993} {LGPL (}
\libdesc {-- Int, long int, real and complex vectors and matrices \\
-- Symmetric positive definite matrix \\
-- Symmetric, banded, triangular and tridiagonal matrices}
{-- Linear systems solving for non-symmetric matrices, symmetric positive
definite systems and solving linear least-square systems; using LU, Cholesky
and QR matrix factorizations \\
-- Symmetric eigenvalues computation \\
-- SVN and QR decompositions}
{-- Blas \\
-- Lapack}
{High performance linear algebra computation}
{-- Linux/Unix: gcc2.95.x, gcc3.x and gcc4.x \\
-- Windows 9x/NT/2000: MinGW and gcc3.x \\
-- Windows 9x/NT/2000: Microsoft Visual Studio, .NET and MSVC \\
-- Mac OS X}
{Requires Blas, Lapack and a Fortran compiler}
{-- Template functions for matrices}
{-- Templates not supported, float supported only for general matrices \\
-- No sparse matrices \\
-- No sparse vectors}
\newpage
\part{Matrix Template Library (MTL)}
MTL is a generic component library developed specially for high
performance numerical linear algebra. MTL includes matrix formats and
functionality equivalent to level 3 Blas.
\libtab {MTL 4} {alpha 1} {Oct. 2007} {Nov. 2008}
{http://www.osl.iu.edu/research/mtl/mtl4/} {in} {4} {1998 (MTL2)} {Copyright
Indiana University (can be modified to become }
\libdesc {-- Dense2D, morton\_dense and sparse matrices \\
-- Arbitrary types can be used for matrix elements (float, double, complex)}
{-- Preconditioners: diagonal inversion, incomplete LU factorization without
fill-in and incomplete Cholesky factorization without fill-in \\
-- Solvers: triangular, conjugate gradient, BiCg, CgSquared and BiCgStab \\
-- Iterative methods for solving linear systems thanks to the Iterative
Template Library (ITL, last release in Oct. 2001): Chebyshev and Richardson
iterations, generalized conjugate residual, generalized minimal residual and
(transpose free) quasi-minimal residual without lookahead}
{Blas (optionally Blitz++ thanks to ITL)}
{-- Natural mathematical notation without sacrificing performances: A = B * C
dispatched to the appropriate Blas algorithm if available; otherwise an
implementation in C++ is provided (also reasonably fast, usually reaching 60
percent peak) \\
-- See benchmarks:
\url{http://grh.mur.at/misc/sparselib_benchmark/report.html},
\url{http://eigen.tuxfamily.org/index.php?title=Benchmark} and
\url{http://projects.opencascade.org/btl/}}
{Can be compiled and used on any target platform with an ANSI C++ compiler \\
-- Linux: g++ 4.0.1, g++ 4.1.1, g++ 4.1.2, g++ 4.2.0, g++ 4.2.1, g++ 4.2.2,
icc 9.0 \\
-- Windows: VC 8.0 from Visual Studio 2005 \\
-- Macintosh: g++ 4.0.1}
{Requires the Boost library included, optionally scons and a Blas library
installed}
{-- Templates and generic programming \\
-- A generic library has been built on top of MTL (ITL:
\url{http://www.osl.iu.edu/research/itl/})\\
-- Developed from scratch but inspired by the design and implementation
details of MTL 2 (interfacing Lapack; supporting sparse, banded,
packed, diagonal, tridiagonal, triangle, symmetric matrices)}
{-- No sparse vectors \\
-- No eigenvalues computation \\
-- No release since the alpha version in Oct. 2007}
\newpage
\part{PETSc}
PETSc, the Portable, Extensible Toolkit for Scientific computation, is a suite
of data structures and routines for the scalable (parallel) solution of
scientific applications modeled by partial differential equations.
\libtab {PETSc} {2.3.3}{May 2007}{Jul. 2008}
{http://www.mcs.anl.gov/petsc/petsc-2/} {in} {11} {Mar. 1999} {Copyright
University of Chicago (}
\libdesc {-- Parallel vectors and matrices \\
-- Several sparse matrices storages \\
-- Symmetric, block diagonal and sequential matrices}
{-- Preconditioners: ILU, LU, Jacobi, block Jacobi, additive Schwartz and ICC
\\
-- Direct solvers: LU, Cholesky and QR \\
-- Krylov subspace methods: GMRES, Chebychev, Richardson, conjugate
gradients (Cg), CGSquared, BiCgStab, two variants of TFQMR, conjugate
residuals and Lsqr \\
-- Nonlinear solvers: Newton-based methods, line search and trust region \\
-- Parallel timestepping solvers: Euler, Backward Euler and pseudo time
stepping}
{Blas, Lapack, ADIC/ADIFOR, AnaMod, BlockSolve95, BLOPEX, Chaco, DSCPACK,
ESSL, Euclid, FFTW, Hypre, Jostle, Mathematica, Matlab, MUMPS, ParMeTiS,
Party, PaStiX, PLapack, Prometheus, Scotch, SPAI, SPOOLES, SPRNG,
Sundial/CVODE, SuperLU, Trilinos/ML, UMFPACK}
{Optimal on parallel systems with high per-CPU memory performance}
{Any compiler supporting ANSI C standard on Unix or Windows}
{Requires Blas, Lapack, MPI and optional packages}
{-- Related packages using PETSc such as TAO, SLEPc, Prometheus, OpenFVM,
OOFEM, DEAL.II and Python bindings (petsc4py and LINEAL) \\
-- Scientific applications in many fields such as nano-simulations,
cardiology, imaging and surgery, geosciences, environment,
computational fluid dynamics, wave propagation and optimization}
{-- No sparse vectors \\
-- Not coded in C++ but in C \\
-- Templates not supported, polymorphism used instead}
\newpage
\part{Seldon}
Seldon is a C++ library for linear algebra. Seldon is designed to be efficient
and convenient, which is notably achieved thanks to template classes.
\libtab {Seldon} {2008-11-12} {Nov. 2008} {Nov. 2008}
{http://seldon.sourceforge.net/} {in} {2} {Nov. 2004} {GPL (}
\libdesc {-- Dense and sparse vectors \\
-- Dense matrices: several formats for rectangular, symmetric, hermitian and
triangular \\
-- Two sparse matrix forms: Harwell-Boeing and array of sparse vectors \\
-- 3D arrays}
{-- Preconditioner of your own or by successive over-relaxations (SOR) \\
-- Direct dense solvers: LU, QR, LQ and SVD decomposition \\
-- Iterative solvers: BiCg, BiCgcr, BiCgStab, BiCgStabl, Cg, Cgne, Cgs,
CoCg, Gcr, Gmres, Lsqr, MinRes, QCgs, Qmr, QmrSym, SYMmetric, Symmlq and
TfQmr
\\
-- Eigenvalues and eigenvectors computation}
{-- Fully interfaced with Blas (level 1, 2 and 3) and Lapack, except for
functions with banded matrices \\
-- Direct sparse solvers: MUMPS, SuperLU and UmfPackLU}
{Very efficient, see benchmarks in section \ref{sec:benchmark_seldon}}
{-- Code fully compliant with the C++ standard \\
-- OS Portable \\
-- GNU GCC $\geq$ 3 (from 3.2 to 4.3 tested); Intel C++ compiler icc
(icc 7.1 and 8.0 tested); compile with Microsoft Visual}
{Requires Blas and CBlas}
{-- Thanks to templates, the solvers can be used for any type of matrix and
preconditioning, not only Seldon matrices: very useful to perform a
matrix-vector product when the matrix is not stored \\
-- Provides a Python interface generated by SWIG \\
-- Exception handling and several debug levels helpful while coding \\
-- Good coverage of the interface to Blas and Lapack routines
(see section~\ref{sec:flens_seldon_trilinos_comparisons}) \\
-- A few alternative functions provided in C++ if Blas is not available \\
}
{-- No band matrices}
\newpage
\part{SparseLib++}
SparseLib++ is a C++ class library for efficient sparse matrix computations
across various computational platforms.
\begin{newMargin}{0cm}{-2cm}
\libtab {SparseLib++} {v.1.7} {after 1996} {1996}
{http://math.nist.gov/sparselib++/} {minimal maintenance, not in} {3 authors,
0} {1994} {Public domain (}
\end{newMargin}
\libdesc {-- Sparse double vectors \\
-- Sparse double matrices with several storage formats: compressed
row/column, compressed diagonal, coordinate formats, jagged diagonal, block
compressed row and skyline}
{-- Preconditioners: incomplete LU, incomplete Cholesky and diagonal scaling
\\
-- Iterative solvers: SOR, Chebyshev iteration, BiCg, BiBgStab, Cg, Cgne,
Cgne, Cgnr, Gmres, MinRes and Qmr \\
-- Sparse triangular system solver}
{Blas}
{Not evaluated}
{Various computational platforms}
{Requires Blas}
{-- Built upon sparse Blas (level 3) \\
-- SparseLib++ matrices can be built out of nearly any C++ matrix/vector
classes (it is shipped with the MV++ classes by default)}
{-- Templates not supported (only double elements) \\
-- No new feature since 1996, only maintenance \\
-- No eigenvalues computation \\
-- No dense vectors and matrices}
\newpage
\part{Template Numerical Toolkit (TNT)}
TNT is a collection of interfaces and reference implementations of numerical
objects useful for scientific computing in C++. JAMA/C++ library (Java Matrix
Package translated into C++) utilizes TNT for the lower-level operations to
perform linear algebra operations.
\begin{newMargin}{-0.4cm}{-2cm}
\libtab {TNT} {v.3.0.11} {Jan. 2008} {2003} {http://math.nist.gov/tnt/}
{active maintenance, not in} {2 authors, 0} {Sep. 1999} {Public domain (}
\end{newMargin}
\libdesc {-- Sparse matrices \\
-- 1D, 2D and 3D arrays (row-major and column-major)}
{Provided by JAMA/C++: \\
-- SVD decomposition \\
-- SVD, LU, QR and Cholesky solvers \\
-- Eigenvalues computation}
{None}
{Not evaluated}
{ANSI C++ compatibility: should work with most updated C++ compilers (tested
by the authors with Microsoft Visual C++ v.5.0)}
{By including header files}
{Templates supported}
{-- No sparse vectors \\
-- No separation between headers and sources (only header files) \\
-- Beta release since Jan. 2008}
\newpage
\part{Trilinos}
Trilinos provides algorithms and enabling technologies within an
object-oriented software framework for large-scale, complex multi-physics
engineering and scientific problems. Trilinos is a collection of interacting
independent packages (package names are in italic). \\
\libtab {Trilinos} {9.0.1} {Oct. 2008} {Oct. 2008}
{http://trilinos.sandia.gov/} {in} {34} {Sept. 2003} {LGPL (}
\libdesc {-- Core kernel package (\textit{Kokkos}) \\
-- Dense, symmetric dense, sparse, block sparse, jagged-diagonal sparse
matrices (\textit{Epetra} and \textit{EpetraExt}) \\
-- Dense vectors and multivectors (\textit{Epetra} and \textit{EpetraExt}),
sparse vectors
(\textit{Tpetra}) \\
-- Integer and double elements (\textit{Epetra} and \textit{EpetraExt}),
templates (\textit{Tpetra})}
{-- Preconditioners: ILU-type (\textit{AztecOO} and \textit{IFPACK}),
Multilevel (\textit{ML}), Block (\textit{Meros}) \\\\
-- Linear solvers:\\
Direct dense solvers (\textit{Epetra} and \textit{Teuchos}: wrappers for
selected Blas and Lapack routines, \textit{Pliris}: LU solver on parallel
platforms);\\
Krylov solvers (\textit{AztecOO}: preconditioned Krylov solver,
\textit{Belos}: iterative solver, \textit{Komplex}: for complex values);\\
Direct sparse solvers (\textit{Amesos}: for DSCPACK, SuperLU, SuperLUDist
and UMFPACK);\\
SVD decomposition (\textit{Epetra}) \\\\
-- Nonlinear solvers: \\
System solvers (\textit{NOX}: globalized Newton methods such as line
search and trust region methods, \textit{LOCA}: computing families of
solutions and their bifurcations for large-scale applications); \\
Optimization (\textit{MOOCHO}: reduced-space successive quadratic
programming (SQP) methods); \\
Time integration/Differential-algebraic equations (\textit{Rythmos}) \\\\
-- Eigensolver: block Krylov-Schur, block Davidson and locally-optimal block
preconditioned conjugate gradient (\textit{Anasazi})}
{-- Uses Blas and Lapack \\
-- Provides interfaces for Metis/ParMetis, SuperLU, Aztec, Mumps, Umfpack
and soon PETSc \\
-- Conjunction possible with SWIG, MPI, Expat (XML parser), METIS and
ParMETIS}
{\textit{Epetra} provides classes to distribute vectors, matrices and other
objects on a parallel (or serial) machine}
{Linux, MAC OS X, Windows (under Cygwin), SGI64, DEC and Solaris}
{Requires Blas and Lapack}
{-- Examples of users: SIERRA (Software Environment for Developing Complex
Multiphysics Applications), SALINAS (structural dynamics finite element
analysis), MPSalsa (high resolution 3D simulations with an equal emphasis
on fluid flow and chemical kinetics modeling), Sundance (finite-element
solutions of partial differential equations), DAKOTA (Design Analysis Kit
for Optimization and Terascale Applications~; coding for uncertainty
quantification, parameter estimation and sensitivity/variance analysis) \\
-- Most of Trilinos functionalities available in a Python script\\
-- A package of basic structures with templated types (\textit{Tpetra},
first release distributed with 9.0.1) \\
-- \textit{EpetraExt} enables objects to be easily exported to MATLAB \\
-- Trilinos is based on established algorithms at Sandia. The effort
required to develop new algorithms and enabling technologies is
substantially reduced because a common base provides an excellent starting
point. Furthermore, many applications are standardizing on the Trilinos
APIs: these applications have access to all Trilinos solver components
without any unnecessary interface modifications.}
{-- Not yet available: templates (except in an isolated
package, \textit{Teuchos}) and sparse vectors (Tpetra is still under heavy
development, release planned in Mar./Apr. 2009)\\
-- Impossible to build Trilinos under Windows without Cygwin (improved
Windows support in a further release) \\
-- Trilinos is a complex collection of interoperable packages and requires
some careful configuration (with a suitable set of packages and options)}
\newpage
\part{uBlas}
uBlas is a C++ template class library that provides Blas level 1, 2, 3
functionality for dense, packed and sparse matrices.
\libtab {uBlas} {1.33.0} {Jul. 2008} {2008}
{http://www.boost.org/doc/libs/1_35_0/libs/numeric/ublas/} {in} {5}
{2008} {Boost Software License (}
\libdesc {-- Dense, unit and sparse (mapped, compressed or coordinate) vectors
\\
-- Dense, identity, triangular, banded, symmetric, hermitian, packed and
sparse (mapped, compressed or coordinate) matrices}
{-- Submatrices and subvectors operations \\
-- Triangular solver \\
-- LU factorization}
{Blas (level 1, 2 and 3)}
{Optimized for large vectors and matrices, see benchmarks:
\url{http://flens.sourceforge.net/session2/tut4.html},\\
\url{http://flens.sourceforge.net/session1/tut9.html},
\url{http://eigen.tuxfamily.org/index.php?title=Benchmark},
\url{http://projects.opencascade.org/btl/} and
and section \ref{sec:benchmark_seldon}}
{OS Independent, requires a modern (ISO standard compliant) compiler such as
GCC 3.2.3, 3.3.x, 3.4.x, 4.0.x; MSVC 7.1, 8.0; ICC 8.0, 8.1; Visual
age 6; Codewarrior 9.4, 9.5}
{Requires Blas}
{-- Templates supported \\
-- Included in Boost C++ libraries \\
-- Mathematical notation to ease the development (use of operator
overloading)}
{-- No eigenvalues computation \\
-- Only basic linear algebra (no linear solving except triangular solver) \\
-- The implementation assumes a linear memory address model \\
-- Tuning focussed on dense matrices \\
-- No separation between headers and sources (only header files)}
\newpage
\part{Other Libraries}
\paragraph{Armadillo++}
is a C++ linear algebra library providing matrices and vectors, interfacing
Lapack and Atlas. : \url{http://arma.sourceforge.net/}. \limitations {No
templates (only double), early development (first release in Apr. 2008), no
portability under Windows without Cygwin.}
\paragraph{Blitz++}
is a C++ class library for scientific computing providing high performance by
using template techniques. The current versions provide dense arrays and
vectors, random number generators, and small vectors and matrices:
\url{http://www.oonumerics.org/blitz/}. \limitations {No sparse matrices and
vectors. Barely relevant for linear algebra.}
\paragraph{CPPScaLapack}
is a C++ class wrapper for BLACS, PBlas and ScaLapack with MPI. CPPScaLapack
provides a user-friendly interface of high-speed parallel matrix calculation
with Blas and Lapack techonologies for programers concerning with large-scale
computing: \url{http://cppscalapack.sourceforge.net/}. \limitations {Still an
alpha program, no sparse matrices and vectors, no templates (only
double-precision vectors and general matrices).}
\paragraph{CVM Class Library}
provides vector and different matrices including square, band, symmetric and
hermitian ones in Euclidean space of real and complex numbers. It utilizes
Blas and Lapack. Contains different algorithms including solving of linear
systems, singular value decomposition, non-symmetric and symmetric eigenvalue
problem (including Cholesky and Bunch-Kaufman factorization), LU
factorization, QR factorization: \url{http://www.cvmlib.com/}. \limitations
{Templates supported but distinction between real and complex types. No sparse
matrices and vectors.}
\paragraph{IML++}
(Iterative Methods Library) is a C++ templated library for solving linear
systems of equations, capable of dealing with dense, sparse, and distributed
matrices: \url{http://math.nist.gov/iml++/}. \limitations {No matrices and
vectors, only iterative solvers.}
\paragraph{LA library}
is a C++ interface to Blas and Lapack providing also a general (sparse) matrix
diagonalization, linear solver and exponentiation templates :
\url{http://www.pittnerovi.com/la/}. \limitations {Portability: tested only on
Linux with a code not fully ANSI C++ compliant.}
\paragraph{LinAl}
The library is based on STL techniques and uses STL containers for the storage
of matrix data furthermore STL algorithms are used where feasible. Low level,
algebraic operators as well as linear solvers and eigenvalue solvers are
implemented, based on calls to Blas, Lapack and CGSOLX and LANCZOS:
\url{http://linal.sourceforge.net/LinAl/Doc/linal.html}. \limitations {No
vectors, no sparse matrices.}
\paragraph{LinBox}
is a C++ template library for exact, high-performance linear algebra
computation with dense, sparse, and structured matrices over the integers and
over finite fields: \url{http://www.linalg.org/}. \limitations {Does not suit
to real and complex values.}
\paragraph{MV++}
is a small set of concrete vector and matrix classes specifically designed for
high performance numerical computing, including interfaces to the
computational kernels found in Blas:
\url{http://math.nist.gov/mv++/}. \limitations{Only building blocks to a
larger-user level library such as SparseLib++ and Lapack++.}
\paragraph{Newmat}
C++ library supports various matrix types. Only one element type (float or
double) is supported. The library includes Cholesky decomposition, QR,
triangularisation, singular value decomposition and eigenvalues of a symmetric
matrix: \url{http://www.robertnz.net/nm_intro.htm}. \limitations {No sparse
matrices and vectors, templates not supported.}
\paragraph{RNM}
by \selectlanguage{francais}Frédéric Hecht \selectlanguage{english} provides
C++ classes for arrays with vectorial operations:
\url{http://www.ann.jussieu.fr/~hecht/}. \limitations {Only general matrices,
no sparse matrices, no vectors, only one linear system solver (conjugate
gradient), no English documentation.}
\paragraph{SL++ (Scientific Library)}
is a C++ numerical library, composed of modules specialized in various fields
of numerical computations:
\url{http://ldeniau.home.cern.ch/ldeniau/html/sl++.html}. \limitations {Not
developed since 1998.}
\paragraph{TCBI (Temporary Base Class Idiom)} templated C++ numerical library
implements basic data structures like complex numbers, dynamic vectors, static
vectors, different types of matrices like full matrices, band matrices, sparse
matrices, etc. It also includes some standard matrix solvers like
Gauss-Jordan, LU-decomposition and Singular Value Decomposition and a set of
powerful iterative solvers (Krylov subspace methods along with
preconditioners). Also interfaces to netlib libraries such as CLapack or
SuperLU. Its specificity is being exclusively written in C++, without needing
to interface to Fortran code. The usual loss of performance associated with
object-oriented languages has been avoided through not as obvious
implementations of numerical base classes, avoiding unnecessary copying of
objects. It can be thought of as some sort of reference counting done by the
compiler at compile time. Supported on Linux/Unix with the GNU compiler, on
Windows with the Microsoft Visual C++ (6, 7) compiler and with the Intel
compiler: \url{http://plasimo.phys.tue.nl/TBCI/}. \limitations {No interface
to Blas and Lapack.}
\newpage
\part{Links and Benchmarks}
\section{Links}
Here are some Web portals related to numerical analysis software and linear
algebra libraries:
-- List of numerical analysis software (Wikipedia)
\url{http://en.wikipedia.org/wiki/List_of_numerical_analysis_software}
-- Numerical computing resources on the Internet (Indiana University)
\url{http://www.indiana.edu/~statmath/bysubject/numerics.html}
-- Scientific computing in object-oriented languages (community resources)
\url{http://www.oonumerics.org/oon/}
-- Scientific computing software (Master's Degrees in Applied Mathematics at
\'Ecole Centrale Paris)
\url{http://norma.mas.ecp.fr/wikimas/ScientificComputingSoftware} \\
\section{Benchmarks}
\subsection{Benchmarks for Linear Algebra Libraries}
~
-- Freely available software for linear algebra on the web -- 2006 comparative
statement \url{http://www.netlib.org/utk/people/JackDongarra/la-sw.html}
-- Benchmark sparse matrices (tests for residual and random order
initialization) -- 2008 \url{http://flens.sourceforge.net/session2/tut4.html}
-- Some Blas Benchmarks -- 2007
\url{http://flens.sourceforge.net/session1/tut9.html}
-- Benchmark 2008 \url{http://eigen.tuxfamily.org/index.php?title=Benchmark}
-- Benchmark of C++ Libraries for Sparse Matrix Computation -- 2007
\url{http://grh.mur.at/misc/sparselib_benchmark/report.html}
-- Benchmark for Templated Libraries project -- 2003
\url{http://projects.opencascade.org/btl/} \\
\newpage
\subsection{Benchmarks including Seldon}
\label{sec:benchmark_seldon}
Platform: Intel Core 2 Duo CPU P9500, 2.53GHz, 6 MB cache, 4 GB
Ram.\\
Compiler: gcc version 4.3.2 (Ubuntu 4.3.2-1ubuntu12).\\
Date: March 2009.\\
\subsubsection {Benchmarks for Dense Matrix}
\label{sec:benchmark_dense}
Adapted from \url{http://flens.sourceforge.net/session1/tut9.html}
\begin{figure}[htpb]
\centering
\includegraphics[width=\textwidth]{aat}
\caption{A x At product for dense matrices.}
\label{fig:dense_aat_product}
\end{figure}
\begin{figure}[htpb]
\centering
\includegraphics[width=\textwidth]{ata}
\caption{At x A product for dense matrices.}
\label{fig:dense_ata_product}
\end{figure}
\begin{figure}[htpb]
\centering
\includegraphics[width=\textwidth]{axpy}
\caption{Y += alpha * X for dense vectors.}
\label{fig:dense_Y+=alpha*X}
\end{figure}
\begin{figure}[htpb]
\centering
\includegraphics[width=\textwidth]{matrix_vector}
\caption{Matrix vector product for dense matrices and vectors.}
\label{fig:dense_matrix_vector_product}
\end{figure}
\begin{figure}[htpb]
\centering
\includegraphics[width=\textwidth]{matrix_matrix}
\caption{Matrix matrix product for dense matrices.}
\label{fig:dense_matrix_matrix_product}
\end{figure}
\newpage
\subsubsection{Benchmarks for Sparse Matrix}
Adapted from \url{http://flens.sourceforge.net/session2/tut4.html}. \\
Compiled with -DNDEBUG -O3 options for g++.\\
Matrices of 1 000 000 lines.
\begin{enumerate}
\item Benchmark for sparse matrix, vector product with random initialization\\
5 non zero elements per line on average. In CRS Flens format, the number of
non zero values stored corresponds to the number of non zero values plus the
number of empty lines (one 0 is inserted on the diagonal).\\
\begin{tabular}{|l|c|c|}
\hline
& Flens & Seldon\\
\hline
initialization & 0.35s & \multirow{2}{*}{1.48s}\\
finalization & 1.23s &\\
\hline
y = Ax & 1.28s & 1.3s\\
y = A'x & 1.61s & 1.49s\\
\hline
\end{tabular}
\item{Benchmark with initialization in order and tridiagonal matrix}~\\
3 non zero elements per line (except on first and last lines with only 2 non
zero elements).\\
\begin{tabular}{|l|c|c|c|}
\hline
& Flens & Seldon & Seldon \\
& & (matrix built by hand) & (matrix built with a generic algorithm *)\\
\hline
initialization & 0.28s & \multirow{2}{*}{0.08s} & \multirow{2}{*}{0.1s} \\
finalization & 0.16s & & \\
\hline
y = Ax & 0.19s & 0.18s & 0.18s \\
y = A'x & 0.22s & 0.22s & 0.22s\\
\hline
\end{tabular}
*using MergeSort on already sorted vectors, instead of QuickSort used in
random initialization case. \\
\item{Benchmark for computation of the residual}~\\
5 non zero elements per line on average.\\
\begin{tabular}{|l|c|c|}
\hline
& Flens & Seldon\\
\hline
initialization & 1.34s & \multirow{2}{*}{2.46s}\\
finalization & 1.25s &\\
\hline
computations r = b - Ax & 1.39s & 1.35s\\
\hline
\end{tabular}
\end{enumerate}
\newpage
\section*{Acknowledgement}
This document benefits from discussions with Vivien Mallet, Dominique Chapelle
and Philippe Moireau, and also from corrections thanks to Xavier Clerc.
\part{Appendix}
These sections deal with Flens distribution RC1 (Jul. 2007), Seldon
distribution 2008-11-12 and Trilinos distribution 9.0.1 (Oct. 2008). The first
two libraries satisfy the main requirements exposed in
section~\ref{part:requirements}, and Trilinos was supposed to be a good
reference as for the coverage of Blas and Lapack.
\section{Flens Overloaded Operator Performance Compared to Seldon}
\label{sec:flens_overloaded}
Flens implements a user-friendly interface to Blas routines by overloading
operators. For example, one can add vectors by using mathematical symbols:
\code{y = a + b + c}. Here are a few tests to check if this does not imply any
loss of performance in the computation.
In table \ref{table:flens_overloaded_operator_performance} are presented CPU
times measured with \code{std::clock()} for several operations, for
RowMajor (R.) and ColMajor (C.) matrices. In the following, lower-case letters
denote vectors and upper-case letters denote matrices. \code{Y = A + B} cannot be
used for sparse matrices in Flens, so this has been tested only on general
dense matrices.
In ColumnMajor format for Flens, the affectation operation \code{Y = A} costs much
more (1.48s) than in RowMajor format (0.35s). This lack of performance slows
down sum operations such as \code{Y = A + B}. This is due to a direct call to Blas to
copy the data (a memory block copy in C++ is used in Seldon).
In ColumnMajor for Seldon, matrix vector product such as \code{Y += transpose(A)*B}
is three times longer than in Flens. This is due to a call to a generic
function instead of the Blas routine. Taking advantage of a local knowledge
of Seldon (one of its authors, Vivien Mallet, is part of my team!), we have
fixed this bug and got the same results as in Flens for \code{y~+=
transpose(A)*b} and \code{y +=2*transpose(A)*b + 1.5*c + d}.
\begin{landscape}
\begin{table}
\caption{Flens and Seldon performance benchmarks using overloaded
operators.}
\label{table:flens_overloaded_operator_performance}
\centering
\begin{tabular}{|l||c|c|c||c|c|c|}
\hline
Operation & Flens code & R.(s) & C.(s) & Seldon code & R.(s) & C.(s) \\
\hline
y = a + b + c & y = a + b + c; & 1.16 & 1.17 & y = a; Add(b,y); Add(c,y); &
1.13 & 1.12 \\
& y = a; y += b; y += c; & 1.16 & 1.18 & & & \\
& y = a + b; y += c; & 1.17 & 1.18 & & & \\
\hline
y = a + b + c + d & y=a+b+c+d; & 1.58 & 1.59 & y=a; Add(b,y); Add(c,y);
Add(d,y) & 1.52 & 1.54\\
& y = a; y += b; y += c; y += d; & 1.6 & 1.58 & & & \\
\hline
y += 4a & y += 4a; & 4.11 & 4.07 & Add(4.,a,y); & 4.09 & 4.09\\
& axpy(4, a, y); & 4.09 & 4.1 & & &\\
\hline
Y = A & Y = A; & 0.35 & 1.48 & Y = A; & 0.32 & 0.32 \\
\hline
Y = A + B & Y = A + B; & 0.78 & 2.92 & Y = A; Add(B,Y); & 0.76 & 1.89\\
& Y = A; Y += B; & 0.78 & 2.93 & & & \\
\hline
Y = A + B + C & Y = A + B + C; & 1.21 & 4.41 & Y = A; Add(B,Y);
Add(C,Y); & 1.24 & 3.44\\
& Y = A + B; Y += C; & 1.22 & 4.35 & & & \\
& Y = A; Y += B; Y += C; & 1.21 & 4.35 & & & \\
\hline
Y = A + B + C + D & Y = A + B + C + D; & 1.63 & 5.78 & Y = A; Add(B,Y); Add(C,Y);
Add(D,Y); & 1.7 & 4.95\\
& Y = A; Y+=B; Y+=C; Y+=D; & 1.62 & 5.78 & & & \\
\hline
y = Ax + b & y = Ax + b; & 1.64 & 2.09 & Copy(b,y); MltAdd(1.,A,x,1.,y); & 1.61 & 2.08\\
\hline
y += A*b & y += A*b; & 1.63 & 2.07 & MltAdd(1.,A,b,1.,y); & 1.63 & 2.08\\
\hline
y += transpose(A)*b & y += transpose(A)*b; & 2.05 & 1.63 &
MltAdd(1.,SeldonTrans,A,b,1.,y); & 6.64/2.05* & 1.53 \\
\hline
y+=2transpose(A)*b+1.5c+d & y += 2*transpose(A)*b+1.5c+d; & 2.06 &
1.64 & MltAdd(2.,SeldonTrans,A,b,1.,y);Add(1.5,c,y);Add(d,y); & 6.66/2.05* &
1.52 \\
\hline
C += 1.5A*transpose(B) & C += 1.5A*transpose(B); & 43.18 & 43.3 &
MltAdd(1.5,SeldonNoTrans, A,SeldonTrans,B,1.,C); & 43.1 & 43.26\\
\hline
\end{tabular}
\end{table}
* with the correction (see above for the explanation).
\end{landscape}
\section{Flens, Seldon and Trilinos Content Comparisons}
\label{sec:flens_seldon_trilinos_comparisons}
\subsection{Available Matrix Types from Blas (Flens and Seldon)}
~
F stands for Flens, S stands for Seldon. A black cell for an existing
structure; a gray cell for no structure.
\\
\begin{tabular}{|lcc|lcc|lcc|}
\hline
& F & S & & F & S & & F & S \\
\hline
GE - General & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & GB - General Band & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.95} & & & \\
SY - SYmmetric & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & SB - Sym. Band & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.95} & SP -
Sym. Packed & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} \\
HE - HErmitian & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & HB - Herm. Band & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & HP -
Herm. Packed & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
TR - TRiangular & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & TB - Triang. Band & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.95} & TP -
Triang. Packed & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} \\
\hline
\end{tabular}
\newpage
\subsection{Available Interfaces to Blas and Lapack Routines (Flens and Seldon)}
~
A black cell for an existing interface to a given routine; a gray cell for no
interface.
\begin{newMargin}{-2cm} {-1cm}
\begin{enumerate}
\item{Blas routines}
\begin{tabular}{|lcc|lcc|lcc|lcc|}
\hline
& Flens & Seldon & & Flens & Seldon & & Flens & Seldon & & Flens & Seldon
\\
\hline
srotg & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & drotg & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & crotg & \cellcolor[gray]{0.95} &
\cellcolor[gray]{0.95} & zrotg & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
srotmg & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & drotmg & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & srot & \cellcolor[gray]{0.95} &
\cellcolor[gray]{0.1} & drot & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
csrot & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zdrot & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & srotm & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & drotm & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
sswap & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & dswap & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & cswap & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zswap & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
sscal & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & dscal & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & cscal & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zscal & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} \\
csscal & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zdscal & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & scopy & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & dcopy & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} \\
ccopy & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zcopy & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & saxpy & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & daxpy & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} \\
caxpy & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zaxpy & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & sdot & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & ddot & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} \\
sdsdot & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & dsdot & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & cdotu & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zdotu & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
cdotc & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zdotc & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & snrm2 & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & dnrm2 & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} \\
scnrm2 & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & dznrm2 & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & sasum & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & dasum & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
scasum & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & dzasum & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & isamax & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & idamax & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} \\
icamax & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & izamax & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & sgemv & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & dgemv & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} \\
cgemv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zgemv & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & sgbmv & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.95} & dgbmv & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.95} \\
cgbmv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zgbmv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & chemv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zhemv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
chbmv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zhbmv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & chpmv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zhpmv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
ssymv & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & dsymv & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & ssbmv & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.95} & dsbmv & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.95} \\
sspmv & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & dspmv & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & strmv & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & dtrmv & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} \\
ctrmv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & ztrmv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & stbmv & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.95} & dtbmv & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.95} \\
ctbmv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ztbmv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & stpmv & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & dtpmv & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} \\
ctpmv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & ztpmv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & strsv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & dtrsv & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} \\
ctrsv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & ztrsv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & stbsv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dtbsv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
ctbsv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ztbsv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & stpsv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & dtpsv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
ctpsv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & ztpsv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & sger & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & dger & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
cgeru & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zgeru & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & cgerc & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zgerc & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
cher & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zher & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & chpr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zhpr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
cher2 & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zher2 & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & chpr2 & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zhpr2 & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
ssyr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & dsyr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & sspr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & dspr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
ssyr2 & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dsyr2 & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sspr2 & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & dspr2 & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
sgemm & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & dgemm & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & cgemm & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zgemm & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} \\
ssymm & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & dsymm & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & csymm & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zsymm & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
chemm & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zhemm & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & ssyrk & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dsyrk & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
csyrk & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zsyrk & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cherk & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zherk & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
ssyr2k & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dsyr2k & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & csyr2k & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zsyr2k & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
cher2k & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zher2k & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & strmm & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & dtrmm & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} \\
ctrmm & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & ztrmm & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & strsm & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & dtrsm & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} \\
ctrsm & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & ztrsm & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & & & & & & \\
\hline
\end{tabular}
\begin{tabular}{|p{4.05cm}|cc|}
\hline
Blas & Flens & Seldon \\
\hline
Total & 44 & 110 \\
Coverage & 30\% & 75\% \\
\hline
\end{tabular}
\newpage
\item{Single precision real Lapack routines}
\begin{tabular}{|lcc|lcc|lcc|lcc|}
\hline
& Flens & Seldon & & Flens & Seldon & & Flens & Seldon & & Flens & Seldon \\
\hline
sgesv & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & sgbsv & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.95} & sgtsv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sposv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
sppsv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & spbsv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sptsv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ssysv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
sspsv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sgels & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.95} & sgelsd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sgglse & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
sggglm & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ssyev & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & ssyevd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sspev & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
sspevd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ssbev & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ssbevd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sstev & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
sstevd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sgees & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sgeev & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & sgesvd & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} \\
sgesdd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ssygv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & ssygvd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sspgv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
sspgvd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ssbgv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ssbgvd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sgegs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
sgges & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sgegv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sggev & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & sggsvd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
sgesvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sgbsvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sgtsvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sposvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
sppsvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & spbsvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sptsvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ssysvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
sspsvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sgelsx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sgelsy & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sgelss & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.95} \\
ssyevx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ssyevr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ssygvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sspevx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
sspgvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ssbevx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ssbgvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sstevx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
sstevr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sgeesx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sggesx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sgeevx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
sggevx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sbdsdc & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sbdsqr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sdisna & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
sgbbrd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sgbcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sgbequ & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sgbrfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
sgbtrf & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.95} & sgbtrs & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.95} & sgebak & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sgebal & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
sgebrd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sgecon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & sgeequ & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & sgehrd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
sgelqf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & sgeqlf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sgeqp3 & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sgeqpf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
sgeqrf & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & sgerfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & sgerqf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sgetrf & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} \\
sgetri & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & sgetrs & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & sggbak & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sggbal & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
sgghrd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sggqrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sggrqf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sggsvp & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
sgtcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sgtrfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sgttrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sgttrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
shgeqz & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & shsein & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & shseqr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sopgtr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
sopmtr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sorgbr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sorghr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sorglq & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
sorgql & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sorgqr & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.95} & sorgrq & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sorgtr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
sormbr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sormhr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sormlq & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & sormql & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
sormqr & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & sormr3 & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sormrq & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sormrz & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
sormtr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & spbcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & spbequ & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & spbrfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
spbstf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & spbtrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & spbtrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & spocon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
spoequ & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sporfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & spotrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & spotri & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
spotrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sppcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sppequ & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & spprfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
spptrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & spptri & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & spptrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sptcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
spteqr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sptrfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & spttrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & spttrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
ssbgst & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ssbtrd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sspcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & sspgst & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
ssprfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & ssptrd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ssptrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & ssptri & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
ssptrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & sstebz & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sstedc & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & sstegr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
sstein & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ssteqr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ssterf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ssycon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
ssygst & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ssyrfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & ssytrd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ssytrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
ssytri & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & ssytrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & stbcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & stbrfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
stbtrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & stgevc & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & stgexc & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & stgsen & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
stgsja & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & stgsna & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & stgsyl & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & stpcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
stprfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & stptri & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & stptrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & strcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
strevc & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & strexc & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & strrfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & strsen & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
strsna & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & strsyl & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & strtri & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & strtrs & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} \\
stzrqf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & stzrzf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & & & & & & \\
\hline
\end{tabular}
\begin{tabular}{|p{4.05cm}|cc|}
\hline
Single real Lapack & Flens & Seldon \\
\hline
Total & 15 & 36 \\
Coverage & 8\% & 19\% \\
\hline
\end{tabular}
\newpage
\item{Double precision real Lapack routines}
\begin{tabular}{|lcc|lcc|lcc|lcc|}
\hline
& Flens & Seldon & & Flens & Seldon & & Flens & Seldon & & Flens & Seldon \\
\hline
dgesv & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & dsgesv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dgbsv & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.95} & dgtsv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
dposv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dppsv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dpbsv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dptsv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
dsysv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dspsv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dgels & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.95} & dgelsd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
dgglse & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dggglm & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dsyev & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & dsyevd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
dspev & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & dspevd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dsbev & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dsbevd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
dstev & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dstevd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dgees & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dgeev & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} \\
dgesvd & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & dgesdd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dsygv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & dsygvd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
dspgv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & dspgvd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dsbgv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dsbgvd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
dgegs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dgges & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dgegv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dggev & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
dggsvd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dgesvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dgbsvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dgtsvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
dposvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dppsvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dpbsvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dptsvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
dsysvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dspsvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dgelsx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dgelsy & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
dgelss & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.95} & dsyevx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dsyevr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dsygvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
dspevx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dspgvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dsbevx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dsbgvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
dstevx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dstevr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dgeesx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dggesx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
dgeevx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dggevx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dbdsdc & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dbdsqr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
ddisna & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dgbbrd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dgbcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dgbequ & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
dgbrfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dgbtrf & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.95} & dgbtrs & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.95} & dgebak & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
dgebal & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dgebrd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dgecon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & dgeequ & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
dgehrd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dgelqf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & dgeqlf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dgeqp3 & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
dgeqpf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dgeqrf & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & dgerfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & dgerqf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
dgetrf & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & dgetri & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & dgetrs & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & dggbak & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
dggbal & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dgghrd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dggqrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dggrqf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
dggsvp & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dgtcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dgtrfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dgttrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
dgttrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dhgeqz & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dhsein & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dhseqr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
dopgtr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dopmtr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dorgbr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dorghr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
dorglq & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dorgql & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dorgqr & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & dorgrq & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
dorgtr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dormbr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dormhr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dormlq & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
dormql & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dormqr & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & dormr3 & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dormrq & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
dormrz & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dormtr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dpbcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dpbequ & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
dpbrfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dpbstf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dpbtrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dpbtrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
dpocon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dpoequ & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dporfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dpotrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
dpotri & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dpotrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dppcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dppequ & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
dpprfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dpptrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dpptri & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dpptrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
dptcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dpteqr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dptrfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dpttrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
dpttrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dsbgst & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dsbtrd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dspcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
dspgst & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dsprfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & dsptrd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dsptrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
dsptri & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & dsptrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & dstebz & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dstedc & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
dstegr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dstein & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dsteqr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dsterf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
dsycon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & dsygst & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dsyrfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & dsytrd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
dsytrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & dsytri & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & dsytrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & dtbcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
dtbrfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dtbtrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dtgevc & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dtgexc & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
dtgsen & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dtgsja & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dtgsna & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dtgsyl & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
dtpcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & dtprfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & dtptri & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & dtptrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
dtrcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & dtrevc & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dtrexc & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dtrrfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
dtrsen & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dtrsna & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dtrsyl & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dtrtri & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
dtrtrs & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & dtzrqf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & dtzrzf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & & & \\
\hline
\end{tabular}
\begin{tabular}{|p{4.05cm}|cc|}
\hline
Double real Lapack & Flens & Seldon \\
\hline
Total & 15 & 38 \\
Coverage & 8\% & 20\% \\
\hline
\end{tabular}
\newpage
\item{Single precision complex Lapack routines}
\begin{tabular}{|lcc|lcc|lcc|lcc|}
\hline
& Flens & Seldon & & Flens & Seldon & & Flens & Seldon & & Flens & Seldon \\
\hline
cgesv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & cgbsv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cgtsv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cposv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
cppsv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cpbsv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cptsv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & csysv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
chesv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cspsv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & chpsv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cgels & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
cgelsd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cgglse & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cggglm & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cheev & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
cheevd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cheevr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & chpev & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & chpevd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
chbev & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & chbevd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cgees & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cgeev & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} \\
cgesvd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & cgesdd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & chegv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & chegvd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
chpgv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & chpgvd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & chbgv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & chbgvd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
cgegs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cgges & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cgegv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cggev & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
cggsvd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cgesvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cgbsvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cgtsvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
cposvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cppsvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cpbsvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cptsvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
csysvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & chesvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cspsvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & chpsvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
cgelsx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cgelsy & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cgelss & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cheevx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
cheevr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & chegvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & chpevx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & chpgvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
chbevx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & chbgvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cgeesx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cggesx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
cgeevx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cggevx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cbdsdc & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cbdsqr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
cgbbrd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cgbcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cgbequ & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cgbrfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
cgbtrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cgbtrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cgebak & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cgebal & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
cgebrd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cgecon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & cgeequ & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & cgehrd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
cgelqf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cgeqlf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cgeqp3 & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cgeqpf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
cgeqrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cgerfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & cgerqf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cgetrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
cgetri & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & cgetrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & cggbak & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cggbal & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
cgghrd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cggqrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cggrqf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cggsvp & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
cgtcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cgtrfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cgttrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cgttrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
chgeqz & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & chsein & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & chseqr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cupgtr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
cupmtr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cungbr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cunghr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cunglq & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
cungql & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cungqr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cungrq & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cungtr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
cunmbr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cunmhr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cunmlq & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cunmql & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
cunmqr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cunmr3 & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cunmrq & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cunmrz & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
cunmtr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cpbcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cpbequ & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cpbrfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
cpbstf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cpbtrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cpbtrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cpocon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
cpoequ & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cporfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cpotrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cpotri & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
cpotrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cppcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cppequ & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cpprfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
cpptrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cpptri & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cpptrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cptcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
cpteqr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cptrfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cpttrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cpttrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
chbgst & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & chbtrd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cspcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & chpcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
chpgst & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & csprfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & chprfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & chptrd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
csptrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & chptrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & csptri & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & chptri & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
csptrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & chptrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & cstedc & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & cstegr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
cstein & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & csteqr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & csycon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & checon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
chegst & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & csyrfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & cherfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & chetrd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
csytrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & chetrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & csytri & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & chetri & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
csytrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & chetrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & ctbcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ctbrfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
ctbtrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ctgevc & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ctgexc & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ctgsen & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
ctgsja & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ctgsna & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ctgsyl & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ctpcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
ctprfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & ctptri & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & ctptrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & ctrcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
ctrevc & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ctrexc & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ctrrfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & ctrsen & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
ctrsna & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ctrsyl & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ctrtri & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & ctrtrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
ctzrqf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ctzrzf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & & & & & & \\
\hline
\end{tabular}
\begin{tabular}{|p{4.05cm}|cc|}
\hline
Single complex Lapack & Flens & Seldon \\
\hline
Total & 1 & 42 \\
Coverage & 1\% & 22\% \\
\hline
\end{tabular}
\newpage
\item{Double precision complex Lapack routines}
\begin{tabular}{|lcc|lcc|lcc|lcc|}
\hline
& Flens & Seldon & & Flens & Seldon & & Flens & Seldon & & Flens & Seldon \\
\hline
zgesv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zgbsv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zgtsv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zposv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
zppsv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zpbsv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zptsv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zsysv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
zhesv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zspsv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zhpsv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zgels & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.95} \\
zgelsd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zgglse & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zggglm & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zheev & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
zheevd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zhpev & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zhpevd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zhbev & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
zhbevd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zgees & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zgeev & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & zgesvd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
zgesdd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zhegv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zhegvd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zhpgv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
zhpgvd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zhbgv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zhbgvd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zgegs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
zgges & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zgegv & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zggev & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zggsvd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
zgesvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zgbsvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zgtsvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zposvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
zppsvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zpbsvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zptsvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zsysvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
zhesvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zspsvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zhpsvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zgelsx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
zgelsy & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zgelss & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.95} & zheevx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zheevr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
zhegvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zhpevx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zhpgvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zhbevx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
zhbgvx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zgeesx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zggesx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zgeevx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
zggevx & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zbdsdc & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zbdsqr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zgbbrd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
zgbcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zgbequ & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zgbrfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zgbtrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
zgbtrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zgebak & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zgebal & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zgebrd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
zgecon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zgeequ & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zgehrd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zgelqf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
zgeqlf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zgeqp3 & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zgeqpf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zgeqrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
zgerfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zgerqf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zgetrf & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} & zgetri & \cellcolor[gray]{0.1} & \cellcolor[gray]{0.1} \\
zgetrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zggbak & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zggbal & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zgghrd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
zggqrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zggrqf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zggsvp & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zgtcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
zgtrfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zgttrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zgttrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zhgeqz & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
zhsein & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zhseqr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zupgtr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zupmtr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
zungbr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zunghr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zunglq & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zungql & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
zungqr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zungrq & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zungtr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zunmbr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
zunmhr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zunmlq & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zunmql & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zunmqr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
zunmr3 & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zunmrq & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zunmrz & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zunmtr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
zpbcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zpbequ & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zpbrfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zpbstf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
zpbtrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zpbtrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zpocon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zpoequ & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
zporfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zpotrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zpotri & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zpotrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
zppcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zppequ & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zpprfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zpptrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
zpptri & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zpptrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zptcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zpteqr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
zptrfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zpttrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zpttrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zhbgst & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
zhbtrd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zspcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zhpcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zhpgst & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
zsprfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zhprfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zhptrd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zsptrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
zhptrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zsptri & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zhptri & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zsptrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
zhptrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zstedc & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zstegr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zstein & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
zsteqr & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zsycon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zhecon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zhegst & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
zsyrfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zherfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zhetrd & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & zsytrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
zhetrf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zsytri & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zhetri & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & zsytrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
zhetrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & ztbcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ztbrfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ztbtrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
ztgevc & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ztgexc & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ztgsen & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ztgsja & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
ztgsna & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ztgsyl & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ztpcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & ztprfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} \\
ztptri & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & ztptrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & ztrcon & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & ztrevc & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
ztrexc & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ztrrfs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & ztrsen & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ztrsna & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
ztrsyl & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & ztrtri & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} & ztrtrs & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.1} &
ztzrqf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} \\
ztzrzf & \cellcolor[gray]{0.95} & \cellcolor[gray]{0.95} & & & & & & & & & \\
\hline
\end{tabular}
\begin{tabular}{|p{4.05cm}|cc|}
\hline
Double complex Lapack & Flens & Seldon \\
\hline
Total & 5 & 49 \\
Coverage & 3\% & 25\% \\
\hline
\end{tabular}
\end{enumerate}
\end{newMargin}
\newpage
\subsection{Available Interfaces to Blas and Lapack Routines (Trilinos)}
~
As a rough guide, here are some results for Trilinos. Several Trilinos
packages offer at least a partial interface to Blas and Lapack routines:
\textit{Epetra}, \textit{Teuchos}, \textit{Amesos}, \textit{AztecOO},
\textit{ML}, \textit{MOOCHO} and \textit{Pliris}. The Trilinos column refers
to all the packages tested together. This result is the maximum coverage and
one should be careful for its interpretation. Indeed, some packages may not
communicate together, with non compatible structures, and therefore could not
be used together. Moreover, some of the interfaces may not be usable directly,
only indirectly through other functions. \\
\begin{newMargin}{-2cm} {-1cm}
\begin{tabular}{|p{4.05cm}|cccccccc|}
\hline
Blas & Trilinos &
\textit{Epetra} & \textit{Teuchos} & \textit{Amesos} & \textit{AztecOO} &
\textit{ML} & \textit{MOOCHO} & \textit{Pliris} \\
\hline
Total & 87 & 28 & 62 & 18 & 21 & 29 & 34 & 28\\
Coverage & 60\% & 19\% & 42\% & 12\% & 14\% & 20\% & 23\% & 19\% \\
\hline
\end{tabular}
\begin{tabular}{|p{4.05cm}|cccccccc|}
\hline
Single real Lapack & Trilinos &
\textit{Epetra} & \textit{Teuchos} & \textit{Amesos} & \textit{AztecOO} &
\textit{ML} & \textit{MOOCHO} & \textit{Pliris} \\
\hline
Total & 57 & 44 & 44 & 4 & 8 & 32 & 1 & 1\\
Coverage & 31\% & 24\% & 24\% & 2\% & 4\% & 17\% & 1\% & 1\% \\
\hline
\end{tabular}
\begin{tabular}{|p{4.05cm}|cccccccc|}
\hline
Double real Lapack & Trilinos &
\textit{Epetra} & \textit{Teuchos} & \textit{Amesos} & \textit{AztecOO} &
\textit{ML} & \textit{MOOCHO} & \textit{Pliris} \\
\hline
Total & 59 & 43 & 44 & 4 & 7 & 32 & 7 & 0 \\
Coverage & 32\% & 23\% & 24\% & 2\% & 4\% & 17\% & 4\% & 0\% \\
\hline
\end{tabular}
\begin{tabular}{|p{4.05cm}|cccccccc|}
\hline
Single complex Lapack & Trilinos &
\textit{Epetra} & \textit{Teuchos} & \textit{Amesos} & \textit{AztecOO} &
\textit{ML} & \textit{MOOCHO} & \textit{Pliris} \\
\hline
Total & 39 & 1 & 38 & 1 & 1 & 1 & 0 & 0 \\
Coverage & 20\% & 1\% & 20\% & 1\% & 1\% & 1\% & 0\% & 0\% \\
\hline
\end{tabular}
\begin{tabular}{|p{4.05cm}|cccccccc|}
\hline
Double complex Lapack & Trilinos &
\textit{Epetra} & \textit{Teuchos} & \textit{Amesos} & \textit{AztecOO} &
\textit{ML} & \textit{MOOCHO} & \textit{Pliris} \\
\hline
Total & 38 & 0 & 38 & 1 & 0 & 0 & 0 & 0\\
Coverage & 20\% & 0\% & 20\% & 1\% & 0\% & 0\% & 0\% & 0\% \\
\hline
\end{tabular}
\end{newMargin}
\newpage
\section{Flens and Seldon Synoptic Comparison}
\begin{newMargin}{-1.5cm} {-1cm}
\begin{tabular}{|p{2.5cm}||p{2.5cm}|p{2.5cm}||p{2.5cm}|p{2.5cm}||p{1.5cm}|}
\hline
& \multicolumn{2}{c||}{Flens} & \multicolumn{2}{c||}{Seldon} & See\\
& ~~~~~~~~~--- & ~~~~~~~~~ + & ~~~~~~~~~ --- & ~~~~~~~~~ + & section \\
\hline
Portability & Not recently tested on Windows & & & & \\
\hline
High-level ~ ~ interface & & & & Python interface generated by SWIG & \\
\hline
C++ ~ ~ ~ ~ ~ templates & & Supported & & Supported & \\
\hline
Matrix types & No hermitian matrices & & No band matrices & & \\
\hline
Sparse ~ ~ ~ ~ matrices & & General and symmetric, compressed row storage & &
Harwell-Boeing and array of sparse vectors & \\
\hline
Sparse vectors & Not Supported & & & Supported & \\
\hline
Syntax & & Natural mathematical notation & & & \\
\hline
Maintenance & Release candidate RC1 in Jul. 2007, last commit Jan. 2009 &
& & Latest release in Nov. 2008 & \\
\hline
Blas and Lapack interface & & & & Good coverage &
\ref{sec:flens_seldon_trilinos_comparisons} \\
\hline
Performance & Costly affectation in ColumnMajor format & & One bug (fixed
in later versions) & Better performance for dense matrix vector product &
\ref{sec:benchmark_seldon} and \ref{sec:flens_overloaded}\\
\hline
Vector and matrix views & & Supported & Not supported & & \\
\hline
Technical mastery & & & & One author in my INRIA team! & \\
\hline
Eigenvalues computation & Not Supported & & & Eigenvalues and
eigenvectors computation & \\
\hline
\end{tabular}
\end{newMargin}
\end{document}
|
\section{Introduction}
Thin liquid films are ubiquitous entities in a variety of settings and display interesting dynamics depending on various forces acting on the liquid film and the surface geometries on which the fluid moves. \cite{Oron97}
A number of studies for the stability of a gravity-driven viscous liquid flowing on an inclined flat plane have been done, beginning from the pioneering works of Benjamin \cite{Benjamin57} and Yih. \cite{Yih63} The stability of a wind-driven liquid film flowing over a horizontal flat plate \cite{Craik66} and airfoils \cite{Yih90, Tsao97} was investigated for small disturbances. In these works an interaction between air and liquid flows was considered, and it was shown that the thin liquid film becomes unstable to small disturbances, and that waves arise due to the variable stresses exerted on the liquid-air interface by the airflow.
On the other hand, glaze (wet) ice formation and icicle growth are the problem of ice growth from a liquid film flow accompanying a phase change. \cite{Farzaneh08} Glaze ice forms when water is collected from the impingement of supercooled water droplets, whereas icicles grow from the water of melting snow and ice at the root of the icicles.
Typically, icicles also make up an important part of the total ice load in freezing rain. \cite{Makkonen98}
The glaze ice and icicle surfaces are covered with a supercooled water film, and ice grows from a part of the water film by releasing latent heat into the ambient air below 0 $^{\circ}$C.
It is well known that a solid surface under a supercooled liquid film is morphologically unstable, resulting in dendritic growth and a material is a microscopic mixture of solid and liquid. When a film of water is supplied by impinging droplets and cooled from its surface by cold air, the growing ice always initially entraps a considerable amount of liquid water. This is called spongy ice. \cite{Makkonen87, Farzaneh08}
It was recently shown that sponginess is a material parameter (70 $\%$ of ice), and is independent of the growth conditions. \cite{Makkonen10}
The remaining unfrozen water flows on the ice surface. It should be noted that water flow is significantly different over an accreting ice layer than a non-accreting substrate and hence the problem of ice accretion in the presence of a flowing water film is highly complex.
Ringlike ripples of a centimeter-scale in wavelength are formed. \cite{Knight80, Maeno94, Farzaneh08}
Although the basic mechanism of icicle growth is well known, \cite{Makkonen88, Maeno94} the mechanism of icicle ripple formation remained unsolved.
Ogawa and Furukawa first attempted a theoretical explanation for the icicle ripple formation in the absence of airflow around icicles, where the icicle surface was covered with a gravity driven supercooled water film with a free surface. \cite{Ogawa02} In an improved model proposed by Ueno, \cite{Ueno03, Ueno04, Ueno07} the influence of the shape of the water-air interface due to the action of gravity and surface tension on the ice growth conditions was newly taken into account, and a quite different mechanism on the origin of ripples on icicles was proposed. The theoretical results obtained by Ueno have recently been compared very favorably with experimental results.\cite{UFYT10}
We extended the above theoretical framework to include natural convection airflow around icicles. \cite{Ueno10} It was found that the enhancement of the rate of latent heat loss from the water-air interface to the surrounding air due to airflow caused the amplification rate of the ice-water interface disturbance to increase. However, the wavelength of ice ripples was not significantly affected by the airflow.
Since the natural convection airflow around icicles was less than 1 m/s in velocity, the free shear stress condition at the water-air interface was still satisfied, and hence driving force of the water film over the ice surface was gravity only. \cite{Ueno10} However, when wind speed is large, the wind drag at the water-air interface also drives the water film.
For example, the combination of two driving mechanism due to gravity and wind drag produces a variety of aufeis (also referred to as icings) morphologies with various surface features. \cite{Streitz02}
Aufeis are spreading and thickening ice accretions that form in cold air when a thin sheet of water flows or trickles over a cold surface. According to the aufeis formation experiments,
an initial morphologies of aufeis appeared essentially wavelike (or terraced) on a planar aluminum and a smooth ice surface, and their spacing and height varied with slope of an inclined plane and wind speed. \cite{Streitz02}
The morphological instability of the ice-water interface under the water film flow due to the two driving forces is also relevant to the surface roughness characteristics associated with
glaze icing formation around aircraft wings and structures. \cite{Shin96, Kollar10}
Furthermore, the morphological instability of the surface of growing ice is closely related to various natural phenomena where a thin layer of moving fluid separates the developing solid from the surrounding air. \cite{Meakin10}
The physical model commonly used in ice accretion codes is mainly based on conservation of energy and mass
within numerical cells along the ice accreting surface. \cite{Poots96, Farzaneh08} However, in glaze icing conditions the numerical results are poor agreement with experimental results. Therefore, some investigators developed theoretical and numerical understanding of the dynamic effect of water film flow on the glaze ice accretion.
Bourgault $\etal$ applied a simple film flow model to the problem of aircraft icing. \cite{Bourgault00}
Myers $\etal$ introduced a mathematical model for ice accretion with water flow driven by air shear, gravity, and surface tension, employing the lubrication approximation to describe the water film flow. \cite{Myers02_1, Myers02_2, Myers04}
A version of the Myers $\etal$ model is used in the ICECREMO commercial aircraft icing code.
The aerodynamic forces, as modified by the accreted ice, are significant in determining the wind drag and lift on iced structures. However, most analyses in current glaze ice accretion models lack the physical motivation for the effects of either surface roughness or profile change of ice on the heat transfer coefficient, and the roughness is treated as input to the code. \cite{Gent00}
The lack of roughness formation in standard icing models indicates that the related surface instabilities must originate from more localized structures in which air flow can interact with the water film. \cite{Tsao98, Tsao00}
A more microscopic, rather than global, energy balance and detailed analysis of the interaction between the air and water flows are required to predict fine details of localized roughness. \cite{Tsao98, Tsao00, Gent00} Therefore, herein we perform a linear stability analysis for ice growth under a supercooled water film driven by a laminar airflow, taking into account the effect of interaction between the air and water flows on the ice growth conditions.
There is a significant difference between the current and previous works, \cite{Ueno03, Ueno04, Ueno07, UFYT10, Ueno10} as follows: Our previous works featured a gravity driven water flow, and the shape of the water-air interface was determined by the action of gravity and surface tension only. In the current model, water flow driven by air shear stress is considered. We will show that when the air and water flows are coupled, tangential and normal air shear stress disturbances as well as gravity and surface tension play an important role in determining the shape of the water-air interface as air stream velocity increases. Without employing any of the empirical methods used in standard icing models, the heat transfer coefficient at the water-air interface is determined explicitly by solving the governing equations for the air and water flows and the air temperature field.
It will be shown that the growth conditions of the ice-water interface disturbance as well as the heat transfer coefficient at the water-air interface are strongly affected by the air shear stress disturbances, which is particularly a new effect not found in gravity driven water flow.
\section{\label{sec:model}Model}
\begin{figure}
\begin{center}
\includegraphics[width=12cm,height=12cm,keepaspectratio,clip]{configuration.eps}
\end{center}
\caption{Physical model of air-water-ice multi-phase system. Vertical height is not to scale.}
\label{fig:ice-water-air}
\end{figure}
The model configuration is shown in Fig. \ref{fig:ice-water-air}, which is based on the experiments of aufeis formation on a smooth ice substrate in a wind tunnel by Streitz and Ettema. \cite{Streitz02} In that experiment, water was supplied through a row of holes located at the upstream end of the wind tunnel in a refrigerated laboratory, and the water film was driven by gravity and wind drag for various plane slopes and wind speeds. The following analysis is restricted to a two-dimensional vertical cross-section. The position $x$ is measured from the leading edge where water is supplied at a constant rate, and the $y$ axis is normal to it. The ice is covered with a thin water film of $\bar{h}_{0}$. A cold air stream flows over the thin water layer, and the airflow is assumed to be laminar. The surface of the water layer moves at a velocity of $u_{la}$ under the influence of wind drag. The air velocity approaches the free stream velocity $u_{\infty}$ at a distance $\delta$ from the water-air interface. Simultaneous water and air boundary layers occur. Since the air temperature at $T_{\infty}$ is lower than the ice-water interface temperature, $T_{sl}$, ice grows from a part of the water film by releasing the latent heat to the air through the water-air interface at temperature $T_{la}$.
In this model, water is supplied from only the leading edge and there is no impingement of supercooled water droplets on the water film surface. The water and air boundary layers start at $x=0$. On the other hand, aircraft icing is primarily due to the impact of supercooled water droplets on a cold surface. One could see ice buildup at the leading edge of the airfoil, taken as $x=0$, where unfrozen water flow is slowest, but the water layer is thickest. In this sense, the model herein is not yet truly relevant to the aircraft icing problems. \cite{Bourgault00, Myers02_1, Myers02_2, Myers04, Gent00, Tsao98, Tsao00}
In addition, the following assumptions are used in the current model:
(1) The water film is driven by wind drag only. Hence the analysis is only valid on a horizontal surface, and the free stream velocity $u_{\infty}$ is constant in space.
(2) Density remains constant through the phase change.
(3) Change in ice shape disturbs the water-air interface, and the flow and temperature fields in the water film and air. A quasi-stationary approximation is used for the disturbed fields and unsteadiness only enters through the Stephan condition, due to the long time scale of the ice-water interface motion.
(4) Heat conduction into a substrate beneath ice sheet is not included. The ice sheet is assumed to be thick and the undisturbed part of temperature gradient in the ice does not exist.
(5) The presence of waves on the water film is ignored because the waves did not interact with the forming ice in any observable manner in the experiments, except for enhancing the spreading of the water over the aufeis surface. \cite{Streitz02}
\subsection{Governing equations}
The velocity components in the $x$ and $y$ directions in the air, $u_{a}$ and $v_{a}$, are governed by \cite{Landau59}
\begin{equation}
\frac{\partial u_{a}}{\partial t}
+u_{a}\frac{\partial u_{a}}{\partial x}
+v_{a}\frac{\partial u_{a}}{\partial y}
=-\frac{1}{\rho_{a}}\frac{\partial p_{a}}{\partial x}
+\nu_{a}\left(\frac{\partial^{2}u_{a}}{\partial x^{2}}
+\frac{\partial^{2}u_{a}}{\partial y^{2}}\right),
\label{eq:geq-ua}
\end{equation}
\begin{equation}
\frac{\partial v_{a}}{\partial t}
+u_{a}\frac{\partial v_{a}}{\partial x}
+v_{a}\frac{\partial v_{a}}{\partial y}
=-\frac{1}{\rho_{a}}\frac{\partial p_{a}}{\partial y}
+\nu_{a}\left(\frac{\partial^{2}v_{a}}{\partial x^{2}}
+\frac{\partial^{2}v_{a}}{\partial y^{2}}\right),
\label{eq:geq-va}
\end{equation}
\begin{equation}
\frac{\partial u_{a}}{\partial x}+\frac{\partial v_{a}}{\partial y}=0,
\label{eq:continuity-air}
\end{equation}
where $p_{a}$ is the air pressure, $\rho_{a}=1.3$ ${\rm kg/m^{3}}$, the density of air, and $\nu_{a}=1.3 \times 10^{-5}$ ${\rm m^{2}/s}$, the kinematic viscosity of air.
The velocity components in the $x$ and $y$ directions in the water layer, $u_{l}$ and $v_{l}$, are governed by \cite{Landau59}
\begin{equation}
\frac{\partial u_{l}}{\partial t}
+u_{l}\frac{\partial u_{l}}{\partial x}
+v_{l}\frac{\partial u_{l}}{\partial y}
=-\frac{1}{\rho_{l}}\frac{\partial p_{l}}{\partial x}
+\nu_{l}\left(\frac{\partial^{2}u_{l}}{\partial x^{2}}
+\frac{\partial^{2}u_{l}}{\partial y^{2}}\right),
\label{eq:geq-ul}
\end{equation}
\begin{equation}
\frac{\partial v_{l}}{\partial t}
+u_{l}\frac{\partial v_{l}}{\partial x}
+v_{l}\frac{\partial v_{l}}{\partial y}
=-\frac{1}{\rho_{l}}\frac{\partial p_{l}}{\partial y}
+\nu_{l}\left(\frac{\partial^{2}v_{l}}{\partial x^{2}}
+\frac{\partial^{2}v_{l}}{\partial y^{2}}\right)
-g,
\label{eq:geq-vl}
\end{equation}
\begin{equation}
\frac{\partial u_{l}}{\partial x}+\frac{\partial v_{l}}{\partial y}=0,
\label{eq:continuity-water}
\end{equation}
where $\nu_{l}=1.8 \times 10^{-6}$ ${\rm m^{2}/s}$ and $\rho_{l}=1.0 \times 10^{3}$ ${\rm kg/m^{3}}$ are the kinematic viscosity and density of water, respectively, $p_{l}$ is the water pressure and $g$ the gravitational acceleration.
The continuity equations (\ref{eq:continuity-air}) and (\ref{eq:continuity-water}) can be satisfied by introducing the stream functions $\psi_{a}$ and $\psi_{l}$ such that
$u_{a}=\partial \psi_{a}/\partial y$,
$v_{a}=-\partial \psi_{a}/\partial x$,
$u_{l}=\partial \psi_{l}/\partial y$, and
$v_{l}=-\partial \psi_{l}/\partial x$.
Neglecting viscous dissipation in the energy equation, the equations for the temperatures in the air $T_{a}$, water $T_{l}$ and ice $T_{s}$ are \cite{Landau59}
\begin{equation}
\frac{\partial T_{a}}{\partial t}
+u_{a}\frac{\partial T_{a}}{\partial x}
+v_{a}\frac{\partial T_{a}}{\partial y}
=\kappa_{a}\left(\frac{\partial^{2} T_{a}}{\partial x^{2}}
+\frac{\partial^{2} T_{a}}{\partial y^{2}}\right),
\label{eq:geq-Ta}
\end{equation}
\begin{equation}
\frac{\partial T_{l}}{\partial t}
+u_{l}\frac{\partial T_{l}}{\partial x}
+v_{l}\frac{\partial T_{l}}{\partial y}
=\kappa_{l}\left(\frac{\partial^{2} T_{l}}{\partial x^{2}}+\frac{\partial^{2} T_{l}}{\partial y^{2}}\right),
\label{eq:geq-Tl}
\end{equation}
\begin{equation}
\frac{\partial T_{s}}{\partial t}
=\kappa_{s}\left(\frac{\partial^{2} T_{s}}{\partial x^{2}}
+\frac{\partial^{2} T_{s}}{\partial y^{2}}\right),
\label{eq:geq-Ts}
\end{equation}
where $\kappa_{a}=1.87 \times 10^{-5}$ ${\rm m^{2}/s}$, $\kappa_{l}=1.33 \times 10^{-7}$ ${\rm m^{2}/s}$ and $\kappa_{s}=1.15 \times 10^{-6}$ ${\rm m^{2}/s}$ are the thermal diffusivities of air, water and ice, respectively.
\subsection{\label{sec:BC}Boundary conditions}
The following boundary conditions are the same as those used in a previous paper \cite{Ueno10} except for the first condition in Eq. (\ref{eq:bc-infinity}) herein.
Ignoring the density difference between ice and water, there is no normal fluid motion at the ice-water interface. Then
both velocity components $u_{l}$ and $v_{l}$ at a disturbed ice-water interface, $y=\zeta(t,x)$, must satisfy the no-slip condition:\cite{Myers02_1, Myers02_2, Myers04}
\begin{equation}
u_{l}|_{y=\zeta}=0,
\hspace{1cm}
v_{l}|_{y=\zeta}=0.
\label{eq:bc-ul-vl-zeta}
\end{equation}
Since there is no impingement of supercooled water droplets on the water film,
the kinematic condition at a disturbed water-air interface, $y=\xi(t,x)$, is \cite{Benjamin57,Oron97}
\begin{equation}
\frac{\partial \xi}{\partial t}+u_{l}|_{y=\xi}\frac{\partial \xi}{\partial x}=v_{l}|_{y=\xi}.
\label{eq:bc-kinematic-xi}
\end{equation}
The continuity of velocities of water film flow and airflow at the water-air interface is \cite{Yih90}
\begin{equation}
u_{l}|_{y=\xi}=u_{a}|_{y=\xi},
\qquad
v_{l}|_{y=\xi}=v_{a}|_{y=\xi}.
\label{eq:bc-ul-vl-ua-va-xi}
\end{equation}
The continuity of tangential and normal stresses at the water-air interface leads to \cite{Landau59, Craik66, Yih90}
\begin{equation}
\mu_{l}\left(\frac{\partial u_{l}}{\partial y}\Big|_{y=\xi}
+\frac{\partial v_{l}}{\partial x}\Big|_{y=\xi}\right)
=\mu_{a}\left(\frac{\partial u_{a}}{\partial y}\Big|_{y=\xi}
+\frac{\partial v_{a}}{\partial x}\Big|_{y=\xi}\right),
\label{eq:bc-shear-stress-xi}
\end{equation}
\begin{equation}
-p_{a}|_{y=\xi}+2\mu_{a}\frac{\partial v_{a}}{\partial y}\Big|_{y=\xi}
-\left(-p_{l}|_{y=\xi}+2\mu_{l}\frac{\partial v_{l}}{\partial y}\Big|_{y=\xi}\right)
=-\gamma\frac{\partial^{2}\xi}{\partial x^{2}}\left[1+\left(\frac{\partial \xi}{\partial x}\right)^{2}\right]^{-3/2},
\label{eq:bc-normal-stress-xi}
\end{equation}
where $\mu_{l}=\rho_{l}\nu_{l}$ and $\mu_{a}=\rho_{a}\nu_{a}$ are the viscosities of water and air, respectively, and $\gamma=7.6 \times 10^{-2}$ N/m is the surface tension.
The curvature term on the right hand side in Eq. (\ref{eq:bc-normal-stress-xi}) determines the magnitude of the surface tension induced stress. Hence, the condition expressed by Eq. (\ref{eq:bc-normal-stress-xi}) is that the capillary force resisting displacement and the normal stress on either side of the water-air interface should be in equilibrium. \cite{Craik66}
The continuity condition of temperature at the ice-water interface is
\begin{equation}
T_{l}|_{y=\zeta}=T_{s}|_{y=\zeta}=T_{i},
\label{eq:Tsl}
\end{equation}
in which the interfacial temperature $T_{i}$ is an unknown to be determined.
The Stephan condition is
\begin{equation}
L\left(\bar{V}+\frac{\partial \zeta}{\partial t} \right)
=K_{s}\frac{\partial T_{s}}{\partial y}\Big|_{y=\zeta}
-K_{l}\frac{\partial T_{l}}{\partial y}\Big|_{y=\zeta},
\label{eq:heatflux-zeta}
\end{equation}
which is based on the assumption that ice grows in proportion to the heat flux difference across the ice-water interface.
Here $L=3.3 \times 10^{8}$ ${\rm J/m^{3}}$ is the latent heat per unit volume, $\bar{V}$ is the undisturbed ice growth rate, and $K_{s}=2.22$ ${\rm J/(m\,K\,s)}$ and $K_{l}=0.56$ ${\rm J/(m\,K\,s)}$ are thermal conductivities of ice and water, respectively.
The continuity condition of temperature at the water-air interface is
\begin{equation}
T_{l}|_{y=\xi}=T_{a}|_{y=\xi}=T_{la},
\label{eq:bc-Tla}
\end{equation}
where $T_{la}$ is the temperature at the water-air interface and will be determined later.
The continuity of heat flux at the water-air interface is
\begin{equation}
-K_{l}\frac{\partial T_{l}}{\partial y}\Big|_{y=\xi}
=-K_{a}\frac{\partial T_{a}}{\partial y}\Big|_{y=\xi},
\label{eq:heatflux-xi}
\end{equation}
where $K_{a}=0.024$ ${\rm J/(m\,K\,s)}$ is the thermal conductivity of air.
Far away from the air boundary layer, the velocities and temperature asymptote to their far-field values:
\begin{equation}
u_{a}|_{y=\infty}=u_{\infty}, \qquad
v_{a}|_{y=\infty}=0, \qquad
T_{a}|_{y=\infty}=T_{\infty}.
\label{eq:bc-infinity}
\end{equation}
\subsection{\label{sec:LSA}Linear stability analysis}
In this paper, the stability analysis will be limited to one-dimensional disturbances. A simple normal-mode analysis is applied to the field variables. Suppose an ice-water interface disturbance with a small amplitude $\zeta_{k}$ resulting in $\zeta(t,x)=\zeta_{k}{\rm exp}[\sigma t+i kx]$,
where $k$ is the wave number and $\sigma=\sigma^{(r)}+i \sigma^{(i)}$, $\sigma^{(r)}$ and $v_{p} \equiv -\sigma^{(i)}/k$ are the amplification rate and phase velocity of the disturbance, respectively.
We separate $\xi$, $\psi_{a}$, $\psi_{l}$, $p_{a}$, $p_{l}$, $T_{a}$, $T_{l}$ and $T_{s}$ into undisturbed steady fields with bar and disturbed parts with prime as follows:
$\xi=\bar{h}_{0}+\xi'$,
$\psi_{a}=\bar{\psi}_{a}+\psi'_{a}$,
$\psi_{l}=\bar{\psi}_{l}+\psi'_{l}$,
$p_{a}=\bar{p}_{a}+p'_{a}$,
$p_{l}=\bar{p}_{l}+p'_{l}$,
$T_{a}=\bar{T}_{a}+T'_{a}$,
$T_{l}=\bar{T}_{l}+T'_{l}$
and
$T_{s}=\bar{T}_{s}+T'_{s}$.
The undisturbed velocities in the air and water are derived from
$\bar{u}_{a}= \partial \bar{\psi}_{a}/\partial y$,
$\bar{v}_{a}=-\partial \bar{\psi}_{a}/\partial x$,
$\bar{u}_{l}= \partial \bar{\psi}_{l}/\partial y$, and
$\bar{v}_{l}=-\partial \bar{\psi}_{l}/\partial x$.
We define
$\bar{G}_{a} \equiv -\partial \bar{T}_{a}/\partial y|_{y=\bar{h}_{0}}$ and
$\bar{G}_{l} \equiv -\partial \bar{T}_{l}/\partial y|_{y=0}$
as temperature gradients at the undisturbed water-air interface and ice-water interface, respectively.
The disturbed field variables are assumed to be expanded in normal mode form, as follows:
\begin{equation}
\left(
\begin{array}{c}
\xi' \\
\psi'_{a} \\
\psi'_{l} \\
p'_{a} \\
p'_{l} \\
T'_{a} \\
T'_{l} \\
T'_{s}
\end{array}
\right)
=
\left(
\begin{array}{c}
\xi_{k} \\
u_{\infty}f_{a}(\eta)\xi_{k} \\
u_{la}f_{l}(y_{*})\zeta_{k} \\
(\rho_{a}u_{\infty}^{2}/\delta_{0})g_{a}(\eta)\xi_{k} \\
(\rho_{l}u_{la}^{2}/\bar{h}_{0})g_{l}(y_{*})\zeta_{k} \\
H_{a}(\eta)\bar{G}_{a}\xi_{k} \\
H_{l}(y_{*})\bar{G}_{l}\zeta_{k} \\
H_{s}(y_{*})\bar{G}_{l}\zeta_{k}
\end{array}
\right)
\exp[\sigma t+i kx].
\label{eq:pertset}
\end{equation}
We introduce the following two dimensionless variables $\eta=(y-\bar{h}_{0})/\delta_{0}$ in the air and $y_{*}=y/\bar{h}_{0}$ in the water layer.
Here
$\delta_{0}=(2\nu_{a}x/u_{\infty})^{1/2}$ is a scaled measure of the air boundary layer thickness, \cite{Schlichting99}
$u_{\infty}$ is the free stream velocity and
$x$ is the distance from the leading edge where water is supplied.
$u_{la}$ in Eq. (\ref{eq:pertset}) is the surface velocity of the water film driven by wind drag.
As shown in Eq. (\ref{eq:h0-ula}) herein, $\bar{h}_{0}$ and $u_{la}$ are functions of $x$.
$\xi_{k}$ is the amplitude of the water-air interface disturbance, and $f_{a}$, $f_{l}$, $g_{a}$, $g_{l}$, $H_{a}$, $H_{l}$ and $H_{s}$ are dimensionless amplitudes of disturbed parts of the stream function $\psi$, pressure $p$ and temperature $T$.
In the following, quasi-stationary approximation is used for the disturbed fields as in previous papers, \cite{Ueno03, Ueno04, Ueno07, UFYT10, Ueno10} and we assume that the undisturbed part of temperature gradient within the ice does not exist, hence $\bar{T}_{s}=T_{sl}$ ($T_{sl}=$0 $^{\circ}$C for pure water).
\subsubsection{Equations and boundary conditions for undisturbed flows and temperatures in the air and water film}
Using the Blasius-type similarity transformations
and substituting
$\bar{\psi}_{a}=u_{\infty}\delta_{0}\bar{F}_{a}(\eta)$ and
$\bar{T}_{a*}=(\bar{T}_{a}-T_{\infty})/(T_{la}-T_{\infty})$
into the partial differential equations (\ref{eq:geq-ua}), (\ref{eq:geq-va}) and (\ref{eq:geq-Ta}),
a set of ordinary differential equations for the dimensionless functions $\bar{F}_{a}$ and $\bar{T}_{a*}$
and their boundary conditions are obtained: \cite{Schlichting99}
\begin{equation}
\frac{d^{3}\bar{F}_{a}}{d\eta^{3}}
=-\bar{F}_{a}\frac{d^{2}\bar{F}_{a}}{d\eta^{2}},
\label{eq:geq-basicFa}
\end{equation}
\begin{equation}
\frac{d^{2}\bar{T}_{a*}}{d\eta^{2}}
=-Pr_{a}\bar{F}_{a}\frac{d\bar{T}_{a*}}{d\eta},
\label{eq:geq-basicTa}
\end{equation}
\begin{equation}
\frac{d\bar{F}_{a}}{d\eta}\Big|_{\eta=0}=0, \qquad
\bar{F}_{a}|_{\eta=0}=0, \qquad
\frac{d\bar{F}_{a}}{d\eta}\Big|_{\eta=\infty}=1, \qquad
\bar{T}_{a*}|_{\eta=0}=1, \qquad
\bar{T}_{a*}|_{\eta=\infty}=0,
\label{eq:bc-basic_air}
\end{equation}
where $Pr_{a}=\nu_{a}/\kappa_{a}$ is the Prandtl number of air.
The first and second equations in Eq. (\ref{eq:bc-basic_air}) are derived from the undisturbed parts in Eq. (\ref{eq:bc-ul-vl-ua-va-xi}) by using the fact that the free stream velocity $u_{\infty}$ is much larger than the water surface velocity $u_{la}$,\cite{Ueno10} as shown in Table \ref{tab:tableI}.
The third equation in Eq. (\ref{eq:bc-basic_air}) is the result of the condition $\bar{u}_{a}|_{y=\infty}=\partial\bar{\psi}_{a}/\partial y|_{y=\infty}=u_{\infty}$.
The boundary conditions $\bar{T}_{a}|_{y=\bar{h}_{0}}=T_{la}$ and $\bar{T}_{a}|_{y=\infty}=T_{\infty}$ yield the fourth and fifth equations in Eq. (\ref{eq:bc-basic_air}).
For water film, we assume the following scaling $\bar{h}_{0}=C_{1}x^{a}$, $u_{la}=C_{2}x^{b}$ and $T_{sl}-T_{la}=C_{3}x^{c}$, and
$\bar{\psi}_{l}=u_{la}\bar{h}_{0}\bar{F}_{l}(y_{*})$.
Here the constants $C_{1}$, $C_{2}$, $C_{3}$, $a$, $b$ and $c$ are determined from the boundary conditions, as follows.
First, if the volumetric water flow rate per width,
\begin{equation}
Q/l_{w}=\int_{0}^{\bar{h}_{0}}\bar{u}_{l}dy=C_{1}C_{2}x^{a+b}\int_{0}^{1}\bar{u}_{l*}dy_{*},
\label{eq:Qoverl}
\end{equation}
is constant, $a+b=0$ must hold, where $\bar{u}_{l*}\equiv\bar{u}_{l}/u_{la}=d\bar{F}_{l}/dy_{*}$.
This is also derived by substituting
$\bar{\psi}_{l}=u_{la}\bar{h}_{0}\bar{F}_{l}(y_{*})$ into the undisturbed part of Eq. (\ref{eq:bc-kinematic-xi}), $\bar{u}_{l}|_{y=\bar{h}_{0}}d\bar{h}_{0}/dx=\bar{v}_{l}|_{y=\bar{h}_{0}}$.
Second, the undisturbed part of Eq. (\ref{eq:bc-shear-stress-xi}) yields
\begin{equation}
\frac{C_{2}}{C_{1}}\mu_{l}\frac{d\bar{u}_{l*}}{dy_{*}}\Big|_{y_{*}=1}x^{b-a}
=\left(\frac{u_{\infty}^{3}}{2\nu_{a}}\right)^{1/2}\mu_{a}\frac{d^{2}\bar{F}_{a}}{d\eta^{2}}\Big|_{\eta=0}x^{-1/2},
\label{eq:C1-C2}
\end{equation}
from which $b-a=-1/2$ must hold.
Finally, the undisturbed part of Eq. (\ref{eq:heatflux-xi}) yields
\begin{equation}
K_{l}\frac{C_{3}}{C_{1}}\frac{d\bar{T}_{l*}}{dy_{*}}\Big|_{y_{*}=1}x^{c-a}
=-K_{a}\left(\frac{u_{\infty}}{2\nu_{a}}\right)^{1/2}T_{\infty}\frac{d\bar{T}_{a*}}{d\eta}\Big|_{\eta=0}x^{-1/2},
\label{eq:C1-C3}
\end{equation}
where $T_{la}-T_{\infty}\approx -T_{\infty}$ is used because we assume $|T_{la}| \ll |T_{\infty}|$.
From Eq. (\ref{eq:C1-C3}) $c-a=-1/2$ must hold.
Substituting
$\bar{\psi}_{l}=u_{la}\bar{h}_{0}\bar{F}_{l}(y_{*})$ and
$\bar{T}_{l*}=(\bar{T}_{l}-T_{sl})/(T_{sl}-T_{la})$
into the partial differential equations (\ref{eq:geq-ul}) and (\ref{eq:geq-Tl}), a set of differential equations for the dimensionless functions $\bar{u}_{l*}$ and $\bar{T}_{l*}$ is obtained:
\begin{equation}
\frac{d^{2}\bar{u}_{l*}}{dy_{*}^{2}}
=2b\frac{\Rey_{l}\bar{h}_{0}}{\Rey_{a}\delta_{0}}\bar{u}_{l*}^{2},
\label{eq:geq-basicFl}
\end{equation}
\begin{equation}
\frac{d^{2}\bar{T}_{l*}}{dy_{*}^{2}}
=2c\frac{\Rey_{l}\bar{h}_{0}}{\Rey_{a}\delta_{0}}Pr_{l}\bar{u}_{l*}\bar{T}_{l*},
\label{eq:geq-basicTl}
\end{equation}
where $\Rey_{l}=u_{la}\bar{h}_{0}/\nu_{l}$ and $\Rey_{a}=u_{\infty}\delta_{0}/\nu_{a}$ are the Reynolds numbers of the water and air, and $Pr_{l}=\nu_{l}/\kappa_{l}$ is the Prandtl number of water.
Since $\Rey_{l}\bar{h}_{0}/(\Rey_{a}\delta_{0}) \ll 1$ for values shown in Table \ref{tab:tableI},
Eqs. (\ref{eq:geq-basicFl}) and (\ref{eq:geq-basicTl}) can be approximated as
$d^{2}\bar{u}_{l*}/dy_{*}^{2}=0$ and $d^{2}\bar{T}_{l*}/dy_{*}^{2}=0$.
The boundary conditions
$\bar{u}_{l}|_{y=0}=0$,
$\mu_{l}\partial\bar{u}_{l}/\partial y|_{y=\bar{h}_{0}}
=\mu_{a}\partial\bar{u}_{a}/\partial y|_{y=\bar{h}_{0}}$,
$\bar{T}_{l}|_{y=0}=T_{sl}$ and
$\bar{T}_{l}|_{y=\bar{h}_{0}}=T_{la}$
can be written as
$\bar{u}_{l*}|_{y_{*}=0}=0$,
$d\bar{u}_{l*}/dy_{*}|_{y_{*}=1}=1$,
$\bar{T}_{l*}|_{y_{*}=0}=0$
and
$\bar{T}_{l*}|_{y_{*}=1}=-1$, respectively, by defining $u_{la}=\mu_{a}u_{\infty}\bar{h}_{0}d^{2}\bar{F}_{a}/d\eta^{2}|_{\eta=0}/(\mu_{l}\delta_{0})$.
Therefore, the solutions of the undisturbed velocity and temperature profiles in the water film are linear in $y_{*}$,
that is,
$\bar{u}_{l*}=y_{*}$ and $\bar{T}_{l*}=-y_{*}$.
This is in agreement with the more usual lubrication approach for describing icing with shear.
\cite{Oron97, Myers02_1, Myers02_2, Myers04}
From $a+b=0$, $b-a=-1/2$ and $c-a=-1/2$, we obtain $a=1/4$, $b=-1/4$, $c=-1/4$. The value of $a$ coincides with that stated in previous papers. \cite{Nelson95,Tsao97}
$C_{1}$, $C_{2}$ and $C_{3}$ are determined from Eqs. (\ref{eq:Qoverl}), (\ref{eq:C1-C2}) and (\ref{eq:C1-C3}).
Hence
$\bar{h}_{0}$ and $u_{la}$
can be expressed as follows:
\begin{equation}
\bar{h}_{0}=\left[
\frac{2\mu_{l}(2\nu_{a})^{1/2}}{\mu_{a}\frac{d^{2}\bar{F}_{a}}{d\eta^{2}}\Big|_{\eta=0}}
\right]^{1/2}(Q/l_{w})^{1/2}u_{\infty}^{-3/4}x^{1/4}, \qquad
u_{la}=\left[\frac{2\mu_{a}\frac{d^{2}\bar{F}_{a}}{d\eta^{2}}\Big|_{\eta=0}}{\mu_{l}(2\nu_{a})^{1/2}}
\right]^{1/2}(Q/l_{w})^{1/2}u_{\infty}^{3/4}x^{-1/4}.
\label{eq:h0-ula}
\end{equation}
It is found that $\bar{h}_{0}$ and $u_{la}$ depend on $Q/l_{w}$ and $u_{\infty}$, and vary slowly with $x$.
We assume that the scaling of $\bar{h}_{0}$ and $u_{la}$ for $x$ holds except for the very vicinity of water source.
Equations (\ref{eq:geq-basicFa}) and (\ref{eq:bc-basic_air}) yield $d^{2}\bar{F}_{a}/d\eta^{2}|_{\eta=0}=0.47$.
The shear rate for the undisturbed water layer is then given by
\begin{equation}
\frac{\partial\bar{u}_{l}}{\partial y}\Big|_{y=0}
=\frac{u_{la}}{\bar{h}_{0}}\frac{d\bar{u}_{l*}}{dy_{*}}\Big|_{y_{*}=0}
=\left(\frac{1}{2\nu_{a}x}\right)^{1/2}
\frac{\mu_{a}}{\mu_{l}}\frac{d^{2}\bar{F}_{a}}{d\eta^{2}}\Big|_{\eta=0}u_{\infty}^{3/2},
\label{eq:ul0overh0}
\end{equation}
and its value is in the range 30.6 to 449.3 ${\rm s}^{-1}$ for the range of $u_{\infty}=5$ to 30 m/s at $x=0.1$ m. Hence the value of the time defined by the inverse of shear rate lies within the range of $2.2\times 10^{-3}$ to $3.3 \times 10^{-2}$ s for the above parameters.
Using Eq. (\ref{eq:C1-C3}), $C_{1}=\bar{h}_{0}x^{-1/4}$ and $\delta_{0}=(2\nu_{a}x/u_{\infty})^{1/2}$, we can express
\begin{equation}
T_{sl}-T_{la}
=-C_{1}\frac{K_{a}}{K_{l}}\left(\frac{u_{\infty}}{2\nu_{a}}\right)^{1/2}\bar{G}_{a*}T_{\infty}x^{-1/4}
=-\frac{K_{a}}{K_{l}}\frac{\bar{h}_{0}}{\delta_{0}/\bar{G}_{a*}}T_{\infty},
\label{eq:Tla}
\end{equation}
where $\bar{G}_{a*}\equiv -d\bar{T}_{a*}/d\eta|_{\eta=0}$.
Since the undisturbed part of temperature gradient within the ice does not exist in the model herein,
the undisturbed part of Eq. (\ref{eq:heatflux-zeta}) yields
$L\bar{V}=K_{l}(T_{sl}-T_{la})/\bar{h}_{0}$,
into which Eq. (\ref{eq:Tla}) is substituted to obtain the undisturbed ice growth rate
\begin{equation}
\bar{V}=-\frac{K_{a}T_{\infty}}{L(\delta_{0}/\bar{G}_{a*})}.
\label{eq:V}
\end{equation}
Equations (\ref{eq:Tla}) and (\ref{eq:V}) are the same form as those in a previous paper, \cite{Ueno10}
but the value of $\bar{G}_{a*}$ is different. From Eqs. (\ref{eq:geq-basicFa}), (\ref{eq:geq-basicTa}) and (\ref{eq:bc-basic_air}), we obtain $\bar{G}_{a*}=0.413$ for $Pr_{a}=0.7$. The variation of $T_{la}$ and $\bar{V}$ against $u_{\infty}$ is shown in Table \ref{tab:tableII}.
\subsubsection{Equations and boundary conditions for disturbed flows and temperatures in the air and water film}
When the assumed forms of $\psi_{a}$ and $T_{a}$ are substituted into the complete equations (\ref{eq:geq-ua}), (\ref{eq:geq-va}) and (\ref{eq:geq-Ta}), the differential equations for the amplitudes $f_{a}$ and $H_{a}$ are obtained:
\begin{eqnarray}
\frac{d^{4}f_{a}}{d\eta^4}
&=&-\bar{F}_{a}\frac{d^{3}f_{a}}{d\eta^{3}}
+\left\{2k_{a*}^{2}-(2-i k_{a*}\Rey_{a})\frac{d\bar{F}_{a}}{d\eta}\right\}\frac{d^{2}f_{a}}{d\eta^{2}} \nonumber \\
&& +\left\{k_{a*}^{2}\left(\bar{F}_{a}+2\eta\frac{d\bar{F}_{a}}{d\eta}\right)
-\frac{d^{2}\bar{F}_{a}}{d\eta^{2}}\right\}\frac{df_{a}}{d\eta}
-\left\{k_{a*}^{4}+i k_{a*}\Rey_{a}\left(k_{a*}^{2}\frac{d\bar{F}_{a}}{d\eta}+\frac{d^{3}\bar{F}_{a}}{d\eta^{3}}\right)\right\}f_{a},\nonumber \\
\label{eq:geq-fa}
\end{eqnarray}
\begin{eqnarray}
\frac{d^{2}(\bar{G}_{a*}H_{a})}{d\eta^{2}}
&=&-Pr_{a}\bar{F}_{a}\frac{d(\bar{G}_{a*}H_{a})}{d\eta}
+\left\{k_{a*}^{2}+Pr_{a}(-1+i k_{a*}\Rey_{a})\frac{d\bar{F}_{a}}{d\eta}\right\}(\bar{G}_{a*}H_{a}) \nonumber \\
&& -i k_{a*}Pr_{a}\Rey_{a}\frac{d\bar{T}_{a*}}{d\eta}f_{a},
\label{eq:geq-Ha}
\end{eqnarray}
where $k_{a*}=k\delta_{0}$ is the dimensionless wave number normalized by the length $\delta_{0}$.
The disturbed part of Eq. (\ref{eq:bc-ul-vl-ua-va-xi}) and the boundary conditions
$u'_{a}|_{y=\infty}=\partial \psi'_{a}/\partial y|_{y=\infty}=0$ and
$v'_{a}|_{y=\infty}=-\partial \psi'_{a}/\partial x|_{y=\infty}=0$
yield
\begin{equation}
\frac{df_{a}}{d\eta}\Big|_{\eta=0}=-\frac{d^{2}\bar{F}_{a}}{d\eta^{2}}\Big|_{\eta=0}, \qquad
f_{a}|_{\eta=0}=0, \qquad
\frac{df_{a}}{d\eta}\Big|_{\eta=\infty}=0, \qquad
f_{a}|_{\eta=\infty}=0.
\label{eq:bc-fa}
\end{equation}
We note that, as shown in Table \ref{tab:tableI}, since the water surface velocity $u_{la}$ is significantly lower than the free stream velocity $u_{\infty}$, $u_{a}|_{y=\xi}=0$ and $v_{a}|_{y=\xi}=0$ are good approximation, from which the first and second equations in Eq. (\ref{eq:bc-fa}) are obtained.
Furthermore, the disturbed part of Eq. (\ref{eq:bc-Tla}) and the boundary condition $T_{a}'|_{y=\infty}=0$ give
\begin{equation}
H_{a}|_{\eta=0}=1, \qquad
H_{a}|_{\eta=\infty}=0.
\label{eq:bc-Ha}
\end{equation}
On the other hand, when the assumed forms of $\psi_{l}$ and $T_{l}$ are substituted into Eqs. (\ref{eq:geq-ul}) (\ref{eq:geq-vl}) and (\ref{eq:geq-Tl}) and neglecting the terms with $\Rey_{l}\bar{h}_{0}/(\Rey_{a}\delta_{0}) \ll 1$, the disturbed parts yield the equation for the amplitudes $f_{l}$ and $H_{l}$:
\begin{equation}
\frac{d^{4}f_{l}}{dy_{*}^{4}}
=\left(2k_{l*}^{2}+ik_{l*} \Rey_{l}\bar{u}_{l*}\right)\frac{d^{2}f_{l}}{dy_{*}^{2}}
-\left\{k_{l*}^{4}+ik_{l*} \Rey_{l}\left(k_{l*}^{2}\bar{u}_{l*}+\frac{d^{2}\bar{u}_{l*}}{dy_{*}^{2}}\right)\right\}f_{l},
\label{eq:geq-fl}
\end{equation}
\begin{equation}
\frac{d^{2}H_{l}}{dy_{*}^{2}}
=\left(k_{l*}^{2}+ik_{l*} \Pec_{l}\bar{u}_{l*}\right)H_{l}
-ik_{l*} \Pec_{l}\frac{d\bar{T}_{l*}}{d y_{*}}f_{l},
\label{eq:geq-Hl}
\end{equation}
where $k_{l*}=k\bar{h}_{0}$ is the dimensionless wave number normalized by the length $\bar{h}_{0}$,
$\bar{T}_{l*}(y_{*})=-y_{*}$ and $\Pec_{l}\equiv u_{la}\bar{h}_{0}/\kappa_{l}$ is the Peclet number.
We note that Eqs. (\ref{eq:geq-fl}) and (\ref{eq:geq-Hl}) are in the same form as found in previous papers, \cite{Ueno03, Ueno04, Ueno07, UFYT10,Ueno10}
but the form of $\bar{u}_{l*}$ is different. In this paper, $\bar{u}_{l*}=y_{*}$ is used.
Using Eq. (\ref{eq:h0-ula}), the Reynolds number and the Peclet number of the water
can be written as $\Rey_{l}=(2/\nu_{l})Q/l_{w}$ and $\Pec_{l}=(2/\kappa_{l})Q/l_{w}$, which are independent of $u_{\infty}$ and $x$.
Linearization of the disturbed parts of Eq. (\ref{eq:bc-ul-vl-zeta}) at $y=0$ and Eqs. (\ref{eq:bc-shear-stress-xi}) and (\ref{eq:bc-normal-stress-xi}) at $y=\bar{h}_{0}$ yield the boundary conditions for $f_{l}$:
\begin{equation}
\frac{df_{l}}{dy_{*}}\Big|_{y_{*}=0}+1=0, \qquad
f_{l}|_{y_{*}=0}=0,
\label{eq:bc-fl0-dfl0}
\end{equation}
\begin{equation}
\frac{d^{2}f_{l}}{dy_{*}^{2}}\Big|_{y_{*}=1}
+\left(k_{l*}^{2}+\Sigma_{a}\right)f_{l}|_{y_{*}=1}=0,
\label{eq:bc-shearstress}
\end{equation}
\begin{eqnarray}
\frac{d^{3}f_{l}}{dy_{*}^{3}}\Big|_{y_{*}=1}
-\left(3k_{l*}^{2}+ik_{l*} \Rey_{l}\right)\frac{df_{l}}{dy_{*}}\Big|_{y_{*}=1} \nonumber \\
+ik_{l*}\Rey_{l}
\left(1+\frac{1}{Fr^{2}}+Wek_{l*}^{2}+\Pi_{a}
\right)f_{l}|_{y_{*}=1}=0,
\label{eq:bc-normalstress}
\end{eqnarray}
where
\begin{equation}
\Sigma_{a}=\frac{\mu_{a}}{\mu_{l}}\frac{u_{\infty}}{u_{la}}\left(\frac{\bar{h}_{0}}{\delta_{0}}\right)^{2}
\frac{d^{2}f_{a}}{d\eta^{2}}\Big|_{\eta=0},
\label{eq:shearstress_a}
\end{equation}
\begin{eqnarray}
\Pi_{a}=-\frac{\rho_{a}}{\rho_{l}}\left(\frac{u_{\infty}}{u_{la}}\right)^{2}\frac{\bar{h}_{0}}{\delta_{0}}
\frac{1+ik_{a*}\Rey_{a}}{1+(k_{a*}\Rey_{a})^{2}}
\left\{
\frac{d^{3}f_{a}}{d\eta^{3}}\Big|_{\eta=0}
+3k_{a*}^{2}\frac{d^{2}\bar{F}_{a}}{d\eta^{2}}\Big|_{\eta=0}
\right\}, \nonumber \\
\label{eq:normalstress_a}
\end{eqnarray}
$Fr=u_{la}/(g\bar{h}_{0})^{1/2}$ is the Froude number, and $We=\gamma/(\rho_{l}u_{la}^{2}\bar{h}_{0})$ is the Weber number.
Linearization of the disturbed part of Eqs. (\ref{eq:bc-Tla}) and (\ref{eq:heatflux-xi}) at $y=\bar{h}_{0}$ yields
\begin{equation}
H_{l}|_{y_{*}=1}+f_{l}|_{y_{*}=1}=0,
\label{eq:Tla-xi}
\end{equation}
\begin{equation}
\frac{dH_{l}}{dy_{*}}\Big|_{y_{*}=1}
-\frac{\bar{h}_{0}}{\delta_{0}}\left(-\frac{dH_{a}}{d\eta}\Big|_{\eta=0}\right)f_{l}|_{y_{*}=1}=0.
\label{eq:heatflux-xi-h0}
\end{equation}
In deriving Eqs. (\ref{eq:bc-shearstress}), (\ref{eq:bc-normalstress}), (\ref{eq:Tla-xi}) and (\ref{eq:heatflux-xi-h0}),
the relation between the amplitude of the water-air interface and that of the ice-water interface,
$\xi_{k}=-f_{l}|_{y_{*}=1}\zeta_{k}$,
is used, which is derived from the linearization of Eq. (\ref{eq:bc-kinematic-xi}) at $y=\bar{h}_{0}$. \cite{Ueno03, Ueno04, Ueno07, UFYT10, Ueno10}
It should be noted that the water flow is affected by the terms $\Sigma_{a}$ in (\ref{eq:bc-shearstress}) and $\Pi_{a}$ in (\ref{eq:bc-normalstress}) due to the tangential and normal air shear stress disturbances, respectively.
The coupling between the air and water flows affects the disturbed part of temperature distribution in the water layer through the boundary conditions found in Eqs. (\ref{eq:Tla-xi}) and (\ref{eq:heatflux-xi-h0}).
\subsubsection{Dispersion relation}
The disturbed parts of Eqs. (\ref{eq:Tsl}) and (\ref{eq:heatflux-zeta}) give
the dimensionless amplification rate
$\sigma_{*}^{(r)}\equiv \sigma^{(r)}/(K_{a}\bar{G}_{a}/L\bar{h}_{0})$,
and the dimensionless phase velocity
$v_{p*}\equiv -\sigma^{(i)}/(kK_{a}\bar{G}_{a}/L)$, \cite{Ueno03, Ueno04, Ueno07, UFYT10, Ueno10}
\begin{equation}
\sigma_{*}^{(r)}=-\frac{dH_{l}^{(r)}}{dy_{*}}\Big|_{y_{*}=0}+K^{s}_{l}k_{l*}(H_{l}^{(r)}|_{y_{*}=0}-1),
\label{eq:amp}
\end{equation}
\begin{equation}
v_{p*}=-\frac{1}{k_{l*}}\left(-\frac{dH_{l}^{(i)}}{dy_{*}}\Big|_{y_{*}=0}+K^{s}_{l}k_{l*}H_{l}^{(i)}|_{y_{*}=0}\right),
\label{eq:phasevel}
\end{equation}
where $H_{l}^{(r)}$ and $H_{l}^{(i)}$ are the real and imaginary parts of $H_{l}$, and $K^{s}_{l}=K_{s}/K_{l}=3.96$ is the ratio of the thermal conductivity of ice to that of water.
The numerical procedure for calculating Eqs. (\ref{eq:amp}) and (\ref{eq:phasevel}) is as follows:
First, Eqs. (\ref{eq:geq-basicFa}), (\ref{eq:geq-basicTa}), (\ref{eq:geq-fa}) and (\ref{eq:geq-Ha}) are simultaneously solved for a given $u_{\infty}$ and $x$ with boundary conditions (\ref{eq:bc-basic_air}), (\ref{eq:bc-fa}) and (\ref{eq:bc-Ha}). The derived solutions $\bar{F}_{a}$ and $f_{a}$ are substituted into Eqs. (\ref{eq:shearstress_a}) and (\ref{eq:normalstress_a}). Using the boundary conditions (\ref{eq:bc-fl0-dfl0}), (\ref{eq:bc-shearstress}) and (\ref{eq:bc-normalstress}), Eq. (\ref{eq:geq-fl}) is solved.
Then Eq. (\ref{eq:geq-Hl}) is solved with the boundary conditions (\ref{eq:Tla-xi}) and (\ref{eq:heatflux-xi-h0}), using solutions $f_{l}$ and $H_{a}$.
Finally, substituting solution $H_{l}$ into Eqs. (\ref{eq:amp}) and (\ref{eq:phasevel}) and replacing $k_{l*}$ with $(\bar{h}_{0}/\delta_{0})k_{a*}$, Eqs. (\ref{eq:amp}) and (\ref{eq:phasevel}) are obtained with respect to $k_{a*}$.
\begin{table}[ht]
\caption{\label{tab:tableI}
Variation of a length, $\delta_{0}$,
thickness of water film, $\bar{h}_{0}$,
water-air surface velocity, $u_{la}$,
inverse of square of the Froude number, $1/Fr^{2}=g\bar{h}_{0}/u_{la}^{2}$,
the Weber number, $We=\gamma/(\rho_{l}u_{la}^{2}\bar{h}_{0})$,
the Reynolds number of air, $\Rey_{a}=u_{\infty}\delta_{0}/\nu_{a}$,
against the free stream velocity, $u_{\infty}$,
for $Q/l_{w}=1000$ $[{\rm (ml/h)/cm}]$ and $x=0.1$ m.
The corresponding values of the Reynolds number and the Peclet number of water are
$\Rey_{l}=u_{la}\bar{h}_{0}/\nu_{l}=31$ and
$\Pec_{l}=u_{la}\bar{h}_{0}/\kappa_{l}=418$, respectively.}
\begin{ruledtabular}
\begin{tabular}{ccccccc}
$u_{\infty}$ (m/s) & $\delta_{0}$ ($\mu$m) & $\bar{h}_{0}$ ($\mu$m) & $u_{la}$ (cm/s)
& $1/Fr^{2}$ & $We$ & $\Rey_{a}$ \\
5 & 721 & 1348 & 4.1 & 7.78 & 33.19 & 277 \\
10 & 510 & 802 & 6.9 & 1.64 & 19.74 & 392 \\
15 & 416 & 591 & 9.4 & 0.66 & 14.56 & 480 \\
20 & 361 & 477 & 11.7 & 0.34 & 11.74 & 555 \\
25 & 322 & 403 & 13.8 & 0.21 & 9.93 & 620 \\
30 & 294 & 352 & 15.8 & 0.14 & 8.66 & 679 \\
100 & 161 & 143 & 39.0 & 0.01 & 3.51 & 1240 \\
\end{tabular}
\end{ruledtabular}
\end{table}
\section{\label{sec:results}Results}
Figures \ref{fig:qoverl-ka-Sigma-Pi} (a) and (b) show the variation of the values $1/Fr^{2}$, $Wek_{l*}^{2}$, $\Sigma_{a}^{(r)}$, $\Sigma_{a}^{(i)}$, $\Pi_{a}^{(r)}$ and $\Pi_{a}^{(i)}$ against the water supply rate per width $Q/l_{w}$ in the range of 10 to 1000 [(ml/h)/cm], in the case of $u_{\infty}=5$ m/s and $u_{\infty}=20$ m/s, respectively. Here, $\Sigma_{a}^{(r)}$, $\Pi_{a}^{(r)}$and $\Sigma_{a}^{(i)}$, $\Pi_{a}^{(i)}$ are the real and imaginary parts of Eqs. (\ref{eq:shearstress_a}) and (\ref{eq:normalstress_a}), respectively.
From Eq. (\ref{eq:h0-ula}), the variation of Eqs. (\ref{eq:shearstress_a}), (\ref{eq:normalstress_a}), and parameters $Fr$ and $We$ with respect to $Q/l_{w}$ and $u_{\infty}$ are as follows:
$\Sigma_{a} \sim (Q/l_{w})^{1/2}$,
$\Pi_{a} \sim (Q/l_{w})^{-1/2}$,
$1/Fr^{2} \sim (Q/l_{w})^{-1/2}$,
$We \sim (Q/l_{w})^{-3/2}$
for a given $u_{\infty}$,
and
$\Sigma_{a} \sim u_{\infty}^{-1/4}$,
$\Pi_{a} \sim u_{\infty}^{-3/4}$,
$1/Fr^{2} \sim u_{\infty}^{-9/4}$,
$We \sim u_{\infty}^{-3/4}$
for a given $Q/l_{w}$.
It is found that as $Q/l_{w}$ increases, the value of $\Sigma_{a}$ increases, while other parameters decrease. Also, as $u_{\infty}$ increases, the value of $\Sigma_{a}$ decreases much slower than the other parameters.
Hence, Fig. \ref{fig:qoverl-ka-Sigma-Pi} (b) shows that $\Sigma_{a}$ is not negligible compared to the other parameters as $Q/l_{w}$ and $u_{\infty}$ increase.
Therefore, we use the value $Q/l_{w}=1000$ [(ml/h)/cm] throughout this paper, which is also in the same order as that employed in the experiments. \cite{Streitz02}
On the other hand, Fig. \ref{fig:qoverl-ka-Sigma-Pi} (c) and (d) show the variation of the values $1/Fr^{2}$, $Wek_{l*}^{2}$, $\Sigma_{a}^{(r)}$, $\Sigma_{a}^{(i)}$, $\Pi_{a}^{(r)}$ and $\Pi_{a}^{(i)}$ against the dimensionless wave number $k_{a*}$ for $Q/l_{w}=1000$ [(ml/h)/cm]. When plotting $Wek_{l*}^{2}$ with respect to $k_{a*}$, the relation $k_{l*}=(\bar{h}_{0}/\delta_{0})k_{a*}$ is used.
In the case of $u_{\infty}=5$ m/s, as shown in Fig. \ref{fig:qoverl-ka-Sigma-Pi} (c), $1/Fr^{2}$ and $Wek_{l*}^{2}$ are dominant terms in Eq. (\ref{eq:bc-normalstress}). As $u_{\infty}$ increases, the values of $1/Fr^{2}$ and $We$ decrease as shown in Table \ref{tab:tableI}. For example, for $u_{\infty}=20$ m/s, Fig. \ref{fig:qoverl-ka-Sigma-Pi} (d) shows that the magnitude of $\Sigma_{a}$ and $\Pi_{a}$ in Eqs. (\ref{eq:bc-shearstress}) and (\ref{eq:bc-normalstress}) are not negligible compared to $1/Fr^{2}$ and $Wek_{l*}^{2}$.
In Figs. \ref{fig:ka-xi-Thetaxi}, \ref{fig:ka-amp-lambda} and \ref{fig:heat-coefficient}
we consider two cases: One takes into account the effect of the air stress disturbances $\Sigma_{a}$ and $\Pi_{a}$, and
the other does not.
We will demonstrate in the following sections that this leads to critically different results for the shape of the water-air interface, the growth conditions of the ice-water interface and the heat transfer coefficient at the water-air interface.
\begin{figure}[ht]
\begin{center}
\includegraphics[width=7cm,height=7cm,keepaspectratio,clip]{qoverl-ua5-ka2-InvFr-We-Sigma-Pi-x0p1.eps}\qquad
\includegraphics[width=7cm,height=7cm,keepaspectratio,clip]{qoverl-ua20-ka2-InvFr-We-Sigma-Pi-x0p1.eps}\\[5mm]
\includegraphics[width=7cm,height=7cm,keepaspectratio,clip]{ka-ua5-InvFr-We-Sigma-Pi-x0p1-q1000.eps}\qquad
\includegraphics[width=7cm,height=7cm,keepaspectratio,clip]{ka-ua20-InvFr-We-Sigma-Pi-x0p1-q1000.eps}
\end{center}
\caption{Variation of $1/Fr^{2}$, $Wek_{l*}^{2}$, $\Sigma_{a}^{(r)}$, $\Sigma_{a}^{(i)}$, $\Pi_{a}^{(r)}$ and $\Pi_{a}^{(i)}$ against $Q/l_{w}$ for $k_{a*}=0.2$ and $x=0.1$ m, in the case of (a) $u_{\infty}=5$ m/s and (b) $u_{\infty}=20$ m/s.
The variation of $1/Fr^{2}$, $Wek_{l*}^{2}$, $\Sigma_{a}^{(r)}$, $\Sigma_{a}^{(i)}$, $\Pi_{a}^{(r)}$ and $\Pi_{a}^{(i)}$ against $k_{a*}$ for $Q/l_{w}=1000$ $[{\rm (ml/h)/cm}]$ and $x=0.1$ m, in the case of (c) $u_{\infty}=5$ m/s and (d) $u_{\infty}=20$ m/s.
Here $k_{a*}=0.2$ corresponds to a wavelength of about 2 cm for $u_{\infty}=5$ m/s and about 1 cm for $u_{\infty}=20$ m/s at $x=0.1$ m.
}
\label{fig:qoverl-ka-Sigma-Pi}
\end{figure}
\subsection{\label{sec:shape}The shape of the water-air interface}
\begin{figure}[ht]
\begin{center}
\includegraphics[width=8cm,height=8cm,keepaspectratio,clip]{ka-xi-ua5-20-q1000.eps}\hspace{5mm}
\includegraphics[width=7cm,height=7cm,keepaspectratio,clip]{Thetaxi-uinf-q1000.eps}\hspace{5mm}
\end{center}
\caption{
For $Q/l_{w}=1000$ $[{\rm (ml/h)/cm}]$, $u_{\infty}=5$ m/s and $u_{\infty}=20$ m/s,
(a) represents the variation of amplitude $|f_{l}|_{y_{*}=1}|$ against $k_{a*}$.
Here $k_{a*}=1.0$ corresponds to a wavelength of about 4.5 mm for $u_{\infty}=5$ m/s, and about 2.3 mm for $u_{\infty}=20$ m/s at $x=0.1$ m.
(b) represents the variation of phase difference $\Theta_{\xi_{*}}$ between the ice-water and water-air interfaces against the free stream velocity $u_{\infty}$ for the most unstable mode.
The solid curves consider the effect of the tangential and normal air shear stress disturbances on the water-air interface. The dashed curves do not consider this effect.
}
\label{fig:ka-xi-Thetaxi}
\end{figure}
It is supposed that a dimensionless small disturbance of the ice-water interface has a sinusoidal form:
\begin{equation}
y_{*}=\zeta_{*}=\delta_{b}\Imag[{\rm exp}(\sigma_{*}t_{*}+ik_{l*}x_{*})]
=\delta_{b}(t_{*})\sin[k_{l*}(x_{*}-v_{p*}t_{*})],
\label{eq:zeta}
\end{equation}
where $\delta_{b}\equiv \zeta_{k}/\bar{h}_{0}$ is an infinitesimal initial amplitude,
$\sigma_{*}=\sigma/(K_{a}\bar{G}_{a}/L\bar{h}_{0})$,
$t_{*}=(\bar{V}/\bar{h}_{0})t$, $x_{*}=x/\bar{h}_{0}$,
$\delta_{b}(t_{*})\equiv \delta_{b}{\rm exp}(\sigma^{(r)}_{*}t_{*})$,
and $\Imag$ denotes the imaginary part of its argument.
Since the water film is very thin, the deformed ice-water interface causes a disturbance of the water-air interface:
\begin{eqnarray}
y_{*}=\xi_{*}&=&1+\Imag[\delta_{t}{\rm exp}(\sigma_{*}t_{*}+ik_{l*}x_{*})] \nonumber \\
&=&1-\delta_{b}(t_{*})
\left\{
f_{l}^{(r)}|_{y_{*}=1}\sin[k_{l*}(x_{*}-v_{p*}t_{*})]
+f_{l}^{(i)}|_{y_{*}=1}\cos[k_{l*}(x_{*}-v_{p*}t_{*})]
\right\} \nonumber \\
&=&1+\delta_{b}(t_{*})|f_{l}|_{y_{*}=1}|\sin[k_{l*}(x_{*}-v_{p*}t_{*})-\Theta_{\xi_{*}}],
\label{eq:xi}
\end{eqnarray}
where
$f_{l}^{(r)}$ and $f_{l}^{(i)}$ are the real and imaginary parts of $f_{l}$, respectively, $|f_{l}|_{y_{*}=1}|=[(f_{l}^{(r)}|_{y_{*}=1})^{2}+(f_{l}^{(i)}|_{y_{*}=1})^2]^{1/2}$ is the amplitude and
$\cos\Theta_{\xi_{*}}=-f_{l}^{(r)}|_{y_{*}=1}/|f_{l}|_{y_{*}=1}|$, $\sin\Theta_{\xi_{*}}=f_{l}^{(i)}|_{y_{*}=1}/|f_{l}|_{y_{*}=1}|$,
$\Theta_{\xi_{*}}$ represents a phase difference between the water-air and ice-water interfaces.
When deriving the second equation in Eq. (\ref{eq:xi}), the relation
$\delta_{t}\equiv \xi_{k}/\bar{h}_{0}=-f_{l}|_{{y_{*}}=1}\delta_{b}$
is used. \cite{Ueno03, Ueno04, Ueno07, UFYT10, Ueno10}
Since $f_{l}|_{y_{*}=1}$ depends on the wave number,
the amplitude and phase change according to the wavelength of the ice-water interface disturbance.
Figure \ref{fig:ka-xi-Thetaxi} (a) shows that the water-air interface tends to become flat as $k_{a*}$ increases, due to the action of gravity, surface tension and tangential and normal air shear stress disturbances on the water-air interface.
In the case of $u_{\infty}=5$ m/s,
since the effect of air shear stress disturbances can be neglected, as shown in Fig. \ref{fig:qoverl-ka-Sigma-Pi} (c), gravity and surface tension are dominant resisting forces for displacement of the water-air interface.
On the other hand, as $u_{\infty}$ increases, a region of $k_{a*}$ appears, where the action of tangential and normal air shear stress disturbances on the water-air interface is dominant compared to that of gravity and surface tension, as shown in Fig. \ref{fig:qoverl-ka-Sigma-Pi} (d). For example, in the case of $u_{\infty}=20$ m/s in Fig. \ref{fig:ka-xi-Thetaxi} (a), there is a region in the solid curve where the amplitude does not decrease with an increase in $k_{a*}$.
If we neglect the air shear stress disturbances, the amplitude is overestimated as shown by the dashed curve in Fig. \ref{fig:ka-xi-Thetaxi} (a). As $k_{a*}$ increases, the difference between the solid and dashed curves becomes small because the surface tension $Wek_{l*}^{2}$ is finally most dominant.
Figure \ref{fig:ka-xi-Thetaxi} (b) shows the variation of the phase difference $\Theta_{\xi_{*}}$ between the ice-water and water-air interfaces against $u_{\infty}$ for the most unstable mode (see {\ref{sec:amplification}), with (solid curve) and without (dashed curve) the effect of air shear stress disturbances. In the case of $u_{\infty}=5$ m/s, an upward phase shift of the water-air interface relative to the ice-water interface is large, as shown in Fig. \ref{fig:heatflux-sla} (a).
The solid curve shows that the phase difference
decreases with increasing $u_{\infty}$ and the sign of $\Theta_{\xi_{*}}$ changes from negative to positive at about $u_{\infty}=27$ m/s.
An example of the configuration of two interfaces at $u_{\infty}=20$ m/s appears in Fig. \ref{fig:heatflux-sla} (b). The decrease of phase difference $\Theta_{\xi_{*}}$ is also due to the effect of air shear stress disturbances on the water-air interface. On the other hand, if we neglect the effect of the air shear stress disturbances, the phase shift is still large even for large $u_{\infty}$, as shown by the dashed curve in Fig. \ref{fig:ka-xi-Thetaxi} (b).
\subsection{\label{sec:amplification}Amplification rate of the ice-water interface disturbance}
Figure \ref{fig:ka-amp-lambda} (a) shows the variation of numerically obtained dimensionless amplification rates $\sigma_{*}^{(r)}$ against the dimensionless wave number $k_{a*}$.
The solid curves represent $\sigma_{*}^{(r)}$, taking into account the effect of tangential and normal air shear stress disturbances on the water-air interface. If we neglect the air shear stress disturbances, the dashed curves are obtained.
In the case of $u_{\infty}=5$ m/s, the difference is negligible.
However, if the air shear stress disturbances are neglected, the magnitude of $\sigma_{*}^{(r)}$ is overestimated with increasing $u_{\infty}$.
One expects to observe an ice pattern with a wave number at which the amplification rate is the maximum.
For example, at $u_{\infty}=20$ m/s, $\sigma_{*}^{(r)}$ acquires a maximum value $\sigma^{(r)}_{*\rm max}=17.7$ at $k_{a*}=0.22$. Since the wave number $k$ is normalized by $\delta_{0}$, the corresponding wavelength of the ice pattern is 1.03 cm from $\lambda=2\pi\delta_{0}/k_{a*}$. Here, the value of $\delta_{0}=(2\nu_{a}x/u_{\infty})^{1/2}=361$ $\mu$m estimated from $x=0.1$ m and $u_{\infty}=20$ m/s is used.
The magnitude of $v_{p*}$ is defined from the wave number at which $\sigma_{*}^{(r)}$ acquires a maximum value.
At $k_{a*}=0.22$, we obtain $v_{p*}=-96.6$, and hence the displacement of the ice-water interface after the time $t_{*}=1/\sigma_{* \rm max}^{(r)}$ is $\Delta x_{*}=v_{p*}/\sigma_{* \rm max}^{(r)}=-5.4$. The ice pattern will move in the direction opposite to the water flow (see Fig. \ref{fig:heatflux-sla}). The variation of $\sigma_{* \rm max}^{(r)}$, $\lambda$, $v_{p*}$ and $\Delta x_{*}$ against $u_{\infty}$ is shown in Table \ref{tab:tableII}.
\begin{figure}[ht]
\begin{center}
\includegraphics[width=8cm,height=8cm,keepaspectratio,clip]{ka-amp-x0p1-q1000.eps}\hspace{5mm}
\includegraphics[width=7cm,height=7cm,keepaspectratio,clip]{lambda-uinf.eps}
\end{center}
\caption{For $Q/l_{w}=1000$ $[{\rm (ml/h)/cm}]$ and $x=0.1$ m,
(a) dimensionless amplification rate $\sigma_{*}^{(r)}=\sigma^{(r)}/(K_{a}\bar{G}_{a}/L\bar{h}_{0})$ versus dimensionless wave number $k_{a*}=k\delta_{0}$,
(b) variation of wavelength of ripples against the free stream velocity $u_{\infty}$.
The solid curves consider the effect of the tangential and normal air shear stress disturbances on the water-air interface, and the dashed curves do not consider this effect.}
\label{fig:ka-amp-lambda}
\end{figure}
It is found from Fig. \ref{fig:ka-amp-lambda} (b) and Table \ref{tab:tableII} that the wavelength shortens with increasing $u_{\infty}$.
Wavelike ice patterns with various roughness spacings and heights were experimentally observed by changing the wind speed and slope of an inclined plane. \cite{Streitz02} For a wind speed of 16 km/h=4.4 m/s, the roughness spacing is increased as the plane slope is decreased (the roughness spacing for smooth-ice base is about 3 cm at about $3^{\circ}$, see Fig. 10 in Ref. \onlinecite{Streitz02}) and for the plane slope of $8^{\circ}$, the roughness spacing is decreased as the wind speed is increased (see Fig. 11 in Ref. \onlinecite{Streitz02}). The latter result is consistent with the theoretical prediction, except that the results herein are only those obtained with a slope of $0^{\circ}$.
Since the values $\sigma_{*}^{(r)}$ in Table \ref{tab:tableII} are much larger than those
predicted by the morphological instability triggered by thermal diffusion at the water-air interface in previous papers, \cite{Ueno03, Ueno04, Ueno07, UFYT10} the ice-water interface instability herein is enhanced by the flow in the water film.
On the other hand, as shown in Figs. \ref{fig:qoverl-ka-Sigma-Pi} (c) and (d), the surface tension $Wek_{l*}^{2}$ is most dominant to suppress the water-air disturbance with increasing $k_{a*}$ and stabilizes the corresponding ice-water interface disturbance.
The solid and dashed curves in Fig. \ref{fig:ka-amp-lambda} (a) show that as $u_{\infty}$ increases, the wave number at which $\sigma_{*}^{(r)}$ vanishes, that is, the neutral stability point is shifted to higher wave number.
This is because the value of the Weber number $We$ decreases with an increase in $u_{\infty}$, as indicated in Table \ref{tab:tableI}, hence the stabilization due to the surface tension $Wek_{l*}^{2}$ becomes more effective for higher wave numbers.
As shown in Figs. \ref{fig:qoverl-ka-Sigma-Pi} (c) and (d), since $\Pi_{a}^{(r)}$ has negative values with respect to $k_{a*}$,
the value of $Wek_{l*}^{2}+\Pi_{a}^{(r)}$ in Eq. (\ref{eq:bc-normalstress}) decreases as $u_{\infty}$ increases.
This means that the stabilization due to the surface tension $Wek_{l*}^{2}$ is weakened by normal air shear stress disturbance $\Pi_{a}^{(r)}$. Hence, the wave number of the neutral stability point in the solid curves is shifted to the higher wave number compared to that in the dashed curves with an increase in $u_{\infty}$.
That is why the wavelength evaluated from the most unstable mode becomes shorter as $u_{\infty}$ increases.
It should be noted that the magnitude of $\sigma_{*}^{(r)}$ is decreased by the effect of the tangential and normal air shear stress disturbances. However, the effect of air shear stress disturbances on the wavelength does not make a significant difference, as shown in Fig. \ref{fig:ka-amp-lambda} (b).
On a static structure, typical wind speed is in the order of 20 m/s. \cite{Myers02_1} The wind speed in the aufeis formation experiments by Streitz and Ettema \cite{Streitz02} was varied up to 48 km/h=13.3 m/s. On the other hand, in aircraft icing a typical value is around 100 m/s. \cite{Myers02_1}
By applying our analysis to the case $u_{\infty}=100$ m/s, assuming that the air stream flow remains laminar as shown in Fig. \ref{fig:ice-water-air}, the results shown in
Tables \ref{tab:tableI} and \ref{tab:tableII} are obtained.
On the other hand, in the limit $u_{\infty}\rightarrow 0$, there is no driving force to move the water film. In this case, $\delta_{0}=(2\nu_{a}x/u_{\infty})^{1/2}$ and $\bar{h}_{0}$ in Eq. (\ref{eq:h0-ula}) have an infinite value and hence the corresponding wavelength is not defined. The same issue arises for gravity driven water flows found in previous papers, \cite{Ogawa02, Ueno03, Ueno04, Ueno07, UFYT10, Ueno10} in which the thickness of water film is determined from $\bar{h}_{0}=[3\nu_{l}/(g\sin\theta)Q/l_{w}]^{1/3}$, where $\theta$ is the inclination angle with respect to the horizontal. In the limit $\theta \rightarrow 0$, there is no driving force to move the water film. Then the wavelength at $\theta=0$ is not defined (see Fig. 8 (a) in Ref. \onlinecite{UFYT10}).
\begin{table}[ht]
\caption{\label{tab:tableII}
Variation of
temperature at the water-air interface, $T_{la}$,
undisturbed ice growth rate, $\bar{V}$,
maximum value of dimensionless amplification rate $\sigma_{*\rm max}^{(r)}$ at a dimensionless wave number $k_{a*}$,
the corresponding wavelength, $\lambda$,
dimensionless phase velocity, $v_{p*}$,
dimensionless displacement of the ice-water interface, $\Delta x_{*}=v_{p*}/\sigma_{*\rm max}^{(r)}$
after the dimensionless time $t_{*}=1/\sigma_{*\rm max}^{(r)}$,
against the free stream velocity, $u_{\infty}$,
for $x=0.1$ m, $Q/l_{w}=1000$ $[{\rm (ml/h)/cm}]$ and $T_{\infty}=-10$ $^{\circ}$C.}
\begin{ruledtabular}
\begin{tabular}{cccccccc}
$u_{\infty}$ (m/s) & $T_{la}$ ($^{\circ}$C) & $\bar{V}$ ($\times 10^{-6}$ m/s)
& $k_{a*}$ &$\sigma_{*\rm max}^{(r)}$ & $\lambda$ (cm) & $v_{p*}$ & $\Delta x_{*}$\\
5 & -0.32 & 0.40 & 0.12 & 18.1 & 3.78 & -70.3 & -3.9 \\
10 & -0.27 & 0.57 & 0.17 & 23.9 & 1.88 & -97.9 & -4.1 \\
15 & -0.25 & 0.70 & 0.21 & 21.0 & 1.25 & -105.1 & -5.0 \\
20 & -0.23 & 0.81 & 0.22 & 17.7 & 1.03 & -96.6 & -5.4 \\
25 & -0.22 & 0.91 & 0.23 & 16.0 & 0.88 & -91.8 & -5.7 \\
30 & -0.21 & 1.00 & 0.23 & 15.2 & 0.80 & -80.1 & -5.3 \\
100 & -0.15 & 1.83 & 0.29 & 13.8 & 0.35 & -68.7 & -5.0 \\
\end{tabular}
\end{ruledtabular}
\end{table}
\subsection{\label{sec:Heat}Heat transfer at disturbed ice-water and water-air interfaces}
Using the assumed forms of $T'_{l}$ and $T'_{s}$ in Eq. (\ref{eq:pertset}) and considering their imaginary parts, the temperature in the water layer and ice can be expressed in dimensionless form as follows:
\begin{eqnarray}
T_{l*}(y_{*})\equiv \frac{T_{l}(y_{*})-T_{sl}}{T_{sl}-T_{la}}
&=& -y_{*}
+\delta_{b}(t_{*})\left\{H_{l}^{(r)}(y_{*})\sin[k_{l*}(x_{*}-v_{p*}t_{*})] \right.
\nonumber \\
&& \left. +H_{l}^{(i)}(y_{*})\cos[k_{l*}(x_{*}-v_{p*}t_{*})]\right\},
\label{eq:Tl}
\end{eqnarray}
\begin{eqnarray}
T_{s*}(y_{*}) \equiv \frac{T_{s}(y_{*})-T_{sl}}{T_{sl}-T_{la}}
&=& \delta_{b}(t_{*}){\rm exp}(k_{l*}y_{*})
\left\{(H_{l}^{(r)}|_{y_{*}=0}-1)\sin[k_{l*}(x_{*}-v_{p*}t_{*})] \right.
\nonumber \\
&& \left. +H_{l}^{(i)}|_{y_{*}=0}\cos[k_{l*}(x_{*}-v_{p*}t_{*})]\right\},
\label{eq:Ts}
\end{eqnarray}
where we used $\bar{T}_{s}=T_{sl}$ and the solution $H_{s}(y_{*})=(H_{l}|_{y_{*}=0}-1){\rm exp}(k_{l*}y_{*})$. \cite{Ueno03, Ueno04, Ueno07, UFYT10, Ueno10}
A microscopic energy balance have to be considered to explain fine details of ice morphology.
We define the disturbed parts of heat flux from the ice-water interface to the water, from the ice to the ice-water interface, and from the water-air interface to the air, as
$q'_{l}\equiv \Imag[-K_{l}\partial T'_{l}/\partial y|_{y=\zeta}]$,
$q'_{s}\equiv \Imag[-K_{s}\partial T'_{s}/\partial y|_{y=\zeta}]$ and
$q'_{a}\equiv \Imag[-K_{a}\partial T'_{a}/\partial y|_{y=\xi}]$, respectively.
These can be expressed in dimensionless form as follows:
\begin{eqnarray}
q'_{l*}\equiv
\frac{q'_{l}}{K_{l}\bar{G}_{l}}
&=&-\delta_{b}(t_{*})\left\{\frac{dH_{l}^{(r)}}{dy_{*}}\Big|_{y_{*}=0}\sin[k_{l*}(x_{*}-v_{p*}t_{*})] \right. \nonumber \\
&& \left. +\frac{dH_{l}^{(i)}}{dy_{*}}\Big|_{y_{*}=0}\cos[k_{l*}(x_{*}-v_{p*}t_{*})]\right\},
\label{eq:ql}
\end{eqnarray}
\begin{eqnarray}
q'_{s*}\equiv
\frac{q'_{s}}{K_{l}\bar{G}_{l}}
&=&-\delta_{b}(t_{*})K^{s}_{l}k_{l*}\left\{(H_{l}^{(r)}|_{y_{*}=0}-1)\sin[k_{l*}(x_{*}-v_{p*}t_{*})] \right. \nonumber \\
&& \left. +H_{l}^{(i)}|_{y_{*}=0}\cos[k_{l*}(x_{*}-v_{p*}t_{*})]\right\},
\label{eq:qs}
\end{eqnarray}
\begin{eqnarray}
q'_{a*}\equiv
\frac{q'_{a}}{K_{l}\bar{G}_{l}}
&=&-\delta_{b}(t_{*})\left\{\left(G'^{(r)}_{a}f_{l}^{(r)}|_{y_{*}=1}-G'^{(i)}_{a}f_{l}^{(i)}|_{y_{*}=1}\right)
\sin[k_{l*}(x_{*}-v_{p*}t_{*})] \right. \nonumber \\
&& \left. +\left(G'^{(r)}_{a}f_{l}^{(i)}|_{y_{*}=1}+G'^{(i)}_{a}f_{l}^{(r)}|_{y_{*}=1}\right)
\cos[k_{l*}(x_{*}-v_{p*}t_{*})]\right\},
\label{eq:qa}
\end{eqnarray}
where
$G'^{(r)}_{a}\equiv (\bar{h}_{0}/\delta_{0})(-dH_{a}^{(r)}/d\eta)|_{\eta=0}$ and
$G'^{(i)}_{a}\equiv (\bar{h}_{0}/\delta_{0})(-dH_{a}^{(i)}/d\eta)|_{\eta=0}$
represent the real and imaginary parts of the disturbed part of the air temperature gradient
$G'_{a}\equiv (\bar{h}_{0}/\delta_{0})(-dH_{a}/d\eta)|_{\eta=0}$ at the water-air interface.
From Eq. (\ref{eq:heatflux-zeta}), the disturbed part of the Stephan condition in dimensionless form
can be written as $\partial \zeta_{*}/\partial t_{*}=q'_{l*}-q'_{s*}$. Substituting Eqs. (\ref{eq:zeta}), (\ref{eq:ql}) and (\ref{eq:qs}) into this condition, Eqs. (\ref{eq:amp}) and (\ref{eq:phasevel}) are obtained.
\begin{figure}[ht]
\begin{center}
\includegraphics[width=8cm,height=8cm,keepaspectratio,clip]{zeta-xi-ua5-tamp0-tamp-x0p1-q1000.eps}\hspace{5mm}
\includegraphics[width=8cm,height=8cm,keepaspectratio,clip]{zeta-xi-ua20-tamp0-tamp-x0p1-q1000.eps}
\end{center}
\caption{For $Q/l_{w}=1000$ [(ml/h)/cm],
(a) and (b) are illustrations of the time evolution of an initial disturbance of the ice-water interface from $t_{*}=0$ to $t_{*}=1/\sigma^{(r)}_{*\rm max}$. The arrows indicate the position of maximum point of disturbed heat flux $q'_{l*}$, $q'_{s*}$ at the ice-water interface and that of $q'_{a*}$ at the water-air interface.
(a) represents the disturbance of $k_{a*}=0.12$ in the case of $u_{\infty}=5$ m/s.
(b) represents the disturbance of $k_{a*}=0.22$ in the case of $u_{\infty}=20$ m/s.
$\Delta x_{*}$ is the displacement of the ice-water interface after the time $t_{*}=1/\sigma^{(r)}_{*\rm max}$.
Vertical height is not to scale.}
\label{fig:heatflux-sla}
\end{figure}
Figures \ref{fig:heatflux-sla} (a) and (b) illustrate the time evolution of the ice-water interface disturbance with an initial amplitude of $\delta_{b}=0.1$ in the case of $u_{\infty}=5$ m/s and $u_{\infty}=20$ m/s, respectively.
The respective wave numbers of disturbance are $k_{a*}=0.12$ in Fig. \ref{fig:heatflux-sla} (a) and $k_{a*}=0.22$ in Fig. \ref{fig:heatflux-sla} (b). These are the fastest growing modes, at which $\sigma^{(r)}_{*}$ acquires a maximum value, as shown by the solid curves in Fig. \ref{fig:ka-amp-lambda} (a).
The phase shift of the water-air interface relative to the ice-water interface in Fig. \ref{fig:heatflux-sla} (b) is negligibly small compared to that in Fig. \ref{fig:heatflux-sla} (a), as shown by the solid curve in Fig. \ref{fig:ka-xi-Thetaxi} (b).
Due to the left-to right air and water flows indicated by arrows in Fig. \ref{fig:heatflux-sla}, the isotherms in the air and water boundary layers are no longer symmetrical around each protruded part of the water-air and ice-water interfaces. Since the isotherms become closer on the upstream side of each protruded part of the interfaces, $q'_{a*}$, $q'_{l*}$ and $q'_{s*}$ are largest on the upstream side of each protruded part, as indicated by the vertical arrows in Fig. \ref{fig:heatflux-sla}. Hence, the ice growth rate on the upstream side of each protruded part is faster than that on the downstream side, and this results in the translation of the ice-water interface in the direction opposite to the water flow. As mentioned in \ref{sec:amplification}, the displacements after the time $t_{*}=1/\sigma^{(r)}_{*\rm max}$ are $\Delta x_{*}=-3.9$ and $\Delta x_{*}=-5.4$ in Figs. \ref{fig:heatflux-sla} (a) and (b), respectively.
We separate the local heat transfer coefficient at the water-air interface, $h_{x}$,
into the undisturbed part
$\bar{h}_{x}=-K_{a}\partial\bar{T}_{a}/\partial y|_{y=\bar{h}_{0}}/(T_{la}-T_{\infty})$ and
the disturbed part
$h'_{x}=\Imag[-K_{a}\partial T'_{a}/\partial y|_{y=\bar{h}_{0}}/(T_{la}-T_{\infty})]$.
The former can be written as
$\bar{h}_{x}=K_{a}/\delta_{0}(-d\bar{T}_{a*}/d\eta|_{\eta=0})$.
Using the value of $\bar{G}_{a*}=-d\bar{T}_{a*}/d\eta|_{\eta=0}=0.413$ obtained in \ref{sec:LSA},
the value of local Nusselt number scaled by $\sqrt{Re_{ax}}$ is
$\bar{Nu}_{x}/\sqrt{Re_{ax}}
=(\bar{h}_{x}x/K_{a})/\sqrt{u_{\infty}x/\nu_{a}}
=-(1/\sqrt{2})d\bar{T}_{a*}/d\eta|_{\eta=0}=0.292$,
from which we obtain $\bar{h}_{x}=0.292K_{a}\sqrt{u_{\infty}/(\nu_{a}x})$.
A similar expression for the laminar convective heat transfer coefficient,
$\bar{h}_{s}=0.296(K_{a}/\sqrt{\nu_{a}})\sqrt{V_{e}^{2.87}/(\int_{0}^{s}V_{e}^{1.87}ds)}$,
is used in aircraft icing models, \cite{Gent00, Myers04}
where $s$ is the surface distance from the stagnation point
and $V_{e}$ is the velocity at edge of air boundary layer.
When $V_{e}=u_{\infty}$ (constant) and replacing $s$ with $x$, $\bar{h}_{s}$ yields the same expression as $\bar{h}_{x}$ except for the very slight difference of numerical factor.
Using $\bar{h}_{x}$, the undisturbed ice growth rate $\bar{V}$ in Eq. (\ref{eq:V}) can be expressed as $\bar{V}=-\bar{h}_{x}T_{\infty}/L$.
On the other hand, the disturbed part of the heat transfer coefficient normalized by the undisturbed one can be written as
\begin{eqnarray}
h'_{x}/\bar{h}_{x}
&=&-\Imag\left[G'_{a}f_{l}|_{y_{*}=1}\delta_{b}{\rm exp}(\sigma_{*}t_{*}+ik_{l*}x_{*})\right] \nonumber \\
&=&\delta_{b}(t_{*})\left[(h'_{x}/\bar{h}_{x})^{(r)}\sin[k_{l*}(x_{*}-v_{p*}t_{*})\right]
+(h'_{x}/\bar{h}_{x})^{(i)}\cos[k_{l*}(x_{*}-v_{p*}t_{*})]]\nonumber \\
&=&\delta_{b}(t_{*})|h'_{x}/\bar{h}_{x}|\sin[k_{l*}(x_{*}-v_{p*}t_{*})-\Theta_{q'_{a*}}].
\label{eq:h'x}
\end{eqnarray}
Equation (\ref{eq:h'x}) becomes the same form as Eq. (\ref{eq:qa}) by putting
$(h'_{x}/\bar{h}_{x})^{(r)}=-(G'^{(r)}_{a}f_{l}^{(r)}|_{y_{*}=1}-G'^{(i)}_{a}f_{l}^{(i)}|_{y_{*}=1})$ and
$(h'_{x}/\bar{h}_{x})^{(i)}=-(G'^{(r)}_{a}f_{l}^{(i)}|_{y_{*}=1}+G'^{(i)}_{a}f_{l}^{(r)}|_{y_{*}=1})$.
Here
\begin{equation}
|h'_{x}/\bar{h}_{x}|=[\{(h'_{x}/\bar{h}_{x})^{(r)}\}^{2}+\{(h'_{x}/\bar{h}_{x})^{(i)}\}^{2}]^{1/2}
\label{eq:amplitude-h'x}
\end{equation}
is the amplitude, and
$\cos\Theta_{q'_{a*}}=(h'_{x}/\bar{h}_{x})^{(r)}/|h'_{x}/\bar{h}_{x}|$ and
$\sin\Theta_{q'_{a*}}=-(h'_{x}/\bar{h}_{x})^{(i)}/|h'_{x}/\bar{h}_{x}|$,
$\Theta_{q'_{a*}}$ is a phase difference between $q'_{a*}=h'_{x}/\bar{h}_{x}$ and the ice-water interface.
\begin{figure}[ht]
\begin{center}
\includegraphics[width=7cm,height=7cm,keepaspectratio,clip]{eachhc-uinf.eps}\hspace{5mm}
\includegraphics[width=7cm,height=7cm,keepaspectratio,clip]{abshc-uinf.eps}\\[5mm]
\includegraphics[width=8cm,height=8cm,keepaspectratio,clip]{har-hai-q1000-x0p1.eps}
\end{center}
\caption{For $Q/l_{w}=1000$ [(ml/h)/cm] and $x=0.1$ m,
(a) represents the variation of the disturbed part of the heat transfer coefficient normalized by the undisturbed heat transfer coefficient against the free stream velocity $u_{\infty}$: $(h'_{x}/\bar{h}_{x})^{(r)}$ is real part and $(h'_{x}/\bar{h}_{x})^{(i)}$ is imaginary part.
(b) represents the variation of $|h'_{x}/\bar{h}_{x}|$ against $u_{\infty}$. The solid curves consider the effect of the tangential and normal air shear stress disturbances on the water-air interface, and the dashed curves do not consider this effect.
(c) represents the variation of $(h'_{x}/\bar{h}_{x})^{(r)}$ (solid curves) and $(h'_{x}/\bar{h}_{x})^{(i)}$ (dashed curves) against dimensionless wave number $k_{a*}$ for free stream velocities $u_{\infty}$=5, 10, 20 m/s.
}
\label{fig:heat-coefficient}
\end{figure}
The solid curves in Fig. \ref{fig:heat-coefficient} (a) show the variations of $(h'_{x}/\bar{h}_{x})^{(r)}$ and $(h'_{x}/\bar{h}_{x})^{(i)}$ against $u_{\infty}$, taking into account the air shear stress disturbances. These values are estimated for $Q/l_{w}=1000$ [(ml/h)/cm] and $x=0.1$ m, and for $k_{a*}$ at which $\sigma_{*}^{(r)}$ acquires a maximum value. It should be noted that $q'_{a*}=h'_{x}/\bar{h}_{x}$ includes $G'_{a}$ and $f_{l}|_{y_{*}=1}$. $\bar{h}_{x}$ depends on only two parameters, free stream velocity $u_{\infty}$ and position $x$. On the other hand, $h'_{x}$ depends on many parameters.
$G'_{a}=(\bar{h}_{0}/\delta_{0})(-dH_{a}/d\eta)|_{\eta=0}$ is determined from the disturbed airflow and temperature fields, and $f_{l}|_{y_{*}=1}$ determines the magnitude of amplitude and phase of the water-air interface
by using the relation $\xi_{k}=-f_{l}|_{y_{*}=1}\zeta_{k}$. The shape of the water-air interface changes by the action of gravity, surface tension and air shear stress disturbances.
As shown in Fig. \ref{fig:ka-xi-Thetaxi}, if we neglect the effect of the air shear stress disturbances, the amplitude and phase of the water-air interface relative to the ice-water interface are not correctly evaluated as $u_{\infty}$ increases. This results in an overestimated value of $|h'_{x}/\bar{h}_{x}|$
compared to that taking into account the effect of air shear stress disturbances with increasing $u_{\infty}$, as shown by the dashed and solid curves in Fig. \ref{fig:heat-coefficient} (b).
The solid curves in Fig. \ref{fig:heat-coefficient} (a) shows that $(h'_{x}/\bar{h}_{x})^{(i)}$ is positive for any $u_{\infty}$, while $(h'_{x}/\bar{h}_{x})^{(r)}$ is negative when $u_{\infty}<10$ m/s and is positive when $u_{\infty}>10$ m/s.
From $\cos\Theta_{q'_{a*}}=(h'_{x}/\bar{h}_{x})^{(r)}/|h'_{x}/\bar{h}_{x}|$ and
$\sin\Theta_{q'_{a*}}=-(h'_{x}/\bar{h}_{x})^{(i)}/|h'_{x}/\bar{h}_{x}|$, the corresponding phase difference between $q'_{a*}=h'_{x}/\bar{h}_{x}$ and the ice-water interface is
$-\pi<\Theta_{q'_{a*}}<-\pi/2$ and $-\pi/2<\Theta_{q'_{a*}}<0$, respectively.
On the other hand, if we neglect the air shear stress disturbances, as shown by the dashed curves in Fig. \ref{fig:heat-coefficient} (a), $(h'_{x}/\bar{h}_{x})^{(r)}$ is negative and $(h'_{x}/\bar{h}_{x})^{(i)}$ is positive for any $u_{\infty}$.
Then, the phase difference between $q'_{a*}=h'_{x}/\bar{h}_{x}$ and the ice-water interface is always $-\pi<\Theta_{q'_{a*}}<-\pi/2$
with respect to any $u_{\infty}$. This means that the position of maximum point of heat flux $q'_{a*}$ or that of the heat transfer coefficient $h'_{x}/\bar{h}_{x}$ depend significantly on the air shear stress disturbances.
Figure \ref{fig:heat-coefficient} (c) shows the variation of $(h'_{x}/\bar{h}_{x})^{(r)}$ (solid curves) and $(h'_{x}/\bar{h}_{x})^{(i)}$ (dashed curves) against $k_{a*}$ for the free stream velocities of $u_{\infty}$=5, 10, 20 m/s.
In the case of $u_{\infty}=5$ m/s, $(h'_{x}/\bar{h}_{x})^{(r)}$ is negative and $(h'_{x}/\bar{h}_{x})^{(i)}$ is positive for any $k_{a*}$. From $\cos\Theta_{q'_{a*}}=(h'_{x}/\bar{h}_{x})^{(r)}/|h'_{x}/\bar{h}_{x}|$ and
$\sin\Theta_{q'_{a*}}=-(h'_{x}/\bar{h}_{x})^{(i)}/|h'_{x}/\bar{h}_{x}|$, the phase difference between $q'_{a*}=h'_{x}/\bar{h}_{x}$ and the ice-water interface is
$-\pi<\Theta_{q'_{a*}}<-\pi/2$
as shown in Fig. \ref{fig:heatflux-sla} (a).
On the other hand, in the case of $u_{\infty}$= 10 and 20 m/s, $(h'_{x}/\bar{h}_{x})^{(i)}$ is positive for any $k_{a*}$, but $(h'_{x}/\bar{h}_{x})^{(r)}$ changes from positive to negative at $k_{a*}=0.17$ and $k_{a*}=0.33$, respectively.
Then, in the case of $u_{\infty}=10$ m/s, the phase difference is
$-\pi/2<\Theta_{q'_{a*}}<0$ for $k_{a*}<0.17$ and
$-\pi<\Theta_{q'_{a*}}<-\pi/2$ for $k_{a*}>0.17$.
Likewise, in the case of $u_{\infty}=20$ m/s, the phase difference is
$-\pi/2<\Theta_{q'_{a*}}<0$ for $k_{a*}<0.33$ and
$-\pi<\Theta_{q'_{a*}}<-\pi/2$ for $k_{a*}>0.33$,
as shown in Fig. \ref{fig:heatflux-sla} (b). This means that the position of maximum point of $q'_{a*}=h'_{x}/\bar{h}_{x}$ on the water-air interface changes according to the wavelength of the ice-water interface disturbance and the free stream velocity.
Furthermore, as shown in Fig. \ref{fig:heatflux-sla}, the position of the maximum point of $q'_{a*}=h'_{x}/\bar{h}_{x}$ moves in the direction opposite to the water flow with time.
\section{Summary and Discussion}
We have proposed a theoretical model for ice growth under a supercooled water film driven by wind drag. The thickness and surface velocity of the water layer are variable by changing air stream velocity and water supply rate.
For a given water supply rate, we investigated the morphological instability of the ice-water interface for various air stream velocities using a linear stability analysis, taking into account the effect of gravity, surface tension and the tangential and normal air shear stress disturbances due to the airflow on the shape of the water-air interface.
Even for the simple model developed here, the form of heat transfer coefficient at the disturbed water-air interface is too complicated, which depends on the disturbed air flow and temperature fields, the shape of the disturbed water-air interface, as well as the shape of the ice-water interface. By considering the interaction between the air and water flows, we have found that the heat transfer coefficient at the water-air interface is significantly affected by the air shear stress disturbances, which suppresses the dimensionless amplification rate of the ice-water interface disturbance as the air stream velocity increases. However, the air shear stress disturbances do not significantly change the wavelength of an ice pattern occurring as a result of morphological instability of the ice-water interface. The model herein predicts that a centimeter scale ice pattern will appear, and its wavelength will decrease with increasing air stream velocity. Moreover, the ice pattern will translate towards the water source with time.
At higher airspeed, the theoretical predictions obtained here might be relevant to the experiments for surface roughness characteristics associated with leading edge ice accretion on a NACA 0012 airfoil at a 0-deg angle of attack. \cite{Shin96} In that experiment, the height and spacing of roughness elements were measured with various icing parameters in glaze icing conditions. It was observed that the roughness spacing is about 1 mm, and that smooth-to-rough zones move upstream towards the stagnation region with time.
Here, first we mention some differences between previous wet icing models \cite{Bourgault00, Myers02_1, Myers02_2, Myers04} and the current model:
(1) The undisturbed part of water film velocity profile derived herein, $\bar{u}_{l*}=y_{*}$, is the same as that used in the shallow-water icing model. \cite{Bourgault00} However, the disturbed part of water flow due to the disturbance of the ice-water interface is taken into account herein.
(2) The current undisturbed part of temperature in the water film has a linear profile, as used in the models. \cite{Myers02_1, Myers02_2, Myers04} However, in this model,
the disturbed part $T'_{l}$ due to the disturbance of the ice-water interface is considered, as shown in Eq. (\ref{eq:Tl}).
Since the Peclet number herein is large, the disturbed part of temperature distribution in the water film is affected by the advection due to $\bar{u}_{l*}$ and $v'_{l}$, as indicated in the terms with $\Pec_{l}$ in Eq. (\ref{eq:geq-Hl}).
(3) In previous icing models, detailed calculations concerning the effect of the interaction between the air and water flows on the temperature distribution in the water film were not carried out. The air shear stress disturbances influence the disturbed part of stream function of the water flow, $f_{l}$. As a result,
the disturbed part of temperature distribution in the water film, $H_{l}$, is affected by both air and water flows.
If we neglect the disturbed part of temperature distribution in the water film flow
and focus on only the influence of the temperature distribution in the air on the growth condition of the ice-water interface disturbance, Eq. (\ref{eq:heatflux-zeta}) may be replaced by
$L(\bar{V}+\partial \zeta/\partial t)
=-K_{a}\partial T_{a}/\partial y|_{y=\xi}$.
Linearizing this equation at $y=\bar{h}_{0}$, the zeroth order yields
$\bar{V}=-K_{a}T_{\infty}/(L\delta_{0}/\bar{G}_{a*})$,
which is identical to Eq. (\ref{eq:V}).
It is found that since the undisturbed part of ice growth rate, $\bar{V}$, does not include any parameter associated with the water film, $\bar{V}$ is determined without considering the details of the water film,
and the heat transfer through the air boundary layer is the deciding factor in $\bar{V}$.
From the first order in $\xi_{k}$, $\partial \zeta_{*}/\partial t_{*}=q'_{a*}=h'_{x}/\bar{h}_{x}$ yields
$\sigma_{*}^{(r)}=(h'_{x}/\bar{h}_{x})^{(r)}$ and
$v_{p*}=-\sigma_{*}^{(i)}/k_{l*}=-(h'_{x}/\bar{h}_{x})^{(i)}/k_{l*}$.
When we neglect the details of the water film, $(h'_{x}/\bar{h}_{x})^{(r)}$ (solid curves) in Fig. \ref{fig:heat-coefficient} (c) directly represents the amplification rate.
However, there is a significant difference between $\sigma_{*}^{(r)}$ in Fig. \ref{fig:ka-amp-lambda} (a) and that in Fig. \ref{fig:heat-coefficient} (c).
$\sigma_{*}^{(r)}$ in Fig. \ref{fig:ka-amp-lambda} (a) takes into account the effect of the disturbed part of temperature distribution in the water layer as well as that in the air boundary layer on the ice growth conditions.
On the other hand, $\sigma_{*}^{(r)}=(h'_{x}/\bar{h}_{x})^{(r)}$ in Fig. \ref{fig:heat-coefficient} (c) is obtained without considering the details of the disturbed temperature distribution in the water film.
This suggests that the wavelength of ice pattern and its translation velocities shown in Table \ref{tab:tableII} cannot be evaluated correctly if we neglect the details of the water film.
It should be emphasized that the same issue arose in a previous paper; \cite{Ueno10}
If we neglected the influence of the disturbed temperature distribution in the water film flow on the growth condition for the ice-water interface disturbance, the amplification rate $\sigma_{*}^{(r)}$ had positive values for all wave numbers and hence a characteristic wavelength of icicle ripples was not obtained.
A centimeter scale of ripples in wavelength was obtained from only $\sigma_{*}^{(r)}$ being taken into account the disturbed temperature distribution in the water film. \cite{Ueno03, Ueno04, Ueno07, UFYT10, Ueno10}
Second, we mention the temperature at the ice-water interface, $T_{i}$, in Eq. (\ref{eq:Tsl}).
Within linear stability analysis, $T_{i}=T_{sl}+\Delta T_{sl}$,
where $T_{sl}$ is the temperature at an undisturbed ice-water interface
and $\Delta T_{sl}$ is a deviation from it when the ice-water interface is disturbed.
Its dimensionless form $\Delta T_{sl*}\equiv\Imag[\Delta T_{sl}/(T_{sl}-T_{la})]$ can be expressed as follows by evaluating Eq. (\ref{eq:Tl}) at the disturbed ice-water interface $y_{*}=\zeta_{*}$:
\begin{equation}
\Delta T_{sl*}
= \delta_{b}(t_{*})\left\{(H_{l}^{(r)}|_{y_{*}=0}-1)\sin[k_{l*}(x_{*}-v_{p*}t_{*})]
+H_{l}^{(i)}|_{y_{*}=0}\cos[k_{l*}(x_{*}-v_{p*}t_{*})]\right\}.
\label{eq:DeltaTsl}
\end{equation}
Figure \ref{fig:isotherms} (a) represents the isotherms in the water film. The real and imaginary parts of the disturbed part of temperature distribution in the water film, $H_{l}^{(r)}$ and $H_{l}^{(i)}$ in Eq. (\ref{eq:Tl}), are determined by solving Eq. (\ref{eq:geq-Hl}) subject to the boundary conditions (\ref{eq:Tla-xi}) and (\ref{eq:heatflux-xi-h0}), which were derived from the continuity of temperature and heat flux at the water-air interface, Eqs. (\ref{eq:bc-Tla}) and (\ref{eq:heatflux-xi}), respectively.
Since the temperature distribution in the water film is affected by both air and water flows, $\Delta T_{sl*}$ varies.
When the ice-water interface is flat, $\Delta T_{sl*}$ must be zero. Indeed, $H_{l}^{(r)}|_{y_{*}=0} \rightarrow 1$ and $H_{l}^{(i)}|_{y_{*}=0} \rightarrow 0$ in the limit $k_{l*} \rightarrow 0$ were numerically confirmed.
\begin{figure}[ht]
\begin{center}
\includegraphics[width=7cm,height=7cm,keepaspectratio,clip]{Tl-Ts-isotherm-ua20-theta0-q1000.eps}\hspace{5mm}
\includegraphics[width=7cm,height=7cm,keepaspectratio,clip]{qs-ql-qa-uinf.eps}\hspace{5mm}
\includegraphics[width=7cm,height=7cm,keepaspectratio,clip]{dbc-Tl-isotherm-ua20-theta0-q1000.eps}
\includegraphics[width=7cm,height=7cm,keepaspectratio,clip]{dbc-ka-amp-x0p1-q1000.eps}
\end{center}
\caption{For $Q/l_{w}=1000$ [(ml/h)/cm], $u_{\infty}=20$ m/s and $\delta_{b}(t_{*})=0.05$,
(a) represents the isotherms in the water film and (b) the isotherms in the ice, for the boundary condition
$T_{l}|_{y=\xi}=T_{a}|_{y=\xi}=T_{la}$ (constant).
(c) represents the variation of disturbed heat flux $|q'_{l*}|$, $|q'_{s*}|$ and $|q'_{a*}|$ against the free stream velocity $u_{\infty}$.
For the boundary condition $T_{l}|_{y=\zeta}=T_{s}|_{y=\zeta}=T_{sl}$ (constant),
(d) represents the isotherms in the water film and
(e) dimensionless amplification rate $\sigma_{*}^{(r)}$ versus dimensionless wave number $k_{a*}$.
The numbers in the isotherms are the values of dimensionless temperatures in water (\ref{eq:Tl}) and in ice (\ref{eq:Ts}).
}
\label{fig:isotherms}
\end{figure}
From Eqs. (\ref{eq:ql}), (\ref{eq:qs}) and (\ref{eq:qa}), we define the magnitude of $q'_{l*}$, $q'_{s*}$ and $q'_{a*}$ as follows:
\begin{eqnarray}
|q'_{l*}|&\equiv & \{(dH_{l}^{(r)}/dy_{*}|_{y_{*}=0})^{2}+(dH_{l}^{(i)}/dy_{*}|_{y_{*}=0})^{2}\}^{1/2}, \\
|q'_{s*}|&\equiv & K^{s}_{l}k_{l*}\{(H_{l}^{(r)}|_{y_{*}=0}-1)^{2}+(H_{l}^{(i)}|_{y_{*}=0})^{2}\}^{1/2}, \\
|q'_{a*}|&\equiv &\{(G'^{(r)}_{a}f_{l}^{(r)}|_{y_{*}=1}-G'^{(i)}_{a}f_{l}^{(i)}|_{y_{*}=1})^{2}
+(G'^{(r)}_{a}f_{l}^{(i)}|_{y_{*}=1}+G'^{(i)}_{a}f_{l}^{(r)}|_{y_{*}=1})^{2}\}^{1/2}.
\label{eq:amplitude-ql-qa-qs}
\end{eqnarray}
Since the water flow is affected by the air flow through the air shear stress disturbances, $|q'_{l*}|$ and $|q'_{s*}|$ as well as $|q'_{a*}|$ depend on the free stream velocity $u_{\infty}$.
Figure \ref{fig:isotherms} (c) represents the variation of $|q'_{l*}|$, $|q'_{s*}|$ and $|q'_{a*}|$ against $u_{\infty}$. These values were estimated for $Q/l_{w}=1000$ [(ml/h)/cm] and $x=0.1$ m, and for $k_{a*}$ at which $\sigma_{*}^{(r)}$ acquires a maximum value for a given $u_{\infty}$.
From Eq. (\ref{eq:heatflux-zeta}), the disturbed part of the Stephan condition is $\partial \zeta_{*}/\partial t_{*}=q'_{l*}-q'_{s*}$. Therefore, the net heat flux $q'_{l*}-q'_{s*}$ determines the amplification rate of the ice-water interface disturbance.
Since $|q'_{a*}|=|h'_{x}/\bar{h}_{x}|$ from Eq. (\ref{eq:amplitude-h'x}), the dashed-dotted curve in Fig. \ref{fig:isotherms} (c) is the same as the solid curve in Fig. \ref{fig:heat-coefficient} (b).
When both ice-water and water-air interfaces are flat and the undisturbed temperature in the water film is the linear profile $\bar{T}_{l*}=y_{*}$, the latent heat released at the ice-water interface, $L\bar{V}$, the undisturbed part of heat flux at the ice-water interface, $K_{l}\bar{G}_{l}$ and that at the water-air interface, $K_{a}\bar{G}_{a}$,
must be equal.
Hence all of the latent heat is released away from the water-air interface through the water film.
However, in the case of the disturbed interfaces, the disturbed part of heat flux at the ice-water
interface, $|q'_{l*}|$, is not necessarily equal to that at the water-air interface, $|q'_{a*}|$. Indeed, as shown in Fig. \ref{fig:isotherms} (c), $|q'_{a*}|$ is much smaller than $|q'_{l*}|$.
The release of latent heat at the water-air interface, $|q'_{a*}|$, is limited not only by the temperature gradient at the water-air interface but also by the shape of the water-air interface. Hence all of the latent heat released at the disturbed ice-water interface cannot be removed at the water-air interface and most of it is carried by the flow in the water film.
The effect of morphological instabilities is to increase the surface area of the phase boundary and hence to enhance the release of latent heat. The latent heat that is not completely removed by air and water flows may lead to a local temperature rise in the supercooled water and ice locally. Also, the flow in the water film can carries a supercooled water in the interior towards the ice-water interface. Hence, in Fig. \ref{fig:isotherms} (a), not only the isotherms in the water film are deformed by the advection terms in Eq. (\ref{eq:geq-Hl}) but also alternating patterns of warming and cooling appear in the neighbourhood of the ice-water interface.
The characteristic time of the deformation associated with the shear rate is $1/(u_{la}/\bar{h}_{0})$.
On the other hand, the thermal diffusion time associated with a wave number $k_{l}$ is $1/(\kappa_{l}k_{l}^{2})$.
For a disturbed ice-water interface with a 1 cm wavelength shown in Figs. \ref{fig:isotherms} (a) and (d),
the condition $1/(u_{la}/\bar{h}_{0}) \ll 1/(\kappa_{l}k_{l}^{2})$ is satisfied.
Hence, the temperature distribution in the neighbourhood of the disturbed ice-water interface is deformed by the advection in the water film.
In order to avoid a temperature discontinuity, $\Delta T_{sl*}$, between water and ice, the disturbed heat flux $q'_{s*}$ due to thermal diffusion occurs in the ice.
As a result, Fig. \ref{fig:isotherms} (b) also shows alternating patterns of warming and cooling in the ice.
From the comparison of Eqs. (\ref{eq:DeltaTsl}) and (\ref{eq:qs}) as well as that of Figs. \ref{fig:isotherms} (b) and \ref{fig:heatflux-sla} (b), if $\Delta T_{sl*}>0$ (warming), then $q'_{s*}<0$, hence the direction of $q'_{s*}$ is from the ice-water interface to the ice. On the other hand, if $\Delta T_{sl*}<0$ (cooling), then $q'_{s*}>0$, hence the direction of $q'_{s*}$ is from the ice to the ice-water interface.
However, it should be noted that this disturbed part of heat flux in the ice exists only in the vicinity of the ice-water interface, as Eq. (\ref{eq:Ts}) shows that the disturbed temperature in the ice is exponentially attenuated, and the ice temperature approaches $T_{s}=T_{sl}$ ($T_{sl}=$0 $^{\circ}$C for pure water) far from the ice-water interface, as shown in Fig. \ref{fig:isotherms} (b).
On the other hand, the isotherms in Fig. \ref{fig:isotherms} (d) are determined from different solutions $H_{l}^{(r)}$ and $H_{l}^{(i)}$, which are obtained by solving Eq. (\ref{eq:geq-Hl}) subject to the boundary conditions $H_{l}|_{y_{*}=0}=1$ and Eq. (\ref{eq:heatflux-xi-h0}). $H_{l}|_{y_{*}=0}=1$ is derived from the condition $T_{l}|_{y=\zeta}=T_{s}|_{y=\zeta}=T_{sl}$ (constant). In this case, the temperature at the water-air interface is an unknown to be determined, which is deviated from $T_{la}$ in Eq. (\ref{eq:bc-Tla})
and hence the first boundary condition in Eq. (\ref{eq:bc-Ha}) is replaced by
\begin{equation}
H_{a}|_{\eta=0}=1-\frac{K_{a}}{K_{l}}-\frac{K_{a}}{K_{l}}H_{l}|_{y_{*}=1}/f_{l}|_{y_{*}=1}.
\label{eq:DeltaTla}
\end{equation}
However, since $K_{a}/K_{l} \ll 1$, $H_{a}|_{\eta=0} \approx 1$ is a good approximation, which is equivalent to the condition $T_{l}|_{y=\xi}=T_{a}|_{y=\xi}\approx T_{la}$ (constant).
As shown in Fig. \ref{fig:isotherms} (d), the isotherms are slightly deformed from the undisturbed temperature distribution $\bar{T}_{l*}=y_{*}$ and the variation of temperature in the neighbourhood of the ice-water and water-air interfaces are strongly affected by the boundary temperatures, $T_{l*}|_{y_{*}=\zeta_{*}}=T_{s*}|_{y_{*}=\zeta_{*}}=0.0$ (constant) and $T_{l*}|_{y_{*}=\xi_{*}}=T_{a*}|_{y_{*}=\xi_{*}}\approx -1.0$ (constant). In particular, the isotherms in the neighbourhood of the ice-water interface in Fig. \ref{fig:isotherms} (d) are significantly different from those in Fig. \ref{fig:isotherms} (a). Although there exists a shear flow in the water film, the temperature distribution is almost symmetric around any protruded part of the ice-water interface. The long arrows at the ice-water interface in Fig. \ref{fig:isotherms} (d) show that the temperature gradient is largest at each protruded part of the ice-water interface. This promotes the ice growth at the protruded part than at the depressed part, always resulting in an unstable growth of the ice-water interface.
Consequently, the amplification rate $\sigma^{(r)}_{*}$ obtained from the boundary condition $T_{l}|_{y=\zeta}=T_{s}|_{y=\zeta}=T_{sl}$ (constant) has positive values for all wave numbers, as shown in Fig. \ref{fig:isotherms} (e).
From $1/(u_{la}/\bar{h}_{0})=1/(\kappa_{l}k_{l}^{2})$, the wave number at which the two time scale equals is given by $k_{l*}=\sqrt{\Pec_{l}}\sim 20$ for $Q/l_{w}=1000$ [(ml/h)/cm]. The corresponding wavelength is 150 $\mu$m from $k_{l*}=k\bar{h}_{0}=20$ for $u_{\infty}=20$ m/s.
The effect of the water flow on the isotherms in such a microscopic length scale is negligible. Instead, taking into account the Gibbs-Thomson effect, the temperature at the ice-water interface is expressed as
$T_{i}=T_{sl}+(T_{sl}\Gamma/L)\partial^{2} \zeta/\partial x^{2}$, from which $H_{l}|_{y_{*}=0}=1-(d_{0}/\bar{h}_{0})(l_{\rm th}/\bar{h}_{0})k_{l*}^{2}$ is derived.
Here $l_{\rm th}=\kappa_{l}/\bar{V}$ and $d_{0}=T_{sl}\Gamma C_{pl}/L^{2}$ are a macroscopic and microscopic characteristic length, respectively, $\Gamma$ is the ice-water interface tension and $C_{pl}$ is the specific heat at constant pressure of the water. Neglecting the advection terms in Eq. (\ref{eq:geq-Hl}) and solving it subject to the boundary conditions $H_{l}|_{y_{*}=0}=1-(d_{0}/\bar{h}_{0})(l_{\rm th}/\bar{h}_{0})k_{l*}^{2}$ and Eq. (\ref{eq:heatflux-xi-h0}) with $f_{l*}|_{y_{*}=1} \approx 0$, from Eq. (\ref{eq:amp}) we obtain the amplification rate,
$\sigma^{(r)}_{*}=k_{l*}\{1-(d_{0}/\bar{h}_{0})(l_{\rm th}/\bar{h}_{0})k_{l*}^{2}(1+K^{s}_{l})\}$,
which is just the result of the Mullins-Sekerka instability. \cite{Langer80}
Here the condition $f_{l*}|_{y_{*}=1} \approx 0$ means that the water-air interface is nearly flat because the surface tension is so dominant in the microscopic length scale of dendrite that the effect of disturbance concerning the dendrite spacing on the shape of the water-air interface is negligible.
$\sigma^{(r)}_{*}$ acquires a maximum value at
$k_{l*}=[\bar{h}_{0}^{2}/\{3l_{\rm th}d_{0}(1+K^{s}_{l})\}]^{1/2} \sim 10$
for $\bar{h}_{0} \sim 5 \times 10^{-4}$ m, $l_{\rm th} \sim 10^{-1}$ m, $d_{0} \sim 10^{-9}$ m and $K^{s}_{l}=3.92$.
The dependence of the microscopic wavelength predicted from the Mullins-Sekerka instability on the free stream velocity $u_{\infty}$ is
$\lambda_{\rm micro}=2\pi\{3l_{\rm th}d_{0}(1+K^{s}_{l})\}^{1/2} \propto u_{\infty}^{-1/4}$
because $l_{\rm th} \sim \delta_{0} \sim u_{\infty}^{-1/2}$ in the model herein. This result is contrast to the dependence of the macroscopic wavelength $\lambda_{\rm macro}$ on $u_{\infty}$. Figure \ref{fig:ka-amp-lambda} (b) shows that $\lambda_{\rm macro} \propto u_{\infty}^{-0.88}$.
It should be noted that the mechanism of the macro-scale morphological instability under a supercooled liquid film herein is quite different from the dendritic growth. $\lambda_{\rm macro}$ depends on the water film thickness $\bar{h}_{0}$ as indicated in Tables \ref{tab:tableI} and \ref{tab:tableII}, while $\lambda_{\rm micro}$ does not depend on it.
The use of the boundary condition $T_{l}|_{y=\zeta}=T_{s}|_{y=\zeta}=T_{sl}$ (constant) caused a serious problem in a model for the icicle ripple formation. \cite{UFYT10} This condition was used in the model proposed by Ogawa and Furukawa \cite{Ogawa02} when determining $H_{l}$. However, our numerical analysis did not reproduce their amplification rate shown in Fig. 4 of Ref. \onlinecite{Ogawa02}. Instead, the numerically obtained amplification rate had positive values for all wave numbers and there was no well-defined maximum amplification rate (for details, see Figure 5 (c) in Ref. \onlinecite{UFYT10}).
The same issue has already arisen in aircraft icing problems. Tsao and Rothmayer developed a physical model to describe the aero-hydro-thermo-dynamic interaction of a cold air boundary layer with glaze ice sheets and water films. \cite{Tsao98} However, their stability analysis showed that the ice-water interface disturbance became unstable for all modes (see Figure 7 in Ref. \onlinecite{Tsao98}), which indicated that there is no dominant amplification rate to select a preferred wavelength.
The assumption that the disturbed ice-water interface is at the equilibrium freezing temperature was used in their model too.
To overcome this issue, the Gibbs-Thomson effect was introduced to stabilize the smallest scale icing disturbances. \cite{Tsao00} However, the length scale predicted by their theory was much smaller than the ice roughness spacing of the order of millimeters observed in the experiments by Shin. \cite{Shin96}
On the other hand, the condition $T_{l}|_{y=\zeta}=T_{s}|_{y=\zeta}=T_{sl}$ (constant) was not used in the model for the icicle ripple formation proposed by Ueno \cite{Ueno03, Ueno04, Ueno07, UFYT10, Ueno10} when determining $H_{l}$. That model was able to predict a centimeter-scale wavelength of icicle ripples and upward ripple migration due to an asymmetry in the temperature distribution between the upstream and downstream side of any protruded part of the ice-water interface, which were confirmed by the experiments. \cite{UFYT10, Chen10}
\begin{figure}[ht]
\begin{center}
\includegraphics[width=10cm,height=10cm,keepaspectratio,clip]{aluminum-configuration.eps}\\[5mm]
\includegraphics[width=8cm,height=8cm,keepaspectratio,clip]{Gsl-ka-amp-uinf20-q1000.eps}
\end{center}
\caption{(a) Schematic of an ice growth on an aluminum plate under air and water flows.
(b) For $Q/l_{w}=1000$ [(ml/h)/cm] and $u_{\infty}=20$ m/s, dimensionless amplification rate $\sigma^{(r)}_{*}$ versus dimensionless wave number $k_{a*}$ for various ice thickness $b_{0}$. The thickness of air boundary layer in (a) is defined by $\delta=\delta_{0}/\bar{G}_{a*}$, where $\bar{G}_{a*}=0.413$ and $\delta_{0}=361$ $\mu$m for $u_{\infty}=20$ m/s.}
\label{fig:aluminum}
\end{figure}
Third, we mention the effect of heat conduction into a substrate beneath an ice sheet of finite thickness on the morphological instability. In the model herein, it was assumed that the ice region is semi-infinite and the undisturbed part of temperature gradient in the ice does not exist. We relax this assumption by including heat conduction into a planer aluminum substrate beneath the ice sheet. In Fig. \ref{fig:aluminum} (a), $b_{0}$ and $l_{\rm sub}$ are the thickness of ice and aluminum plate, respectively, $T_{\rm sub}$ is the temperature between the ice and aluminum plate and $T_{\rm sub0}$ is the temperature of other side of surface of the aluminum plate. Then, the undisturbed temperature gradient in the ice is
$\bar{G}_{s}=(T_{sl}-T_{\rm sub})/b_{0}$, and
Eq. (\ref{eq:amp}) is replaced by
\begin{equation}
\sigma_{*}^{(r)}=-\frac{dH_{l}^{(r)}}{dy_{*}}\Big|_{y_{*}=0}
+K^{s}_{l}k_{l*}\frac{\cosh(k_{l*}b_{0}/\bar{h}_{0})}{\sinh(k_{l*}b_{0}/\bar{h}_{0})}
\left(-G^{s}_{l}+H_{l}^{(r)}|_{y_{*}=0}-1\right),
\label{eq:aluminum-amp}
\end{equation}
where $G^{s}_{l}\equiv \bar{G}_{s}/\bar{G}_{l}$ is the ratio of the undisturbed temperature gradient at the ice-water interface in ice to that in water. The ice thickness $b_{0}$ is determined by integrating the following equation subject to an initial condition of $b_{0}=0$ at $t=0$: \cite{UFYT10}
\begin{equation}
\frac{db_{0}}{dt}=\frac{K_{s}}{L}\frac{T_{sl}-T_{\rm sub}}{b_{0}}
+\frac{K_{a}}{L}\frac{T_{sl}-T_{\infty}}{b_{0}/\bar{G}_{a*}}.
\label{eq:b0}
\end{equation}
If other side of surface of the aluminum plate is exposed to ambient cold air, i.e. assuming $T_{\rm sub0}=T_{\infty}$, $G^{s}_{l}$ can be expressed as \cite{UFYT10}
\begin{equation}
G^{s}_{l}=\frac{K_{l}}{K_{a}}\frac{\delta_{0}}{b_{0}}\Big/\left(1+\frac{K_{l}}{K_{\rm sub}}\frac{l_{\rm sub}}{b_{0}}\right),
\label{eq:Gsl}
\end{equation}
where $K_{\rm sub}=237$ ${\rm J/(m\,K\,s)}$ is the thermal conductivity of the aluminum plate.
Figure \ref{fig:aluminum} (b) shows the dimensionless amplification rate $\sigma^{(r)}_{*}$ versus dimensionless wave number $k_{a*}$ for various ice thickness $b_{0}$.
In the case of $b_{0}=\delta_{0}$, $\sigma^{(r)}_{*}$ is negative for all wave numbers, which means that the ice-water interface disturbances diminish with time.
On the other hand, in the case of $b_{0}=10\delta_{0}$, the ice-water interface disturbances in a finite range of wave numbers become unstable and $\sigma^{(r)}_{*}$ acquires a maximum value at a wave number. Noting that $G^{s}_{l}$ in Eq. (\ref{eq:Gsl}) is zero in the limit $b_{0} \gg \delta_{0}$ and $\bar{h}_{0}$ is the same order as $\delta_{0}$ as indicated in Table \ref{tab:tableI}, when $b_{0} \gg \delta_{0}$ Eq. (\ref{eq:aluminum-amp}) reduces to Eq. (\ref{eq:amp}) and the solid curve in Fig. \ref{fig:aluminum} (b) is the same as that for $u_{\infty}=20$ m/s in Fig. \ref{fig:ka-amp-lambda} (a).
When the ice thickness is small during the ice growth, the morphological instability of the ice surface is suppressed because the removal of the latent heat due to the conduction into the aluminum plate is dominant than that due to the convection by air and water flows. Even in the presence of the undisturbed part of temperature gradient $\bar{G}_{s}$ in ice, the morphological instability occurs when the ice thickness exceeds a critical value.
Finally, we mention some limitations of the proposed model.
First, freshwater icing sponginess containing non-negligible amount of liquid water is observed in icicles \cite{Makkonen88, Knight80} and aufeis. \cite{Streitz02} The spongy icing phenomenon is also well-known to the in-flight icing community. \cite{Blackmore03, Lozowski05} In the experiments of icing wind tunnel using a NACA 0012 airfoil, a considerable variation in sponginess (or liquid fraction) with air temperature, wind speed and liquid water content was observed. \cite{Lozowski05} The model herein cannot explain these experimental results. Therefore, it needs to modify an air-water-ice multi-phase system in Fig. \ref{fig:ice-water-air} to an air-water-spongy ice multi-phase system in Fig. \ref{fig:spongyice-water-air}, where a spongy layer is introduced in between a fully water region and a fully ice region.
However, since the water region is thin liquid film and the latent heat transfer is strongly affected by the existence of the water-air interface, the configuration in Fig. \ref{fig:spongyice-water-air} is fundamentally different from the mathematical models of flow-induced morphological instability of mushy layers developed in previous studies, \cite{Worster00, Worster92, Feltham99, Chung01, Neufeld08} where the liquid region was taken to be semi-infinite.
The conventional Stefan problem cannot describe the pattern formation observed in nature, because the dimensional information needed to set the scale of a crystal growth is absent. \cite{Langer80}
In other words, if the temperature at the ice-water interface is kept at 0 $^{\circ}$C and neglecting surface energy, the morphological instabilities occur on arbitrarily small length scales given any amount of supercooling. \cite{Worster00} In practice, surface energy limits the instability at some scale and an ice surface under a supercooled water film results in dendritic growth.
As a result, there can be a possibility of spongy ice formation, in which a portion of the surface liquid is incorporated into the dendritic ice matrix. \cite{Blackmore03} Hence the spongy layer is a mixture of ice and water and its temperature is a thermodynamic equilibrium one. \cite{Makkonen87} The use of this local equilibrium assumption is appropriate in the interior of the spongy layer at some distance from the tips of dendrites, where the increase in specific surface area of micro-scale phase boundaries promotes the release of latent heat into the interstices (pores) and hence the level of non-equilibrium can be kept very small. \cite{Worster92}
\begin{figure}[ht]
\begin{center}
\includegraphics[width=10cm,height=10cm,keepaspectratio,clip]{air-water-spongyice.eps}
\end{center}
\caption{Schematic of air-water-spongy ice multi-phase system.}
\label{fig:spongyice-water-air}
\end{figure}
In Fig. \ref{fig:spongyice-water-air}, since the interface between the spongy ice region and water region does not have a well defined position on the micro scale of dendritic growth, the spongy ice-water interface is defined as the envelope (suitably smoothed) of the dendrite ice matrix. \cite{Worster00, Worster92}
If the ice-water interface in Figs. \ref{fig:isotherms} (a) and (b) is replaced by the spongy ice-water interface, $T_{i}$ in Eq. (\ref{eq:Tsl}) is interpreted as the temperature at the spongy ice-water interface.
The local equilibrium assumption is likely to break down in the neighbourhood of the spongy ice-water interface, because if the spongy layer is postulated as a permeable material, \cite{Beavers67} a significant effect of the tangential and normal air shear stress disturbances found herein on the spongy layer through the thin water film can be expected and the latent heat advected by the water flow penetrates into near the spongy ice-water interface. The level of this non-equilibrium effect in the macro-scale is negligible at some distance from the spongy ice-water interface, where the temperature is 0 $^{\circ}$C, as shown in Fig. \ref{fig:isotherms} (b).
In order to gain a clear understanding of these viewpoints,
it is necessary to add equations governing local liquid fraction and the internal evolution of the spongy layer to the model herein and the effects of the shear stress at a disturbed spongy ice-water interface on the distribution of liquid fraction, permeability and penetration depth should be investigated along with a study on non-equilibrium coexistence of crystal and shearing liquid flow. \cite{Butler02} Furthermore, the dependence of the liquid fraction on the icing parameter such as air temperature, wind speed and liquid water content in the experiments by Lozowski et al.\cite{Lozowski05} must be explained.
Second, in the linear stability analysis, the amplitude of the ice-water interface disturbance of the most unstable mode increases with time:
$\delta_{b}(t_{*})=\exp(\sigma_{* \rm max}^{(r)}t_{*})\delta_{b}$,
which affects the magnitude of $h'_{x}/\bar{h}_{x}$ in Eq. (\ref{eq:h'x}).
In order to evaluate the value of $\sigma_{*\rm max}^{(r)}$, it is necessary to determine the disturbed flow and temperature fields in the water film which are influenced by surrounding airflow and temperature fields.
However, the linear theory is unable to clarify further features related to the development of disturbance.
We have to generalize the equation $d\delta_{b}(t_{*})/dt_{*}=\sigma^{(r)}_{* \rm max}\delta_{b}(t_{*})$ to a nonlinear amplitude evolution equation.
Third, the magnitude of $h'_{x}/\bar{h}_{x}$ depends on the shape of the ice-water interface.
In the normal mode analysis presented here, the values of $h'_{x}/\bar{h}_{x}$ depend on the supposed sinusoidal form in Eq. (\ref{eq:h'x}).
It is necessary to extend the current model to the problem of ice morphology produced on an arbitrary three-dimensional surface such as aircraft wings and overhead line cables, \cite{Myers02_2} taking into account water flow driven by both gravity and wind drag simultaneously.
Removing restrictions mentioned above and further extension of the current model to practical icing problems
will be discussed in later papers.
\begin{acknowledgements}
This study was carried out within the framework of the NSERC/Hydro-Qu$\acute{\rm e}$bec/UQAC Industrial Chair on Atmospheric Icing of Power Network Equipment (CIGELE) and the Canada Research Chair on Engineering of Power Network Atmospheric Icing (INGIVRE) at the Universit$\acute{\rm e}$ du Qu$\acute{\rm e}$bec $\grave{\rm a}$ Chicoutimi.
The authors would like to thank all CIGELE partners (Hydro-Qu$\acute{\rm e}$bec, Hydro One, R$\acute{\rm e}$seau Transport d'$\acute{\rm E}$lectricit$\acute{\rm e}$ (RTE) and $\acute{\rm E}$lectricit$\acute{\rm e}$ de France (EDF), Alcan Cable, K-Line Insulators, Tyco Electronics, Dual-ADE, and FUQAC) whose financial support made this research possible. The authors would also like to thank H. Tsuji and anonymous referees for their useful comments.
\end{acknowledgements}
|
\section{\hskip -1em.~~#1}\vspace{-3pt}}
\newcommand{\SubSection}[1]{\vspace{-3pt}\subsection{\hskip -1em.~~#1}
\vspace{-3pt}}
\pagestyle{empty}
\begin{document}
\twocolumn[
{\LARGE \bf Modeling the Envelope of the Secular Light Curve of the Comet 1P/Halley }
\vskip10mm
\begin{raggedright}
{\small \centering{{\bf Eduardo Rond\'on} (1,2) and Ignacio Ferr\'in (2)}\\
\vskip5mm
(1) N\'ucleo Alberto Adriani, Facultad de Ingenier\'ia, Universidad de Los Andes, M\'erida, Venezuela, (2) Centro de F\'isica Fundamental, Facultad de Ciencias, M\'erida, Venezuela (<EMAIL> ).}
\end{raggedright}
\vskip10mm %
]
\thispagestyle{empty}
\section*{Abstract}
We have been modeling the envelope of the secular light curve (Ferrín, 2010)\cite{Ferrin10} of comet 1P/Halley. As the first step we have modeled the water production rate of the comet, for several active regions as a function of visual albedo and infrared albedo, and then we have used the correlation between the water production rate and the reduced magnitude (Jorda, 2008)\cite{Jorda08} for modeling the envelope of the secular light curve of the comet. We obtain probable orientations of the rotation axis (I,$\Phi$). These orientations are compared with several solutions by several authors. We have calculated the surface temperature for the orientation of the rotation axis with minimal standar deviation. We have found that for near and far values of the perihelion the surface temperature as a function of the latitude is constant.
\Section{Introduction}
Comet Halley has a large number of observed apparitions, and many visual observations have been published in the literature. Ferr\'in (2010)\cite{Ferrin10} has published brightness observation of several comets, considering that the envelope of the data describes the behavior of the brightness in time. Several secular light curves show asimmetry. We have developed a model that predicts this asimmetry. The orientation of the rotation axis explains the asimmetry in the secular light curve (Rond\'on \& Ferrín, 2010; Rond\'on, 2007)\cite{Rondon10}\cite{Rondon07}. The importance of studying the secular light curve is that it gives a large amount of physical informations of comet. By modeling this curve we can explain the behavior of the brightness mathematically and we can predict the physical parameters of the curve.
\Section{Model Calculations}
The first step for modeling the light curve is to calculate the sublimation rate of water. The equation that describes the vaporization rate of a comet is the energy conservation equation, given by:
\begin{eqnarray}
F_{0}(1-A_{v})r_{H}^{-2}\overline{cos(\theta)}= (1-A_{ir})\sigma T^{4}+ \nonumber \\
\hspace{2cm} Z(T)L(T)+ K \frac{\partial(T)}{\partial(z)}
\end{eqnarray}
Where A$_{ir}$ is the infrared albedo, r$_H$ Sun-comet distance, $\overline{cos(\theta)}$ is the projection factor (Cowan\&A'Hearn, 1979)\cite{Cowan79}, $\sigma$ is the Steffan Bolztmann constant, T is the temperature, Z(T) is the sublimation function, L(T) is the latent heat function, K is the thermal conductivity constant, z is the layer depth.
\begin{equation}
Z(T)=\frac{P(T)m}{2(\pi)kT}
\end{equation}
where $P(T)$ is the vapor pressure function, m is the molecular weight, k is the ideal gas constant.
If we know the vapor pressure function and the latent heat function, we can solve for the energy conservation equation(1).
\begin{equation}
Z_{total}=\overline{Z(i)}=\frac{1}{2}{\int_{-\pi/2}^{\pi/2} (Z(i,b)cos(b)db)}
\end{equation}
We have modeled the water production rate for comet 1P/Halley using the observational data given by (Schleicher, 1998)\cite{Schleicher98}, and we have found that when considering an active region of 80 km² we can predict the average of the observational data (Fig.1a). However, the secular light curve considers the envelope of the data (Ferrin, 2005)\cite{Ferrin05}. For this case we have taken an active region of 180 km²(Fig.1b). In both cases we have assumed an visual albedo $A_v=0.009$, an infrared albedo $A_ir=0.5$ and a thermal conductivity K=0.
The correlation equation between the reduce visual magnitude and the water production rate is known (Jorda, 2008)\cite{Jorda08} and is given by:
\begin{equation}
m=125.051-4.077\log(Z_{total})
\end{equation}
\\
The secular light curve of comet Halley is assymetric. This can be explained through of the orientation of the rotation axis of comet (Rond\'on, 2007)(Rond\'on \& Ferr\'in, 2010)\cite{Rondon07}\cite{Rondon10}.
\begin{figure}[!ht]
\vspace{-0.1 cm}
\begin{center}
\includegraphics[width=2.7 in]{Simulacion_Curva_Agua_Area80.jpg}
\vskip-5mm
\includegraphics[width=2.7 in]{Simulacion_Curva_Agua_Area180.jpg}
\vskip-5mm
\caption{Water production rate vs log(r), a) with an active region of 80 km² b) with an active region of 180 km².}
\end{center}
\end{figure}
\begin{figure}[!ht]
\vspace{0.2 cm}
\begin{center}
\includegraphics[width=2.5 in]{Stadistic2.jpg}
\caption{Standard deviation for each of the orientation of the rotation axis for the comet 1P/Halley.}
\label{fig2}
\end{center}
\end{figure}
\vskip -1.5mm
We have calculated the standar deviation for our model with the observational data, and found that the solution with minimal standar deviation is for a obliquity, I = 90° and a pole orbital longitude, $\Phi$ = 112°, (Fig, \ref{fig2}). Several authors obtained solutions for the orientation of the rotation axis (Sekanina, 1986) \cite{Sekanina86}.
\begin{figure}[!ht]
\vspace{-0.3 cm}
\begin{center}
\includegraphics[width=2.7 in]{Simulacion_Curva_de_Luz.jpg}
\vskip -3mm
\caption{Secular light curve of the Halley comet with a orientation of the rotation axis of I = 90°, $\phi$ = 112°. In the upper envelope curve has been considered an active region of 180 km² and in the lower envelope curve has been considered an active region of 80 km².}
\label{fig3}
\end{center}
\end{figure}
\vskip-7mm
\
\\
In Figure \ref{fig2} we can see that a region of the solution of (Grun, 1986)\cite{Grun86} has a standard deviation of only 1 magnitude. In the Figure \ref{fig3} we have modeling the secular light curve assuming an orientation of the rotation axis of I= 90°, $\Phi$=112° for an active region of 80 km²(blue line), and 180 km², (red line). In this graph we can see that before perihelion the model fits the observational data, but after perihelion the fits is not very good. This behavior is due to the thermal conductivity effect in the nuclear surface that has not been considered in the model.
\
In Figure 4 we can see that the surface temperature as a function of the latitude is constant for values of the heliocentric distance near and very far away of perihelion. This behavior near of perihelion is that the coma of comet tends to stabilize the temperature, while for heliocentric distance very far away perihelion the sun radiation is small and therefore the whole cometary nucleus cools.
\
\\
\\
\begin{figure}[!ht]
\vspace{-0.3 cm}
\begin{center}
\includegraphics[width=2.5 in]{Comportment_of_the_Surface_Temperature.jpg}
\vskip 1mm
\caption{Surface temperature as a function of heliocentric distance, assuming an active region of 180 km² and a orientation of the rotation axis of I= 90°, $\Phi$ = 112°.}
\label{fig4}
\end{center}
\end{figure}
\
\Section{Summary and Conclusions}
We have developed a theoretical model capable of reproducing the observational data of the secular light curve and water production rate of comet Halley, with minimal standar deviation $\sigma$ = 0.3 mag, for an orientation of the rotation axis I = 90°, $\phi$ = 112°. The region of the solution shown by (Grun, 1986)\cite{Grun86} explain the secular light curve with a standard deviation $\sigma$=0.6 5 mag. We also have found that the surface temperature as a function of the latitude is constant for values of the heliocentric distance near and very far away of perihelion. \
|
\section{Introduction}
\CVS is a member of the RVSb$_{3}$ (R=rare-earth) family, with an orthorhombic crystal structure. Two systematic studies \cite{Hartjes97,Sefat08} of this family showed interesting physical properties such as high anisotropy with a quasi two-dimensional crystal structure, and different types of magnetic order when the rare-earth is changed. Similarly complex properties were also observed in other binary and ternary rare-earth antimonide families, such as RSb$_{2}$ \cite{Budko98}, RCrSb$_{2}$ \cite{Leonard99} and RAgSb$_{2}$ \cite{Myers99}.
\CVS is the only ferromagnetic compound from the RVSb$_{3}$ family, with \Tc around 4.6~K \cite{Sefat08}. It may be considered as a moderately heavy fermions system as its $\gamma$ value is found to be 162~mJ/mol K$ ^{2} $ below 2~K \cite{Sefat08}. Only a few studies involving this compound have been reported \cite{Hartjes97,Sefat08,Vannette08}.
Similar ferromagnetic, Ce-based compounds, such as CeNiSb$_{3}$ or CeAgSb$_{2}$ were studied under pressure by resistivity measurements \cite{Sidorov03,Nakashima03,Sidorov05Ce}, and revealed complex phase diagrams, with ferromagnetic transitions evolving into antiferromagnetic ones under pressure. Among the Ce-based ferromagnets studied under pressure to date, none have exhibited superconducting behavior.
The reported increase of \Tc for applied pressures up to 1~GPa \cite{Sefat08} motivated us to continue investigations on \CVS at higher pressures; ultimately it is reasonable to expect \Tc to pass through a local maximum value and then decrease \cite{Doniach77}. Ideally this would present a good opportunity to study possible quantum criticality in a Ce-based ferromagnet. We present here resistivity and ac-calorimetry measurements under pressure up to 8 GPa, in a Bridgman anvil cell modified to use a liquid medium and a diamond anvil cell, respectively. \Tc behaves as expected from the Doniach model \cite{Doniach77} with an initial increase with pressure up to a maximum above which a fast decrease and ultimately disappearance of \Tc is observed. The low temperature power law fits of the resistivity are in agreement with the disappearance of the magnetic transition at a quantum critical point.
We observed discrepancies in the \Tcf\textit{(P)} behavior between pressure cells using different pressure media, and attributed it to the differences of pressure conditions, with a slight uniaxial component existing along the cell axis in the Bridgman anvil cell. Further measurements in the modified Bridgman cell with a different, more hydrostatic, pressure medium, confirmed this assumption. With differently oriented samples we have studied the \Tcf\textit{(P)} dependencies for different directions of the uniaxial component of pressure.
\section{Experimental details}
Single crystals of \CVS were grown out of antimony flux as detailed by Sefat \textit{et al.} \cite{Sefat08}. Resistivity and specific heat measurements were performed on these crystals up to 7.6 and 6.9~GPa, respectively.
The resistivity samples were measured by a four probe method using the AC-transport option of a Quantum Design Physical Property Measurement System (PPMS) down to 1.8~K or a LakeShore 370 AC resistance bridge with a $ ^{3} $He cryostat down to 400~mK. Four, 12.5~$ \mu m $ diameter, gold wires were spot-welded to each polished and cut crystal with typical dimensions of 600 $\times$ 150 $\times$ 40~$\mu m^{3}$. Unless specified otherwise, the resistivity was measured along the \textit{c}-axis with the sample larger dimensions respectively along the \textit{c} and \textit{b}-axis. The measurement current was 1~mA and the frequency was 17~Hz. Before each sample was loaded into the pressure cell, the resistivity was measured at ambient pressure on a standard PPMS puck. A reproducible \Tc of 4.56~K was deduced from a sharp peak in the resistivity derivative, similarly to an averaged value of 4.6~K found previously \cite{Sefat08}.
Before performing studies under pressure, we measured several samples at ambient pressure with current flowing along each of the three crystallographic directions of the orthorhombic structure (at least two samples for each direction). We observed a good reproducibility in the resistivity behavior, although the uncertainties associated with measuring the relatively small sample dimensions lead to an error in the resistivity value at room temperature of up to 30-40~\% . Each sample's orientation was identified from the crystal's morphology, as was discussed by Sefat \textit{et al.} \cite{Sefat08} without any further X-ray Laue measurements. The reproducibility in resistivity from one sample to another was considered an indication that contributions from the other components of the resistivity were low or absent.
In addition, thermal expansion was measured at ambient pressure using a capacitive dilatometer constructed of OFHC copper, mounted in a Quantum Design PPMS instrument. A detailed description of the dilatometer is presented elsewhere \cite{Schmiedeshoff06}. The samples were lightly polished so as to have parallel surfaces approximately parallel to the different crystallographic axis directions. The dimensions range from 0.5~mm to a few mm. Measurements were performed on warming. We define the thermal expansion coefficients as $ \alpha_{i}=\dfrac{1}{L_{i}}\dfrac{dL_{i}}{dT} $ with $L_{i}$ being one of the 3 sample's principle crystallographic orientations, and the volume thermal expansion coefficient $ \beta=\Sigma(\alpha_{i}) $.
Resistivity measurements under pressure were performed using a Bridgman cell modified to use with a liquid pressure medium \cite{Colombier07,Colombier09}, either a Fluorinert mixture (1:1 FC70:FC77) or 1:1 n-pentane:isopentane. When not specified, the medium used was 1:1 FC70:FC77. A piece of lead, to use as a manometer, and the sample were inserted in a pressure chamber of 1.4~mm inner diameter. The typical transition widths for lead were 15~mK and 40~mK, respectively for 1:1 n-pentane:isopentane and 1:1 FC70:FC77.
Although, ideally, we would like pressure to be hydrostatic (i.e. isotropic), even with a medium that is a liquid at ambient conditions, once the media freezes at room temperature the application of pressure is expected to give rise to non-hydrostaticity that, as a first approximation, can be thought of as an application of hydrostatic pressure as well as a smaller uniaxial pressure. Given our cell geometry, if such a uniaxial component exists, it is anticipated to be in the direction perpendicular to the thin-disk-like sample space volume, i.e. along the cell axis. If we align the sample with one of its crystallographic axes along this direction, then we will say that "pressure is applied along this direction" to identify this potential uniaxial direction. For example, we use in the following the notation $ \rho_{c,P//a} $ to refer to the resistivity measured with current along the \textit{c}-axis and with pressure applied along the \textit{a}-axis of the sample.
Although the pressure is not purely hydrostatic, it is reproducible. Three samples were measured with similar pressure conditions (Bridgman cell filled with Fluorinert), current along the \textit{c} axis and pressure applied along the \textit{a} axis. The reproducibility of results was confirmed by similar \textit{T(P)} phase diagram data.
1:1 n-pentane:isopentane offers pressure conditions much closer to hydrostaticity, compared to 1:1 FC70:FC77 in the Bridgman cell pressure range, as it is known to freeze above 5~GPa at 300~K instead of below 1~GPa for the Fluorinert mixture \cite{Sidorov05}. However, it is more difficult to handle because of its high compressibility in the low pressure range \cite{Stella_bientot} (below 2~GPa) and because its boiling point is close to room temperature (28.5$^{\circ}$C for isopentane). Due to these difficulties, one of the resistivity data sets, in 1:1 n-pentane:isopentane media, was taken in a three wire configuration after the failure of one of the wires. The resulting three wires resistivity measurement gave limited quantitative information, but a sharp transition was still observable, and its derivative, (shown in figure \ref{Derivative}.a. below) looks very similar to those obtained from four wires measurements, once the data for the first pressure are scaled to ambient pressure.
The specific heat under pressure was measured in a diamond anvil cell \cite{Salce00,Demuer00} up to pressures of 7~GPa cooled down to 1.5~K using a $^{4}$He cryostat. The culet size of the anvils was 0.7~mm. The pressure, changed \textit{in-situ} at low temperatures \cite{Salce00}, was read using the ruby fluorescence method. Argon was chosen as a pressure medium. Albeit solidified at 1.4~GPa and 300~K, argon provides close to hydrostatic conditions due to its weak interatomic interactions (i.e a very soft solid). Three different pressure runs were performed. For one of these, two pressure cycles were realized by decreasing pressure in one step after a first run with increasing pressure. The ac-calorimetry method \cite{Demuer00} was used; a quasi-sinusoidal excitation was applied to the sample by a laser via a mechanical chopper. The temperature oscillations of the sample (inversely proportional to the specific heat) were measured with a Au/AuFe (0.07\%) thermocouple which was spot-welded on the sample. We estimated the amplitude of temperature oscillations of the sample $T_{ac} $ from the thermocouple voltage measured $V_{ac} $ and the thermoelectric power of the thermocouple $S_{th} $: $ T_{ac}= \vert V_{ac} \vert / S_{th} $.
\section{Results}
\subsection{Ambient pressure}
Our examination of \CVS (\textit{a}=13.172 \AA , \textit{b}=6.2419\AA , \textit{c}=6.0327 \AA) \cite{Sefat08} under pressure includes the study of the anisotropic properties of \CVS and in particular its sensitivity to slight uniaxial strains. In order to accomplish this, we first investigated the anisotropic resistivity with current \textit{i} flowing along the three crystalline directions of this compound at ambient pressure (figure \ref{anisot}.a). The resistivity ratios between 300~K and 2~K range from 2, when the current flows along the \textit{a}-axis, to 5.5 along the \textit{b}-axis. The results for current along the \textit{b} and \textit{c} axis are consistent with the study from Sefat \textit{et al.} \cite{Sefat08}, although resistivity values are lower in our measurements.
\begin{figure}[!ht]
\begin{center}
\includegraphics[angle=0,width=120mm]{fig1a.eps}
\includegraphics[angle=0,width=120mm]{fig1b.eps}
\end{center}
\caption{(Color online) a. Resistivity, at ambient pressure, of \CVS along the \textit{b} and \textit{c} axis. The resistivity along the third direction is added to the two other in the inset. b. Anisotropic thermal expansion coefficients of \CVSf , inset shows expanded, low temperature range.}
\label{anisot}
\end{figure}
A clear local maximum is observed at around 16~K for \textit{i} along the \textit{b}-axis, and is barely detected for \textit{i} along the \textit{c}-axis. A more striking anisotropy is the resistivity measured with current along the \textit{a}-axis, roughly 10 times higher than along the two other directions.
The resistivity thus tends to be quasi two-dimensional, and in the following we study the resistivity along the \textit{b} or \textit{c} axis, depending on the geometry needed.
In addition, a strong anisotropy in the thermal expansion coefficients is shown in figure \ref{anisot}.b. The broad local maximum of the volume thermal expansion coefficient around 10~K may be related to the Kondo temperature. We applied the Ehrenfest relation for second order phase transitions,
\begin{center}
$ \dfrac{dT_{C}}{dP_{i}}=\dfrac{V_{m}\Delta \alpha_{i}T_{C}}{\Delta C_{p}} $ ; $ \dfrac{dT_{C}}{dP}=\dfrac{V_{m}\Delta \beta T_{C}}{\Delta C_{p}} $
\end{center}
where $V_{m}$ is the molar volume, $ \Delta \alpha_{i} $ and $ \Delta \beta $ are respectively a change in the linear or volume thermal expansion coefficients at the phase transition, and $ \Delta C_{p} $ is a change in the specific heat \cite{Sefat08} at the phase transition. From this relation, we deduced a substantial uniaxial pressure dependence anisotropy for \Tcf , of 0.4 K/GPa, 0.2 K/GPa and 0.7 K/GPa when the pressure is respectively applied along the \textit{a}, \textit{b} and \textit{c} axes. The addition of these three components gives \textit{d}\Tcf \textit{/dP}=1.4~K/GPa, very close to the low pressure slope \textit{d}\Tcf \textit{/dP}=1.2~K/GPa found from the pressure temperature phase diagram (see figure \ref{PD} below).
\subsection{Resistivity under pressure}
The main interest of the present study is in the investigation of the evolution of \Tc under pressure. The modified Bridgman cell used with Fluorinert as a pressure medium is known to present a slight uniaxial component in addition to the expected isotropic pressure due in part to its low, room temperature freezing pressure, below 1~GPa. As an example, the iron arsenide superconductors, recently measured in this cell \cite{Colombier09,Colombier10}, are known to be sensitive to these uniaxial stresses which stabilize the superconducting phase. This superconducting phase is then observed in a broader pressure range of the phase diagram in presence of a uniaxial component of pressure. Since \CVS is an orthorhombic compound with clear anisotropy and some degree of electronic correlation, we decided to check its sensitivity to uniaxial component of pressure associated with non-hydrostaticity. The grown crystals are relatively large and mechanically sturdy making them easy to polish to three different geometries to allow for this study. (This is in contrast to iron arsenides, which were soft and easily exfoliated along their tetragonal, \textit{c}-axis.)
The temperature dependent resistivity data of \CVS measured with the pressure successively applied along the three crystallographic directions are shown in figure \ref{Resistivity}. We had to measure the resistivity along two different directions so as to fully investigate the response of the crystal to slight uniaxial stresses, but the evolution of the anisotropy of resistivity under hydrostatic pressure was not the main purpose of this work.
\begin{figure}[!ht]
\begin{center}
\includegraphics[angle=0,width=80mm]{fig2a.eps}
\includegraphics[angle=0,width=80mm]{fig2b.eps}
\includegraphics[angle=0,width=80mm]{fig2c.eps}
\end{center}
\caption{(Color online) Resistivity measurement of \CVS under pressure. Sketches illustrate the sample orientations in the pressure cells. Insets: low temperature resistivity. a. with current along the \textit{c}-axis and pressure applied along the \textit{a}-axis. b. with current along the \textit{c}-axis and pressure applied along the \textit{b}-axis. c. with current along the \textit{b}-axis and pressure applied along the \textit{c}-axis.}
\label{Resistivity}
\end{figure}
In all cases, the resistivity above \Tc increases with pressure. \Tc itself initially increases with pressure, reaches a maximum value, and then decreases with pressure and finally disappears. The transition is sharp at ambient pressure and broadens progressively. It is difficult to distinguish it as \Tc drops towards 0~K. The resistivity curves presented in figures \ref{Resistivity}. a. and b. are obtained with the same current direction, but the transition temperature increase is slower in b (see \textit{T(P)} phase diagram in figure \ref{PD} below). This shows evidence for anisotropy of the pressure response of the crystals, as the reproducibility of results was checked for three pressure runs in similar conditions. For current along the \textit{b}-axis (figure \ref{Resistivity}.c), the local maximum observed at ambient pressure is still present under pressure; it progressively broadens as it is shifted up to higher temperatures. For each direction of applied pressure, there is a clear and consistent increase of $ \rho _{300K} $ over the measured pressure range.
\begin{figure}[!ht]
\begin{center}
\includegraphics[angle=0,width=53mm]{fig3a.eps}
\includegraphics[angle=0,width=53mm]{fig3b.eps}
\includegraphics[angle=0,width=53mm]{fig3c.eps}
\includegraphics[angle=0,width=53mm]{fig3d.eps}
\end{center}
\caption{(Color online) Resistivity derivative $ d\rho (T)/dT $ of \CVS under pressure. Sketches outline the sample orientations in the pressure cells. a. with current along the \textit{c}-axis and pressure applied along the \textit{a}-axis in a cell filled with 1:1 n-pentane:isopentane (in $\mu \Omega$ $cm /K $ at 0~GPa and arbitrary units under pressure). b. with the same orientation, but filled with 1:1 FC70:FC77 c. with current along the \textit{c}-axis and pressure applied along the \textit{b}-axis with 1:1 FC70:FC77 as a pressure medium. d. with current along the \textit{b}-axis and pressure applied along the \textit{c}-axis with 1:1 FC70:FC77 as a pressure medium.}
\label{Derivative}
\end{figure}
The low temperature resistivity derivative data, $ d\rho (T)/dT $, are compared in figure \ref{Derivative}, for the three different cells configuration shown in figure \ref{Resistivity} as well as an additional cell filled with 1:1 n-pentane:isopentane. The influence of sample orientation on \Tc is even more obvious when the data are presented in this manner. The highest \Tc value is observed in fig. \ref{Derivative}.d, for the \textit{c}-axis of the sample aligned with the cell axis. In the graphs a. and b., the samples' orientations are the same but two different pressure media are used: 1:1 n-pentane:isopentane and 1:1 FC70:FC77, respectively. We observe a strong dependence on pressure conditions. Whereas the feature remains sharp until the highest pressure of 4.5~GPa with 1:1 n-pentane:isopentane (figure \ref{Derivative}.a), it is already broadened significantly at a similar pressure in Fluorinert (figure \ref{Derivative}.b), and \Tc is much lower. In the experiment with Fluorinert, the transition temperature broadens significantly for pressures above 4~GPa.
\subsection{Phase diagram}
Figure \ref{PD} shows the phase diagrams obtained from several runs with different pressure conditions and different crystal orientations in the modified Bridgman cell, together with data points inferred from the piston-cylinder cell magnetization data and the diamond anvil cell specific heat data.
\begin{figure}[!ht]
\begin{center}
\includegraphics[angle=0,width=90mm]{fig4a.eps}
\includegraphics[angle=0,width=90mm]{fig4b.eps}
\end{center}
\caption{(Color online) \textit{T(P)} phase diagram of \CVSf . We added data from magnetization measurements performed in a piston-cylinder cell \cite{Sefat08}. For a given symbol, experimental conditions were similar, and the different colors refer to different runs. a. Comparison between the different axis orientations in the Bridgman anvil cell. Diamond anvil cell $C _{P} $ data points are also shown. Crosses are the lowest measured temperature for the lowest pressure for which no phase transition could be detected. $P_{1}$, $P_{2}$, $P_{3}$ and $P_{4}$ are given by downwards arrows and refer to the critical pressures estimated from figure \ref{LTfit}. b. Comparison between the different pressure media used.}
\label{PD}
\end{figure}
For each pressure run, we observe a similar dome-shaped phase diagram. However, the data from runs with different media and orientations are somewhat scattered. Whereas all curves overlap below at least 2~GPa, differences in \Tcf \textit{(P)} are observed at higher pressures.
We observe obvious differences between the 3 crystal orientations measured in the modified Bridgman cell, figure \ref{PD}.a. The maximum values of \Tc range from 7.9~K to 9.3~K and the corresponding pressures from 3.2~GPa to 4.2~GPa. More importantly, the critical pressure, the pressure at which the \textit{T(P)} curve extrapolates to zero, ranges from roughly 5.5 to 7~GPa. As this behavior is reproducible within 0.3~GPa for three different runs when the cell axis coincides with the \textit{a} crystallographic axis, we assume the differences seen come from an anisotropic response to the slight uniaxial component present in the modified Bridgman cell. This strong anisotropy and particular sensitivity to uniaxial component of pressure along the \textit{a} crystallographic axis is confirmed when we use a more hydrostatic pressure medium. Two runs in the modified Bridgman cell with the same crystal orientation (which appears to be the most sensitive to uniaxial pressure) are shown figure \ref{PD}.b, one with Fluorinert and one with 1:1 n-pentane:isopentane as a pressure medium. Here again we observe differences between the two runs in the maximum value of \Tcf , its corresponding pressure, and the critical pressure. 1:1 n-pentane:isopentane is known to freeze at room temperature above 5~GPa and Fluorinert freezes below 1~GPa \cite{Sidorov05}. This means that, contrary to the other medium, 1:1 n-pentane:isopentane was always liquid at room temperature in this experiment, and so was much closer to hydrostaticity. (A conclusion established by the superconducting transitions widths of the lead manometers, given in the experimental details section.) The basic agreement between the 1:1 n-pentane:isopentane data and the $C_{p} $ data taken in argon (discussed below) and the higher pressure deviation of the Fluorinert data from this manifold is further evidence that the discrepancies in the phase diagram can be attributed to an anisotropic sensitivity of the sample to a uniaxial component of pressure. Keeping in mind the strong sensitivity of \CVS to pressure conditions, we try to be very cautious about the impact of pressure conditions to our results.
To estimate the evolution of the samples' sensitivity to pressure conditions, we checked the broadening of the magnetic transition. The lead, as a soft material, is not very sensitive to deviations to hydrostaticity and the broadening of the superconducting transition is modest \cite{Colombier09}. The transition broadening of \CVS would indeed be a more efficient clue as long as we are able to estimate also the effects intrinsic to the magnetism. We estimated in figure \ref{Delta_TCurie} the broadening of the transition by comparing two different criteria for \Tcf : the maximum of the peak in the $ d\rho /dT$ derivative and the onset of this peak from two asymptotes (as shown by dashed lines in the inset). By comparing cells measured in different pressure conditions, we get a good sense of pressure effect versus intrinsic properties of the compound. As $ d\rho /dT$ appears to be very similar to the heat capacity feature around the transition, similar criteria were applied to $C_{P}(T)$ data.
\begin{figure}[!ht]
\begin{center}
\includegraphics[angle=0,width=90mm]{fig5.eps}
\end{center}
\caption{(Color online) Evolution with pressure of the difference in temperature between two criteria for the magnetic transition. Several curves are shown for different resistivity measurement conditions and one for the specific heat. Three data sets are shown for \textit{P//a} (stars). The inset shows the definition for these two criteria, for the resistivity derivative and the specific heat.}
\label{Delta_TCurie}
\end{figure}
At ambient pressure, the difference between both criteria is around 200~mK in resistivity and it increases only slightly up to around 400~mK at 3~GPa. Above 3 or 4~GPa, the transition broadens strongly, to above 4~K in the modified Bridgman cell filled with Fluorinert. The broadening observed with the 1:1 n-pentane:isopentane set of measurements is at least a factor of two smaller, compared to Fluorinert, with only a slight increase at the highest pressure. We did not reach pressures above 4.5~GPa, because of the low compressibility of this medium and a pressure medium leak observed at the lowest pressure. Whereas the effect of non-hydrostatic conditions on the transition broadening is obvious, some degree of broadening may be an intrinsic property of the material once \Tc starts to decrease at high pressures, as it generally seems to occur at pressures above the maximum of the magnetic transition temperature.
To further our investigation of the influence of pressure non-hydrostaticity, a noble-gas as a pressure medium was used to provide a near hydrostatic reference. Even when it is solid, its low interatomic interactions indeed allow excellent pressure conditions, as can be seen in figure \ref{Delta_TCurie}. This experiment entailed the measurement of specific heat in a diamond anvil cell with argon as a pressure medium, and is described below.
\subsection{ac-calorimetry}
In figure \ref{Cp}, we present the temperature dependent specific heat curves of \CVS obtained from one of the three pressure runs.
\begin{figure}[!ht]
\begin{center}
\includegraphics[angle=0,width=90mm]{fig6.eps}
\end{center}
\caption{(Color online) Specific heat of \CVS (in arbitrary units) under pressure, measured in the diamond anvil cell. Pressures are given in GPa.}
\label{Cp}
\end{figure}
The transition at the lowest pressures is sharp with a shape similar to the ambient pressure measurement \cite{Sefat08}. The 1.0~GPa \Tc value inferred from the data presented in figure \ref{Cp} is in good agreement with that inferred from the magnetization data at 1.0~GPa \cite{Sefat08}. \Tc progressively increases with pressure until 4.3 GPa and then decreases. The transition progressively broadens, and its amplitude also seems to decrease, although the background and the signal amplitude might be a little different from one measurement to another. Whereas at pressures of 6.0 and 6.3~GPa, a feature is still clearly seen, we can just barely resolve a broad bump in the 6.9~GPa data. We can see in figure \ref{Delta_TCurie} the stronger broadening of the transition when \Tc decreases, similarly to what is found in the Bridgman cell resistivity data. However, as the pressure conditions are closer to hydrostaticity compared to a Bridgman cell filled with Fluorinert, the transition is twice as sharp, in particular at the highest pressures as seen in figure \ref{Delta_TCurie}. The agreement is much better when the pressure medium is 1:1 n-pentane:isopentane with sharp transitions in both cases, in the overlapping pressure range.
The \Tcf \textit{(P)} values obtained are also shown in the phase diagrams in figure \ref{PD}. The diamond anvil cell can be considered as the reference since the pressure conditions are presumed to be the closest to hydrostaticity. We observe an increase of \Tc from 4.6~K to as high as 9.7~K when the pressure increases from 0~GPa to 4.3~GPa. It then decreases and we expect to have a critical pressure around 7-7.5~GPa. From one run to another, only differences in maximum \Tc are noticed, and they are below 1~K. These differences may also be linked to pressure conditions. Light and medium gray triangles in figure \ref{PD} respectively refer to the first and the second pressure increase realized with the same diamond anvil cell in specific heat. We observe a slightly lower maximum \Tcf , around 0.5~K, for the second run, when the sample may be more strained. These differences between runs even with a noble gas as a pressure medium emphasize here again the extreme sensitivity of \CVS to pressure conditions. The diamond anvils cell axis (along which the load is applied) is coincident with the \textit{a}-crystallographic axis, and the phase diagram is moved up to higher pressures and temperatures, compared to the Bridgman cell measurement using the same sample orientation. Compared to 1:1 n-pentane:isopentane, difference are more subtle, and seem to mainly consist in a lower value of maximum \Tc in the Bridgman cell.
\subsection{Quantum criticality}
Whereas differences in the pressure dependence are noticed for the several pressure runs shown in the phase diagram figure \ref{PD}, the general behavior and in particular the way the magnetic transition is suppressed are similar. As our main interest is to determine if we observe a quantum critical point, we performed further low temperature measurements in a $ ^{3} $He cryostat. From these measurements we made low temperature resistivity fits using the equation: $ \rho (T)=\rho _{0}+AT^{n}$, where either (i) \textit{n} equals to 2 or (ii) \textit{n} was treated as a free fitting parameter. As measurements in a PPMS down to $ ^{4} $He temperatures are much more convenient, we performed only a few measurements down to 0.35~K, so as to check that the fits down to 1.8~K gave qualitatively similar results. Two $ ^{3} $He measurements were performed above P$ _{c} $ when the pressure is applied along the \textit{c}-axis of the crystal and another whole set of measurements was made for \textit{P} $>$ 3~GPa with pressure applied along the \textit{a}-axis. Figure \ref{LTfit} presents fit data from when \textit{n} was left as a free parameter. We determined the temperature range of the fit either by a progressive increase of the maximum fit temperature, or by checking the linear behavior of the law $ Log(\rho (T) )=Log(\rho _{0} )+nLog(A)Log(T) $, when $\rho _{0}$ was slightly modified. The temperature ranges and fit results obtained from both methods were in good agreement, the maximum fit temperature being around 3.5-4~K for the measurements in a $ ^{4} $He cryostat. The results of fits performed in the $ ^{3} $He and in the $ ^{4} $He cryostat are in a good qualitative agreement but the parameters values (specifically \textit{n}) can differ of as much as 40\% around the critical pressure (where the magnetic transition disappears).
\begin{figure}[!ht]
\begin{center}
\includegraphics[angle=0,width=60mm]{fig7a.eps}
\includegraphics[angle=0,width=60mm]{fig7b.eps}
\includegraphics[angle=0,width=60mm]{fig7c.eps}
\end{center}
\caption{(Color online) Pressure dependence of the parameters obtained from a low temperature fit $ \rho (T)= \rho _{0} + A T^{n} $. The arrows labeled $P_{1} $, $P_{2} $, $P_{3} $ and $P_{4} $ are the estimated critical pressures (see text). The triangles refer to fits down to $ ^{3} $He temperatures. The colors are chosen the same as in figure \ref{PD}. a. \textit{A} coefficient. b. Temperature exponent, \textit{n}. c. Residual resistivity, $\rho _{0}$. }
\label{LTfit}
\end{figure}
\begin{figure}[!ht]
\begin{center}
\includegraphics[angle=0,width=90mm]{fig8a.eps}
\includegraphics[angle=0,width=90mm]{fig8b.eps}
\end{center}
\caption{(Color online) Pressure dependence of the parameters obtained from a low temperature fit $ \rho (T)= \rho _{0} + A T^{2} $ from $^{3}$He data of $ \rho_{c,P//a} $. a. \textit{A} coefficient. The inset shows the maximum temperature where this fit applies. b. Residual resistivity, $\rho _{0}$. }
\label{LTfitT2}
\end{figure}
The general behavior of \textit{A} as well as \textit{n}, shown in figures \ref{LTfit}.a and \ref{LTfit}.b, is similar but shifted in pressure for the three orientations of the sample axes with respect to the cell axis. The \textit{A} parameter presents a strong peak around the pressure where the magnetic transition disappears. Roughly at the same pressure, a local minimum of the \textit{n} parameter can be observed. We estimated the critical pressures (labeled $P _{1} $, $P_{2} $, $P_{3} $ and $P_{4} $ in each panel of figure \ref{LTfit}) from the pressure average of the estimated local maximum of \textit{A} and minimum of \textit{n}.
The critical pressures obtained this way with fits down to 1.8~K (cf figure \ref{LTfit}) were $P_{1} \approx$5.3~GPa (+/- 0.2~GPa), $P _{2} \approx$5.6~GPa (+/- 0.3~GPa), $P _{3} \approx$6.7~GPa(+/- 0.5~GPa) and $ P_{4} \approx$6.8~GPa(+/- 0.5~GPa). The error is due to the data spacing and the difference when $P_{c}$ is estimated from a fit of the \textit{T(P)} phase diagram \ref{PD}.
$\rho_{0}$ behaves similarly when the cell axis is along the \textit{a} or \textit{b} crystallographic axis, with a slight increase around the pressure where \Tc disappears and a stronger decrease above. When the cell axis is along the \textit{c} crystallographic axis, the behavior is different, with a continuous increase which is faster in $ \sim $ 5-7~GPa range of pressures, once \Tc decreases. For this orientation, the current is along the \textit{b}-axis, instead of \textit{c}, which may have a strong influence of the pressure dependence of $\rho_{0}$. The \textit{RRR} for $ \rho _{b,P//c} $ decreases from 7.2 at 0~GPa to 3.5 close to the critical pressure. This is in contrast with the \textit{RRR} in the two other directions which monotonically increases from 4-5 at 0~GPa to nearly 8 above 7.5~GPa.
Given the essentially complete $ \rho_{c,P//a} $ data set from our $ ^{3}$He run we can also try forcing the temperature exponent to be exactly equal to two at the lowest temperatures. Figure \ref{LTfitT2} presents the pressure dependence of \textit{A} and $\rho_{0}$ as well as the temperature range over which the $T^{2} $ fit to the data could be made. These results are consistent with those presented in Figure \ref{LTfit}: there is a divergence in \textit{A} near 5.1~GPa and the temperature range that the data can be fit with a quadratic pressure dependence drops below our minimum temperature between 5 and 5.5~GPa.
\section{Discussion}
\subsection{Anisotropy}
RVSb$ _{3} $ materials respond to chemical and applied pressure anisotropically. The lattice parameter decrease of RVSb$_{3}$ is anisotropic when R goes from La to Dy. Sefat \textit{et al.} \cite{Sefat08} found a decrease from 0.9\% to 5.4\% along the \textit{b} and \textit{a} axis, respectively. The thermal expansion of \CVS at ambient pressure (figure \ref{anisot}.b) is also clearly anisotropic and we deduced, from the Ehrenfest relation, a uniaxial pressure dependence anisotropy for \Tcf .
These observations motivated us to take advantage of the deviations from hydrostaticity in the modified Bridgman cell, to measure our samples with a slight additional uniaxial pressure component, successively applied along each of the three crystallographic axes.
We already observed from figure \ref{PD} and \ref{LTfit} that the critical pressure is different depending on the lattice direction along which the pressure is applied. This difference is significantly larger than any cell-to-cell variation. When the uniaxial stresses are applied along a stiffer axis, the crystal may be subject to smaller distortions and a higher pressure would be needed to suppress the magnetic transition. From this picture, the \textit{c}-axis can be considered as the least sensitive to the Bridgman cell uniaxial component and give results closest to the ones obtained in the more hydrostatic diamond anvil cell.
\begin{figure}[!ht]
\begin{center}
\includegraphics[angle=0,width=90mm]{fig9a.eps}
\includegraphics[angle=0,width=90mm]{fig9b.eps}
\end{center}
\caption{(Color online) Dependence on a scaled pressure of the parameters obtained from a low temperature fit $ \rho (T)= \rho _{0} + A T^{n} $ ($T \geq $1.8~K). The critical pressure values, $ P_{c} $, used here were given in the results section from the low temperature fits of the resistivity. a. \textit{A} coefficient. b. Temperature exponent, \textit{n}.}
\label{LTfit_param}
\end{figure}
Whereas the uniaxial pressure dependence anisotropy deduced from the Ehrenfest relation gave us a clue to measure our samples with the pressure applied along several different crystallographic axis, the predicted anisotropy was not retrieved from our measurements at the lower pressures. We do not observe any deviations between the \Tcf \textit{(P)} curves below 3~GPa. This might be due to relatively good hydrostaticity in this pressure range.
It is interesting to notice that whereas deviations in hydrostaticity tend to modify the pressure dependence of \Tcf , the low temperature functional dependence of the resistivity appears to be similar at comparable distances from the critical point. Indeed if we define an effective pressure parameter as $ \dfrac{P-P_{c}}{P_{c}} $ (with $ P_{c} $, critical pressure determined above) we can plot all of the \textit{A} and \textit{n} data on this universal scale (Figure \ref{LTfit_param}). The fact that both the \textit{A} and \textit{n} data fall onto common manifolds indicates that the quantum critical behavior is inherent to the system and only depends upon the distance from the critical value of the tuning parameter, pressure in this case. This result implies that slight uniaxial components of pressure may be utilized to further tune the criticality without fundamentally changing the underlying physics.
\subsection{Phase diagram}
When pressure is applied, the magnetic ordering temperature first increases, passing through a maximum before decreasing at a faster rate. No magnetic transition is observed for pressures above 7~GPa. This behavior is consistent with what we expect from the competition between the Kondo effect and the Ruderman-Kittel-Kasuya-Yosida interaction. The phase diagram (figure \ref{PD}) has a shape in very good agreement with the Doniach model \cite{Doniach77}.
A goal of this study was to determine the presence of a possible quantum critical point. Although the suppression of \Tc seems continuous, we can not clearly follow the transition for $T < $ 1.5~K. Even between 1.5~K and at least 4~K, in the modified Bridgman cell as well as in the diamond anvils cell, the peak used to infer the transition temperature is broad and its amplitude is small. This broader transition and lower amplitude may be an additional evidence of the progressive weakening of the magnetic transition once the pressure is high enough to reduce \Tcf , even in good pressure conditions. It has also to be mentioned that the transition broadening is in part related to the fact that above $ \sim $ 4~GPa, \Tcf\textit{(P)} line is becoming steeper with pressure. The $ \Delta $\Tc resulting from fixed experimental uncertainties will consequently increase. In the present case then, it is useful to evaluate the pressure evolution of the fit parameters $ \rho (T)= \rho _{0} + A T^{n} $ obtained at very low temperature to find further evidences for a quantum critical point.
As we already showed, it was found to be acceptable to fit only down to 1.8~K ($ ^{4} $He cryostat temperatures) at least to get a qualitative behavior. The results from the low temperature fits $ \rho (T)= \rho _{0} + A T^{n} $ presented in figure \ref{LTfit} are consistent with a presence of pressure induced quantum critical point. A sharp peak is observed in the \textit{A(P)} graph and the \textit{n(P)} graph drops sharply to $n \sim $1 as the critical pressure is approached. At low pressures, the \textit{n} exponent is above 2 as expected in the magnetic phase for a Kondo lattice system, and it tends to increase with the magnetic transition temperature. It then decreases until the critical pressure. At that point, \textit{n} is around 1.35-1.4, when measured in a $^{3} $He cryostat. This value is very close to 4/3, given by the spin fluctuation model in the case of a two-dimensional ferromagnet. However the lowest temperature obtained to determine \textit{n} was 0.35~K, which might be too high when close to the critical pressure. \textit{n} is then probably a little underestimated (from our estimations, \textit{n} tends to increase when the temperature decreases in this pressure range). The \textit{A} and $ \rho _{0} $ parameters appear to be less sensitive to the fit temperature range. $\rho _{0} $ slightly increases while approaching $P_{c} $, and then present a stronger decrease. However, its behavior is much different when pressure is applied along the \textit{c}-axis and current along the \textit{b}-axis, probably because of the resistivity anisotropy. It is interesting to notice that at higher pressures, far enough from the critical pressure, $\rho _{0} $ is even lower than at ambient pressure.
No superconductivity was observed in this compound down to the lowest temperature of 0.35~K reached in this work. This may be due to the ferromagnetic order, as no superconductivity was found in any other Ce-based ferromagnetic compounds such as CeNiSb$_{3}$\cite{Sidorov05Ce} or CeAgSb$_{2}$\cite{Sidorov03,Nakashima03}, presenting many similarities. Antiferromagnetic order is indeed known to be more propitious to superconductivity, compared to a ferromagnetism \cite{Mathur98}, and it has been shown that \textit{d}-wave singlet pairing in nearly antiferromagnetic metals is generally much stronger than \textit{p}-wave triplet pairing in nearly ferromagnetic metals \cite{Monthoux99}. On another hand, at least four U-based ferromagnetic compounds, which are all Ising-type ferromagnets, were found to be superconductors, at ambient pressure for URhGe \cite{Aoki01} and UCoGe \cite{Huy07}, or under pressure for UGe$_{2}$ \cite{Saxena00} and UIr \cite{Akazawa04}. The residual resistivity ratio of \CVS is low (below 10 over the whole pressure range) with a rather high residual resistivity, above 10 \mWf , compared to other superconducting Ce compounds. This may evidence a too strong scattering for the occurrence of exotic superconductivity. As an example, the residual resistivity should not be higher than a few \mW in the case of CePd$_{2}$Si$_{2}$ and CeIn$_{3}$ to observe superconductivity \cite{Mathur98}. On the other hand, for the ferromagnet CeAgSb$_{2}$, no superconductivity was observed in high quality samples with $\rho_{0}$ below 0.5~\mW \cite{Sidorov03,Nakashima03}.
Finally, the lowest temperatures reached of 0.35~K might also be too high to observe any eventual superconductivity.
\section{Conclusion}
We determined the pressure-temperature phase diagram of the ferromagnetic compound \CVSf. An initial increase of \Tc with pressure up to 4.5~GPa (for hydrostatic pressure medium) is observed, followed by the transition being progressively suppressed with further increase of pressure, in agreement with the Doniach model. From the extrapolation of \Tc to zero and the low temperature fits of the resistivity, we find a quantum critical point around 7~GPa. No superconductivity was observed down to 0.35~K. We took advantage of the uniaxial component in the modified Bridgman anvils cell and applied successively the pressure along the three axis. Discrepancies were noticed in the \Tc \textit{(P)} behavior when this slight uniaxial component is applied. The \textit{c}-axis seems to be stiff enough not be sensitive to the uniaxial component, and present a behavior in agreement with pressure conditions closer to hydrostaticity.
Whereas the modified Bridgman cell filled with Fluorinert was not suitable by itself to perform this study, it was shown to be very useful to evaluate the anisotropy in the uniaxial pressure dependence of the crystal. The use of 1:1 n-pentane:isopentane brought a strong improvement in pressure conditions and we are currently working to be able to consistently use it up to 8~GPa with the modified Bridgman cell.
\begin{acknowledgments}
This work was performed in part at Ames Laboratory, US DOE, under contract $\#$ DE-AC02-07CH11358 (E. Colombier, E.~D. Mun, S. L. Bud'ko, and Paul C. Canfield). This project has been supported by the French ANR programs DELICE and CORMAT (G. Knebel, and B. Salce). Part of this work was carried out at the Iowa State University and supported by the AFOSR-MURI grant $ \# $FA9550-09-1-0603 (X. Lin, and P.~C. Canfield). S. L. Bud'ko was also partially supported by the State of Iowa through the Iowa State University.
We would also like to acknowledge Stella Kim for her assistance with pressure cells measurements.
\end{acknowledgments}
\clearpage
|
\section{Introduction}
Strongly interacting matter at high temperature or large net baryon
number density is expected to undergo a rapid transition from a phase
with hadrons as dominant degrees of freedom to a phase where
partonic degrees of freedom prevail. At vanishing baryon chemical
potential ($\mu_B=0$), this transition is a true second order phase
transition only in the limit of vanishing light quark masses. For
$\mu_B > 0$, however, a second order phase transition point, the so-called
chiral critical point, may
exist also for physical values of the quark masses. A large experimental
as well as theoretical effort is put into the exploration of the QCD phase
transitions and the development of appropriate tools and observables
that can provide a univocal signal for the existence of the phase transitions
and their universal properties. The analysis of fluctuations of various
physical observables \cite{koch}, in particular of the net baryon
number, may serve this purpose \cite{ste,raj}. Theoretically, the
properties of the corresponding susceptibilities are well understood at
high temperature ($T$) and small values of the baryon chemical potential
($\mu_B$). In this regime, they are suitable observables for localizing the
phase boundary in the $\mu_B$-$T$ plane.
Critical behavior is signaled by long range correlations and
increased fluctuations, owing to the appearance
of massless modes at a second order phase transition.
Fluctuations of baryon number and electric charge
have been shown to be sensitive indicators for such critical behavior
\cite{lattice}.
In the exploration of the QCD phase diagram at non-zero temperature
and baryon chemical potential, higher order cumulants of baryon number
fluctuations play a particularly important role. They diverge on
the chiral phase transition line $T_c(\mu_B,m_q\equiv 0)$ as well as at
the elusive chiral critical point \cite{Stephanov}.
In heavy ion experiments, a lot of information has been collected on
particle yields in a wide range of beam energies \cite{data}. The
particle multiplicities are well described in a thermal model using
the partition function of a hadron resonance gas (HRG) \cite{HRG}.
Ratios of particle yields at a given beam energy can be
characterized by a few thermal parameters, e.g. temperature and chemical
potentials for baryon number, electric charge and strangeness.
These parameters define the freeze-out conditions, {\it i.e.} the thermal
parameters corresponding to the last interaction of the hadrons
participating in the collective expansion and cooling
of the hot and dense matter formed in a heavy ion collision. Data obtained at small values
of the baryon chemical potential suggest that the freeze-out curve $T_f(\mu_B)$
is close to the expected QCD phase boundary. In particular, at $\mu_B=0$
the chemical freeze-out seems to occur at or very near the QCD transition
region for a physical quark mass spectrum \cite{BraunMunzinger}.
At larger values of
$\mu_B/T$, however, there is a discrepancy between the slope of the
freeze-out curve and current lattice QCD results on the curvature of
the chiral phase transition line \cite{Mukherjee}.
The HRG model, which is based on the observed hadron spectrum,
does not exhibit critical behavior nor does it
reflect the sudden change of degrees of freedom in the
transition to the partonic phase of QCD. In the chiral limit, close to the
phase transition line $T_c(\mu_B,m_q\equiv 0)$ fluctuations of e.g. the
net baryon number density are expected to reflect the universal properties \cite{Allton}
of the 3-dimensional, $O(4)$ symmetric spin model \cite{Engels}.
The $O(4)$ scaling relations for cumulants of net baryon number fluctuations
differ significantly from predictions based on the HRG model.
Lattice calculations of cumulants of baryon number and electric charge
fluctuations, performed in the transition region at vanishing baryon chemical potential
and non-zero quark mass, do
indeed show that these cumulants differ qualitatively from those of the HRG model and reflect the
basic feature expected close to the chiral phase transition \cite{lattice}.
This suggests that at physical values of the light and strange quark masses
these cumulants are sensitive to the critical dynamics in the chiral limit
and may be employed to characterize also the
crossover transition in strongly interacting matter.
Thus, also at vanishing baryon chemical potential, {\it i.e.}
under the conditions approximately realized in the high energy runs at RHIC or LHC,
the question arises to what extent a refined analysis of the freeze-out
conditions can establish the existence of the chiral phase transition.
At $\mu_B/T \simeq 0$, net quark number fluctuations and their higher cumulants
can be computed within the framework of lattice QCD~\cite{lattice, Gavai:2008zr}. Such calculations will
eventually provide a complete theoretical characterization of the
thermal conditions in the crossover region. This may then be used to unravel
the relation of the freeze-out conditions at RHIC and LHC energies to the
pseudo-critical line in the QCD phase diagram, provided the system
remains close to thermal equilibrium during freeze-out.
At present, however, lattice calculations provide only limited
information on cumulants up to eighth order.
In particular, controlled predictions on their properties in the continuum limit are
still lacking; their characteristic features are obtained on a qualitative
level, but quantitative results are not yet available.
Viable alternatives for discussing qualitative
features of the net baryon number fluctuations is offered by O(4) scaling theory and by
chiral models, like e.g. the Nambu-Jona-Lasinio (NJL) model. In particular, the effective models have
the advantage
that they can be extended to $\mu_B > 0$ with minimal effort. On the other hand,
a clear disadvantage of NJL-type chiral models is that
they do not account for the potentially
large contribution from resonances in the hadronic phase.
In this paper we will discuss the robust features of cumulants
of net baryon number fluctuations that can be extracted from
considerations based on $O(4)$ universality, on existing lattice
calculations and on model calculations. In the next section we
discuss higher order cumulants at vanishing baryon chemical potential
making use mainly of $O(4)$ universality. In Section 3 we
extend these considerations to $\mu_B/T > 0$ using results from
model calculations. In Section 4 we summarize the relevance of our
findings for experimental studies of baryon number fluctuations at LHC
and RHIC and in Section 5 we give our conclusions.
\section{Charge fluctuations at \boldmath $\mu_B=0$}
\subsection{\boldmath $O(4)$ scaling functions and net baryon number
fluctuations}
Close to the chiral limit and at temperatures near the chiral phase
transition temperature $T_c$, higher order derivatives of the free
energy density ($f$) with respect to temperature or chemical potential
are increasingly sensitive to the non-analytic (singular) part
($f_s$). We may represent the free energy density in terms of the
singular and regular contributions
\begin{equation}
f(T,\mu_q,m_q) = f_s(T,\mu_q,m_q) + f_r(T,\mu_q,m_q) \; .
\label{freeenergy}
\end{equation}
In addition to the dependence on temperature $T$, we also introduce here an explicit
dependence on the light quark chemical potential, $\mu_q=\mu_B/3$,
and the (degenerate) light quark masses $m_q\equiv m_u = m_d$. For
simplicity we do not take the chemical potentials for electric charge and strangeness into account, nor do we introduce an explicit dependence on the
strange quark mass.
The singular part of the free energy may be written as
\begin{equation}
\frac{f_s(T,\mu_q,h)}{T^4} =
A h^{1+1/\delta} f_f(z) \ ,\ z \equiv t/h^{1/\beta\delta} \; ,
\label{singular}
\end{equation}
where $\beta$ and $\delta$ are critical exponents of the 3-dimensional
$O(4)$ spin model~\cite{Engels} and
\begin{eqnarray}
t &\equiv& \frac{1}{t_0}\left( \frac{T-T_c}{T_c} +
\kappa_q \left( \frac{\mu_q}{T}\right)^2
\right)
\ ,
\nonumber \\
h &\equiv& \frac{1}{h_0} \frac{m_q}{T_c} \ .
\label{scalingfields}
\end{eqnarray}
Here $T_c$ is the critical temperature in the chiral limit
and $t_0$, $h_0$ are non-universal scale parameters (as is $T_c$).
We use the chiral transition temperature $T_c$ also to set the
scale for the explicit symmetry breaking, introduced by the non-vanishing
light quark masses\footnote{In Ref.~\cite{Ejiri} the strange
quark mass was used to set the scale for the symmetry breaking term.}.
The amplitude $A$ is fixed by the relation of the scaling function
for the free energy density, $f_f(z)$, to the more
commonly used scaling function $f_G(z)$, which characterizes the
scaling properties of the chiral condensate, or in general the order
parameter ($M$) in $O(4)$ symmetric models, $M = h^{1/\delta} f_G(z)$,
where
\begin{equation}
f_G(z) = -\left( 1+\frac{1}{\delta} \right) f_f(z) + \frac{z}{\beta\delta}
f'_f(z) \; .
\label{fG}
\end{equation}
The scaling function $f_f(z)$ and its derivatives $f_f^{(n)}(z)$ have recently been
determined for $n\le 3$ using high precision Monte
Carlo simulations of the 3-dimensional $O(4)$ spin model and the known asymptotic
series expansions \cite{Engels_2011}.
We will use these results as a starting point
for a discussion of the generic structure of higher order cumulants of
the net baryon number fluctuations.
Note that the reduced temperature $t$, introduced in
Eq.~(\ref{scalingfields}), depends explicitly
on the quark chemical potential. The constant
$\kappa_q\simeq 0.06$, which controls the curvature of the chiral phase
boundary for small values of $\mu_q/T$, was recently determined in
a scaling analysis of (2+1)-flavor QCD \cite{Mukherjee}. A comparison
with other lattice results for $\kappa_q$~\cite{Endrodi,Philipsen}, suggests that
this parameter is only weakly dependent on the quark mass
and the number of flavors.
In this paper we focus on the properties of the net baryon number
fluctuations. The corresponding cumulants are obtained from Eq.~(\ref{freeenergy}) by taking
derivatives with respect to $\hat{\mu}_q = \mu_q/T$,
\begin{equation}
\chi_{n}^{B} = - \frac{1}{3^{n}}
\frac{\partial^{n}f/T^4}{\partial\hat{\mu}_q^n} \; .
\label{obs}
\end{equation}
From Eq.~(\ref{singular})
it is apparent that, in the vicinity of the critical temperature, the susceptibilities $\chi_{n}^{B}$ show a strong dependence on the explicit symmetry breaking term, the quark mass,
\begin{equation}
\chi_n^B \sim
\begin{cases}
-(2\kappa_q)^{n/2} h^{(2-\alpha -n/2)/\beta\delta} f_f^{(n/2)}(z)
& ,\ {\rm for}\ \mu_q /T = 0,\
{\rm and}\ n\ {\rm even} \\
- (2\kappa_q)^n \left( \frac{\mu_q}{T} \right)^n
h^{(2-\alpha -n)/\beta\delta} f_f^{(n)}(z)
&,\ {\rm for}\ \mu_q/T > 0\,
\end{cases}
\label{fluct_mass}
\end{equation}
where we used the scaling relation $2-\alpha = \beta\delta (1+1/\delta)$.
Because $\alpha$ is negative in the 3-dimensional $O(4)$ universality class
($\alpha = -0.2131 (34)$ \cite{Engels}), the specific heat, or equivalently
the fourth order cumulants of the net baryon number fluctuations,
does not diverge at the chiral transition
temperature, {\it i.e.} at $z=0$, in the chiral limit. At $\mu_q/T=0$ the first divergent cumulant
is obtained for $n= 6$, while at $\mu_q/T > 0$ this happens already for $n= 3$.
We note that the singular structure appearing
in $n$-th order cumulants for $\mu_q/T > 0$ is identical to that of $(2n)$-th order
cumulants at $\mu_q/T =0$, since
the chemical potential enters quadratically in the reduced
temperature $t$.
This characteristic has been used in Refs.~\cite{Asakawa, Skokov:2010}
to exploit properties of third order cumulants at non-zero baryon
chemical potential as signatures for critical behavior.
In this case, however, the singular contributions
are suppressed by a factor $(\mu_q/T)^{n/2}$
relative to the sixth order cumulants at $\mu_q/T =0$. Consequently, in the mean-field analysis of Ref.~\cite{Asakawa}, it was found that qualitative changes of the third-order cumulant, e.g. a change
of sign in $\chi_3^B$, is found only at rather large values of the chemical potential, $\mu_B/T > 4$.
Using Eq.~(\ref{fluct_mass}), one finds the leading
singularity in the chiral limit,
\begin{equation}
\chi_n^B \sim
\begin{cases}
-(2\kappa_q)^{n/2} |t|^{2-\alpha -n/2} f^{(n/2)}_\pm
& ,\ {\rm for}\ \mu_q /T = 0,\
{\rm and}\ n\ {\rm even} \\
-(2\kappa_q)^n \left( \frac{\mu_q}{T} \right)^n
|t|^{2-\alpha -n} f^{(n)}_\pm &,\ {\rm for}\ \mu_q/T > 0\; ,
\end{cases}\
\label{fluct}
\end{equation}
where
\begin{equation}
f^{(n)}_\pm = \lim_{z\rightarrow \pm \infty} |z|^{-(2-\alpha -n)} f_f^{(n)}(z)
\; .
\label{limit}
\end{equation}
The singular part of $\chi_4^B$, which is proportional to
$f_f^{(2)}$, has
the same singular structure as the specific heat; it is proportional to
the second derivative of the free energy with respect to temperature.
Thus, using the convention of Ref.~\cite{Engels_Cv}, we may write $\chi_4^B(t)$
in the chiral limit at $\mu_q/T=0$,
\begin{equation}
\chi_4^B(t) \sim \chi_{r} +\frac{A^\pm}{\alpha} |t|^{-\alpha} \; ,
\label{chi4}
\end{equation}
where $A^{+}$ is the amplitude above and $A^{-}$ below the critical temperature. The amplitudes $A^\pm$ are positive and the ratio $A^+/A^- \simeq 1.8$. This implies that
the cumulants $\chi_n^B(t)$ are positive for all $n > 4$ and $t<0$,
while for $t>0$ they alternate in sign. At non-zero values of the quark
mass, $h>0$, we thus expect $\chi_6^B$ to change sign in the transition
region and $\chi_8^B$ to do so twice.
For a given $h>0$, this is reflected
in the $z$-dependence of the scaling functions $f_f^{(n)}(z)$, as shown\footnote{
From the result for $f_f^{(3)}(z)$ \cite{Engels_2011}
we also constructed an estimate
for the next derivative, $f_f^{(4)}(z)$, which required some smoothening
of the interpolations that entered the determination of $f_f^{(3)}(z)$.
The resulting scaling function and the resulting quark mass dependence
of $\chi_8^B(t)$ is shown in Fig.~\ref{fig:c8}.
We want to use it here to point out the qualitative structure that
does arise within the $O(4)$ universality class, but do not consider
this figure as being correct on the quantitative level, {\it i.e.} as far
as the accurate location of the minima and the height of peaks is concerned.}
in Figs.~\ref{fig:generic} and \ref{fig:c8}.
In fact, the temperature and quark mass dependence of the singular parts of
the net baryon number fluctuations is directly related to the scaling functions $f_f^{(n)}(z)$ of the 3-dimensional $O(4)$ model \cite{Engels_2011} .
Thus, the generic structure of the fourth and sixth order cumulants can
be obtained from the known $O(4)$ scaling functions,
in the chiral limit as well as for non-zero values of the quark mass.
\begin{figure}
\begin{center}
\vspace*{-0.8cm}
\includegraphics*[width=6.7cm]{c4_O4_t.eps}
\includegraphics*[width=6.7cm]{c6_O4_t.eps}
\caption{
Scaling of the non-analytic contributions to
$\chi_4^B$ (left) and $\chi_6^B$ (right) arising from
second and third derivatives of the singular part of the free energy.
Shown are results for different values of the symmetry breaking
parameter $h_0h = m_q/T_c$; $h_0$ and $z_0=h_0^{1/\beta\delta}/t_0$
are non-universal scale parameters. Note that for $h_0h=1$ the
abscissa is the scaling variable $z$.
The corresponding curve thus directly shows the $O(4)$
scaling function.
}
\label{fig:generic}
\end{center}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics*[width=7.5cm]{c8_O4_t.eps}
\caption{Same as Fig.~\protect\ref{fig:generic} but for the
non-analytic contributions to $\chi_8^B$.
}
\label{fig:c8}
\end{center}
\end{figure}
In the chiral limit, the non-analytic contribution to $\chi_4^B$ vanishes
at the chiral transition temperature, $t=0$. Consequently, in the transition region,
the regular terms dominate in $\chi_4^B$ .
Nonetheless, the non-analytic term in $\chi_4^B$ varies rapidly with
temperature, leading to a pronounced maximum in the transition region,
observed in lattice as well as model calculations.
The temperature at which $\chi_6^B$ changes sign is non-universal since it depends on the
magnitude of the regular terms. However, in the scaling regime the location of the extrema and
the corresponding amplitudes follow universal scaling laws. Moreover, we note
that the positions of the two extrema of $f_f^{(3)}(z)$, $z^-<0$ and $z^+>0$, provide bounds on the
critical temperature in the chiral limit.
It is evident from Figs.~\ref{fig:generic} and \ref{fig:c8} that the $O(4)$
scaling functions
show much more structure and a stronger quark mass dependence in the
symmetric phase, $t>0$, than in the broken phase, $t<0$. In the latter
case, the divergence at $t=0$ builds up much more slowly than on the high
temperature side. The scaling function $f_f^{(3)}(z)$ changes sign
close to $z=0$, while $f_f^{(4)}(z)$ does so already
at $z^-< 0$. Furthermore, we note that the position of the maximum
of $f_f^{(3)}(z)$ is at $z^+\simeq 1.45$ \cite{Engels_2011}, which is close to
the peak position of the chiral susceptibility, $z_p\simeq 1.33(5)$
\cite{Engels}.
The location of these extrema define pseudo-critical temperatures $T_{6\pm}$,
which converge to that of the second order chiral phase transition
temperature $T_c$ in the chiral limit,
\begin{equation}
\frac{T_{6\pm}(m_q)}{T_c} = 1+ \frac{z^\pm}{z_0} \left( \frac{m_q}{T_c}\right)^{1/\beta\delta} \; ,
\label{T6}
\end{equation}
where $1/\beta\delta \simeq 0.55$. The proportionality constant $z_0$ is
uniquely determined in the scaling regime of QCD with two light quarks
\cite{Ejiri}
and is the same constant, which also controls the scaling of the
pseudo-critical temperature $T_\chi$ determined from the position of
the peak in the chiral susceptibility,
\begin{equation}
\frac{T_{\chi}(m_q)}{T_c} = 1+ \frac{z_p}{z_0} \left( \frac{m_q}{T_c}\right)^{1/\beta\delta}
\; .
\label{Tchiral}
\end{equation}
With decreasing quark mass the minimum of $\chi_6^B$ decreases as,
\begin{equation}
\chi^B_{6,min} \sim -\ \left( \frac{m_q}{T_c}\right)^{-(1+\alpha)/\beta\delta}
\simeq -\ \left( \frac{m_q}{T_c}\right)^{-0.66} \; .
\label{c6_min}
\end{equation}
We also note that corrections to $\chi^B_{6,min}$ at non-zero
$\mu_q/T$ start at ${\cal O}( (\mu_q/T)^4)$. We thus expect that the
basic structure of higher cumulants persists also for $\mu_B/T>0$.
\subsection{Sixth order cumulant and the QCD transition temperature}
Based on the generic structure of the $O(4)$ scaling functions we thus can
understand the basic features of the temperature dependence of higher
cumulants of the baryon number fluctuations at vanishing baryon chemical
potential. We focus on the properties of the sixth order cumulant,
$\chi^B_{6}(T)$, or correspondingly the ratio of cumulants
$R_{6,2}^B(T) = \chi^B_{6}(T)/\chi^B_{2}(T)$. In the hadronic phase,
$\chi^B_{6}(T)$ first grows with increasing temperature. It exhibits a
maximum in the
hadronic phase, close to the transition region and then drops rapidly. In the entire
high temperature regime, $\chi^B_{6}(T)$ and consequently $R_{6,2}^B(T)$ remain
negative. Lattice calculations of $\chi^B_{6}(T)$ \cite{Allton,Schmidt}
suggest that in QCD with physical quark masses, these
basic features, which are due to the singular terms in
$\chi^B_n$, persist. In particular,
$\chi^B_{6}(T)< 0$ in the vicinity of the pseudo-critical temperature
for chiral symmetry restoration.
We thus expect the sixth order
cumulant to be a very sensitive probe for the temperature at
which the freeze-out of hadrons in heavy ion collisions occurs.
In fact, it is conceivable
that hadronic freeze-out occurs in a temperature regime above
the QCD phase transition temperature in the chiral limit, {\it i.e.} it may
occur close to the pseudo-critical temperature that signals the onset of chiral
symmetry restoration. In that case we would expect to observe
\begin{equation}
R_{6,2}^B(T)\equiv \frac{\chi^B_{6}(T)}{\chi^B_{2}(T)} < 0
\; \text{at freeze-out at LHC and RHIC high energy runs.}
\nonumber
\label{prediction}
\end{equation}
This would be in striking contrast to conventional HRG model calculations,
which predict $R_{6,2}^B = 1$. A similar conclusion can be
drawn for the eighth order cumulant $\chi^B_8$ or, equivalently, the ratio
$R_{8,2}^B$.
\begin{figure}
\vspace*{-0.8cm}
\begin{center}
\includegraphics*[width=6.5cm]{c62_schematic.eps}
\vspace*{-0.6cm}
\caption{Schematic plot of the temperature dependence of the ratio
of the sixth and second order cumulants of the net baryon number fluctuations
in units of the phase transition temperature $T_c$ in the chiral limit.
The vertical lines show the chiral phase transition temperature and
the pseudo-critical temperature $T_{pc}$, corresponding to a peak in
the chiral susceptibility in QCD with physical light quark
masses.
}
\label{fig:schematic}
\end{center}
\end{figure}
In more general terms we suggest that a determination of $R_{6,2}^B$
at large collision energies at RHIC or LHC will provide a characteristic signature for
the location of the freeze-out temperature relative to the QCD phase
transition. This is illustrated in Fig.~\ref{fig:schematic}, where we show a schematic plot of $R_{6,2}^B$ along with the critical temperature in the chiral limit ($T_c$) and the crossover
temperature for chiral symmetry restoration for physical
light quark masses ($T_{pc}$).
In the following we extend the above considerations to the case of
non-vanishing baryon chemical potential. We do this in the framework
of a chiral model, analyzing the critical behavior of the
Polyakov loop extended quark meson (PQM) model in the functional
renormalization group approach \cite{Skokov:2010}. This allows us
to compute the dependence of higher cumulants on $\mu_q/T$ and $T$. We
determine the line of minima of $R_{6,2}^B$
in the $\mu_q$-$T$ plane and compare this with the
chiral transition line as well as with the pseudo-critical line obtained
at a physical pion mass.
\section{Higher cumulants of charge fluctuations at \boldmath $\mu_B/T > 0$}
The universal features of the 3-dimensional $O(4)$ scaling functions
discussed in the previous section also are reflected in effective chiral
model calculations where the Polyakov loop is coupled to
the fermion sector, thus generating many of the characteristics of the
confinement-deconfinement transition \cite{Fukushima,Fukushima:strong,Ratti,Sasaki:2006ww}.
The Polyakov loop extended quark meson (PQM) model~\cite{Schaefer:PQM} is one variant. It shares with QCD a global $O(4)$ symmetry, which
is spontaneously broken at low and restored at high temperatures.
Thus, in the light quark mass limit, this model reproduces the
universal scaling functions, discussed in the previous section.
We use the PQM model to implement the basic features
of the transition from hadronic matter to a quark gluon plasma into the
temperature dependence of the cumulants of net baryon number fluctuations and
to analyze their properties also at $\mu_q/T > 0$.
The Lagrangian of the PQM model \cite{Schaefer:PQM},
\begin{eqnarray}\label{eq:pqm_lagrangian}
{\cal L} &=& \bar{q} \, \left[i\gamma_\mu D^\mu - g (\sigma + i \gamma_5
\vec \tau \vec \pi )\right]\,q
+\frac 1 2 (\partial_\mu \sigma)^2+ \frac{ 1}{2}
(\partial_\mu \vec \pi)^2
\nonumber \\
&& \qquad - U(\sigma, \vec \pi ) -{\cal U}(\ell,\ell^{*})\ ,
\end{eqnarray}
involves interactions of mesons and gluons with fermionic fields.
The coupling between the effective
gluon field and quarks is implemented through the covariant derivative
\begin{equation}
D_{\mu}=\partial_{\mu}-iA_{\mu},
\end{equation}
where $A_\mu=g_s\,A_\mu^a\,\lambda^a/2$ and the spatial components of the gluon
field are neglected, i.e. $A_{\mu}=\delta_{\mu0}A_0$.
The purely mesonic potential of the model,
\begin{equation}
U(\sigma,\vec{\pi})=\frac{\lambda}{4}\left(\sigma^2+\vec{\pi}
^2-v^2\right)^2-c\sigma,
\end{equation}
contains a term linear in the $\sigma$ field, which explicitly breaks the chiral symmetry and is used to reproduce the physical pion mass.
The effective potential for the gluon field
is expressed in terms of the thermal expectation values of the color trace of the Polyakov loop and its conjugate,
\begin{equation}\label{pot}
\frac{{\cal U}(\ell,\ell^{*})}{T^4}=
-\frac{b_2(T)}{2}\ell^{*}\ell
-\frac{b_3}{6}(\ell^3 + \ell^{*3})
+\frac{b_4}{4}(\ell^{*}\ell)^2\,
\end{equation}
where
\begin{equation}
\ell=\frac{1}{N_c}\vev{\mathrm{Tr}_c\, L(\vec{x})},\quad \ell^{*}=\frac{1}{N_c}\vev{\mathrm{Tr}_c\,
L^{\dagger}(\vec{x})},
\end{equation}
and
\begin{eqnarray}
L(\vec x)={\mathcal P} \exp \left[ i \int_0^\beta d\tau A_4(\vec x , \tau)
\right]\,.
\label{pot_l}
\end{eqnarray}
Here ${\mathcal P}$ stands for the path ordering, $\beta=1/T$ and $A_4=i\,A_0$.
The potential ${\cal U}(\ell,\ell^{*})$ preserves the $Z(3)$ symmetry of the gluonic sector of QCD.
In the Appendix, we give further details on the choice of parameters for the
Polyakov loop potential and present the relevant calculational steps within
the functional renormalization group (FRG) approach
\cite{Wetterich, Morris,Ellwanger,Berges:review}. Within this formalism we compute
the free energy density of the PQM model
\cite{Skokov:2010,Skokov:2010wb},
\begin{equation}
f_{\rm PQM}(\ell, \ell^*;T, \mu) \equiv \Omega(\ell, \ell^*;T, \mu)
= -\frac{T}{V}\ln Z(\ell, \ell^*;T, \mu)\; ,
\label{fPQM}
\end{equation}
as a function of temperature,
chemical potential as well as the Polyakov loop variables, $\ell$ and $\ell^*$.
The latter two are then fixed by the stationarity condition:
\begin{eqnarray}
\label{eom_for_PL_l}
\frac{ \partial }{\partial \ell} \Omega(\ell, \ell^*;T, \mu) =0,~~
\frac{ \partial }{\partial \ell^*} \Omega(\ell, \ell^*;T, \mu) =0 \; .
\label{eom_for_PL_ls}
\end{eqnarray}
The cumulants of the net baryon number fluctuations are obtained
by taking suitable derivatives of the free energy, as specified in Eq.~(\ref{obs}). In practice
these derivatives have been implemented directly into the analysis
of the flow equations (see Appendix).
\begin{figure}
\begin{center}
\includegraphics*[width=6.5cm]{c_6.eps}
\includegraphics*[width=6.8cm]{c_8.eps}
\end{center}
\caption{The sixth and eighth order cumulants of the
net baryon number fluctuations at $\mu_q/T=0$ in the PQM
model. The temperature is given in units of the pseudo-critical
temperature $T_{pc}(m_\pi)$ corresponding to a maximum of the the chiral susceptibility.
The shaded area indicates the chiral crossover region.
}\protect\label{fig:PQM}
\end{figure}
\begin{figure}[t]
\begin{center}
\vspace*{-0.5cm}
\includegraphics*[width=6.5cm,]{c6c2.eps}\hspace*{0.3cm}
\includegraphics*[width=6.2cm,]{c8c2.eps}
\vskip 0.0cm\caption{The temperature
dependence of the fourth, sixth and eighth order cumulants of the net baryon
number fluctuations $\chi_{n}^{B}$ relative to the second order one. The
temperature is given in units of the chiral crossover temperature. The shaded
area indicates the region of the chiral crossover transition at $\mu_q/T=0$.
The calculations were done in the PQM model within the FRG approach.
}
\label{fig:PD0}
\end{center}
\end{figure}
\begin{figure}[t]
\begin{center}
\hspace*{-0.5cm}
\includegraphics*[width=6.5cm,]{c6c2mu02.eps}\hspace*{0.3cm}
\includegraphics*[width=6.5cm,]{c8c2mu02full.eps}
\vskip 0.0cm\caption
{
Comparison of the temperature dependence of the ratios $\chi_{6}^B$/$\chi_{2}^B$ and $\chi_{8}
^B/\chi_{2}^B$ for various $\mu_q/T$ corresponding to values at chemical
freeze-out in heavy ion collisions at RHIC.
The shaded
area indicates the region of the chiral crossover transition at $\mu_q/T=0$.
}
\label{fig:PD}
\end{center}
\end{figure}
In Fig.~\ref{fig:PQM} we show the sixth and eighth order cumulants of
the net baryon number fluctuations computed at $\mu_q/T=0$ within the PQM
model for physical values of the pion mass. The basic features dictated
by $O(4)$ symmetry restoration, as discussed in the previous sections,
are readily identified in the figure.
Moreover, the positions of the two extrema of $\chi_6^B$ correspond
approximately to the zeros
of $\chi_8^B$. This confirms that in the transition region, two derivatives
with respect to $\mu_q/T$ are indeed equivalent to one derivative with
respect to $T$.
From these calculations, as well as from calculations of the lower order
cumulants $\chi^B_2$ and $\chi^B_4$, we obtain the ratios $R_{n,m}^B$ of
the $n$-th and $m$-th cumulants.
Results obtained for $\mu_q/T=0$ and $\mu_q/T > 0$ are shown
in Figs.~\ref{fig:PD0} and \ref{fig:PD}, respectively. We note that these
ratios approach unity at low temperatures, as it is the case also in the
hadron resonance gas model. In the transition region, they reflect the
expected $O(4)$ scaling properties; they have a shallow maximum close
to the transition region before they drop sharply. In particular, they
show pronounced
minima with $R_{n,2}^B<0$ in the vicinity of the chiral crossover
temperature. The exact location of these minima and their depth is
to some extent model dependent. However,
we note, that in the transition region the second order cumulant
used in these ratios for normalization is dominated by non-singular
contributions which are positive.
The minima
in $R_{n,2}^B$ therefore mainly reflect the strong temperature dependence of
higher cumulants $\chi_{n}^B$. We also note that these minima become more pronounced
with increasing $\mu_q/T$. In fact, the structure of e.g. $R_{6,2}^B$
becomes similar to that of $\chi_{8}^B$ at large $\mu_q/T$. This is easily understood
in terms of the Taylor expansion of $R_{6,2}^B$, where
the dominant correction at non-zero $\mu_q/T$ is due to $\chi_{8}^B$ ,
\begin{eqnarray}
R_{6,2}^B(\mu_q/T) &=& R_{6,2}^B(0) +\frac{1}{2}\left( \frac{\mu_q}{T}
\right)^2 \left( R_{8,2}^B(0) -R_{6,2}^B(0) R_{4,2}^B(0) \right)
\nonumber \\
&&+ {\cal O}((\mu_q/T)^4) \; .
\label{Taylor}
\end{eqnarray}
This
also makes it clear why for $\mu_q/T >0$ the location of the minimum
of $R_{6,2}^B(\mu_q/T)$ is shifted to lower temperatures relative
to that of the chiral crossover temperature. Similarly, at non-zero $\mu_q/T$,
the ratio $R_{8,2}^B(\mu_q/T)$ shows more pronounced oscillations in the
transition region, due to contributions from higher order cumulants, which
oscillate more rapidly in the transition region. The amplitude of the maximum
at high temperatures becomes compatible in magnitude with that of the minimum.
\section{Freeze-out and the QCD transition}
The analysis of universal scaling functions that control the thermodynamics
in the vicinity of a phase transition in the universality class of
3-dimensional $O(4)$ symmetric theories (section 2),
of model (section 3)
and of lattice calculations
\cite{Schmidt,Cheng} suggest that at vanishing baryon chemical potential
the sixth order cumulant of the net baryon number fluctuations is negative
in the entire high temperature phase. In fact, the $O(4)$ scaling functions
for higher cumulants turn negative already in the vicinity of the chiral
($m_q=0$) critical temperature, {\it i.e.} below the crossover
temperature ($m_q>0$), which is relevant for the transition
at non-zero values of the light quark masses. The regular terms in the QCD
free energy may shift the onset of the negative regime to higher temperatures.
However, model \cite{Schaefer:2009st}
as well as lattice \cite{Schmidt,Cheng} calculations suggest that
the regime of negative sixth order cumulants starts below but close to the
QCD crossover transition temperature.
At non-zero baryon chemical potential, the temperature interval of negative
$\chi_6^B$, or equivalently $R^B_{6,2}(\mu_q/T)$, shrinks and
follows the crossover transition line. This is illustrated in
Fig.~\ref{fig:lines},
which shows the temperature interval, closest to the hadronic phase,
where the sixth
and eighth order cumulants of the net baryon number fluctuations are negative,
as obtained in the FRG approach to the PQM model.
It is evident that the sixth order cumulant $\chi_6^B$ is
negative in a wide range of temperatures which
extends into the symmetry broken phase. This is even more the case for
the eighth order cumulant as expected from the structure of the corresponding
$O(4)$ scaling function. Except for a small range of chemical potential
values close to $\mu_q/T = 0$, the eighth order cumulant is, however,
positive again on the crossover line.
\begin{figure}[t]
\begin{center}
\includegraphics*[width=6.5cm,]{negative_c6.eps}
\includegraphics*[width=6.5cm,]{negative_c8.eps}
\vskip -0.0cm\caption
{
The chiral crossover line [dashed line] and the first minima
in $\chi_6^B$ (left) and $\chi_8^B$ (right) [solid line]. The bands show
the parameter range for which $\chi_6^B$ and $\chi_8^B$, respectively,
are negative in the neighborhood of these minima.
}
\label{fig:lines}
\end{center}
\end{figure}
\section{Discussion and Conclusions}
We have shown that higher order cumulants of the net baryon number fluctuations
are sensitive probes for the analysis of freeze-out conditions in heavy ion
collisions and may allow to clarify their relation to the QCD phase transition.
This is the case at LHC energies as well as at the entire regime
of beam energies covered by the low energy run at RHIC. If in heavy ion
collisions, particles are produced from a thermalized system, the analysis
of higher cumulants of the net baryon number fluctuations does provide constraints on the
location of the freeze-out
temperature relative to the chiral transition temperature. We find that the
most robust statements, which become rigorous in the chiral limit,
can be made for the sixth order cumulants of charge fluctuations for small values of the quark chemical potential:
{\it If freeze-out occurs close to the chiral crossover temperature the
sixth order cumulant of the net baryon number fluctuations will be negative
at LHC energies as well as for RHIC beam energies $\sqrt{s_{NN}}\ \raise0.3ex\hbox{$>$\kern-0.75em\raise-1.1ex\hbox{$\sim$}}\ 60$~GeV,
corresponding to $\mu_B/T\raise0.3ex\hbox{$<$\kern-0.75em\raise-1.1ex\hbox{$\sim$}} 0.5$. This is in contrast to hadron
resonance gas model calculations which yield a positive sixth order
cumulant.}
We also note that the basic features discussed here for
net baryon number fluctuations also carry over one-to-one to
electric charge fluctuations. In the chiral limit, the most singular
component in the cumulants of electric charge fluctuations is proportional
to the singular part of the corresponding cumulant of
net baryon number fluctuations. In fact, lattice~\cite{Cheng} and
model~\cite{Fu:2009wy} calculations of the
sixth order cumulant of electric charge fluctuations at $\mu_q/T=0$
show that this cumulant is negative in the high temperature phase of
QCD. The ratio of the sixth and second order cumulants,
$\chi_6^Q/\chi_2^Q$ rapidly drops in the transition region from
the HRG value $(\chi_6^Q/\chi_2^Q)_{HRG}\simeq 10$ to about $-4$
at the chiral crossover temperature. Also this ratio is therefore a
sensitive probe of the conditions at freeze-out and
their relation to the critical behavior in strongly interacting matter.
We finally comment on the fourth order cumulants, in particular on the
ratio $\chi_4^B/\chi_2^B$ recently measured by STAR \cite{STAR} as well
as the corresponding ratio for cumulants for electric charge fluctuations,
$\chi_4^Q/\chi_2^Q$. The former is consistent
with the HRG value $(\chi_4^B/\chi_2^B)_{HRG}=1$ \cite{Karsch_2010,Gavai_2010}.
If freeze-out
occurs in the hadronic phase, one expects to find also large values
for ratios of cumulants of electric charge fluctuations,
$(\chi_4^Q/\chi_2^Q)_{HRG}\simeq 1.8$ \cite{Karsch_2010}. On the other hand,
if freeze-out occurs at the crossover temperature, also these ratios
will deviate from the HRG values. Lattice calculations suggests that
both ratios drop rapidly in the transition region, but stay positive
also in the high temperature phase \cite{Cheng}.
\begin{table}[t]
\begin{center}
\begin{tabular}{|l|c|c|c|c|}
\hline
~freeze-out conditions & $\chi_4^B/\chi_2^B$ &$\chi_6^B/\chi_2^B$ &
$\chi_4^Q/\chi_2^Q$ &$\chi_6^Q/\chi_2^Q$ \\
\hline
HRG& 1 & 1 & $\sim$ 2& $\sim$ 10 \\
QCD: $T^{freeze}/T_{pc}\raise0.3ex\hbox{$<$\kern-0.75em\raise-1.1ex\hbox{$\sim$}} 0.9$ & $\raise0.3ex\hbox{$>$\kern-0.75em\raise-1.1ex\hbox{$\sim$}} 1$ & $\raise0.3ex\hbox{$>$\kern-0.75em\raise-1.1ex\hbox{$\sim$}} 1$ & $\sim 2$&
$\sim 10$ \\
QCD: $T^{freeze}/T_{pc}\simeq 1$ & $\sim 0.5$ & $ < 0$ & $\sim 1 $ & $< 0$ \\
\hline
\end{tabular}
\end{center}
\caption{\label{tab:scenario}
Values for ratios of cumulants of net baryon number (B) and
electric charge (Q) fluctuations for the case that freeze-out appears
well in the hadronic phase (third row) or in the vicinity of the chiral
crossover temperature (fourth row). We give results
based on current lattice calculations \cite{Schmidt,Cheng} and on the calculations
presented here. In the second row we give results
of a HRG model calculation \cite{Cheng}. We also note that unlike the
cumulants of net baryon number fluctuations the ratios of cumulants of
electric charge fluctuations vary somewhat as a function
of the baryon chemical potential along the freeze-out line.
}
\end{table}
In Table~\ref{tab:scenario} we summarize two different freeze-out scenarios
characterized by various ratios of the cumulants of the net baryon number and electric
charge fluctuations, respectively. These scenarios assume that in heavy ion
collisions the freeze-out of hadrons occurs from a thermalized system
characterized by the phenomenologically determined freeze-out curve
$T_f(\mu_B^f)$. Table~\ref{tab:scenario} shows, that ratios of cumulants of
charge fluctuations are very sensitive to possible
differences in freeze-out and crossover temperature.
Finally, we note that these results do not account for possible finite volume
effects nor for possible effects of the evolution from chemical towards thermal freeze-out
in heavy ion collisions.
\section*{Acknowledgments}
We gratefuly acknowledge discussions with J\"urgen Engels on the $O(4)$
scaling functions.
The work of F.K. was supported in part by contract DE-AC02-98CH10886
with the U.S. Department of Energy, by the BMBF under grant 06BI401 and the
GSI Helmholtzzentrum f\"ur Schwerionenforschung under grant BILAER. K.R. acknowledges
partial support by the Polish Ministry of Science (MEN). B.F. and K.R. were supported in part by the
ExtreMe Matter Institute (EMMI). V.S. acknowledges support by the Frankfurt Institute for
Advanced Studies (FIAS).
\addcontentsline{toc}{section}{Acknowledgements}
|
\section{Introduction}
The main virtue of general-purpose Monte Carlo event generators
(sometimes called ``shower'' Monte Carlos, although they are normally
relied on for many other physics aspects as well) is their ability to
provide a complete and fully differential picture of collider final
states, down to the level of individual particles. This
allows them to be used as detailed --- albeit approximate ---
theoretical references for measurements performed at accelerators like the LHC,
against which models of both known and `new' physics can be tested.
The achievable accuracy depends both on the
inclusiveness of the chosen observable (with more
inclusive observables
generally being more precisely predicted) and on the
sophistication of the simulation itself. An important driver for the
latter is obviously the development of improved theoretical models,
e.g., by including matching to higher-order matrix elements, more
accurate resummations, or better non-perturbative models; but it
also depends crucially on the available constraints on the remaining
free parameters of the model. Using existing data to constrain these is
referred to as generator tuning.
The recent minimum-bias measurements from the LHC experiments, in
particular those at centre-of-mass energy $\sqrt{s} = 7$ TeV,
have highlighted the question to what extent the energy
scaling of total cross sections and differential distributions
are consistent with model-based extrapolations from lower energies, or
whether they exhibit any non-trivial departures from such
predictions.
Most of the LHC collaborations have already gone some way
towards studying this question, by including comparisons of specific
models and tunes to the data in their publications. In the short term,
such comparisons are useful both as immediate tests of commonly used
models, and to illustrate the current amount of theoretical
uncertainty surrounding a particular distribution. They also
provide a set of well-defined theoretical reference
curves for future studies. However, the conclusions that can be drawn
from comparisons of individual tunes of specific models on single
distributions are necessarily limited. In order to obtain more general
conclusions, a more coherent and over-arching look at both the data
and the models is needed.
In this study, we shall make use of the \textsc{Professor}\xspace tuning tool
\cite{Buckley:2009bj} to provide such a look. Specifically, rather than
performing one global fit to all the data, as is usually done, we
instead use \textsc{Professor}\xspace to perform several independent
optimizations of the model parameters for a range of different collider
energies. At each energy, we use the same set of minimum-bias
observables, modulo the limitations imposed by
different detector acceptances, trigger
conditions, and correction procedures.
We thereby seek to obtain a data-driven map of the preferred energy dependence
of each of the tuned parameters. This can then
be compared to the functional dependence assumed
by the underlying model, thereby furnishing not just a ``best fit'' of the model
parameters, but also a consistency check on the universality of the
underlying physics model itself.
We emphasize that this sort of consistency check is not limited to
energy scaling alone. In all generality, a consistency check
on the underlying physics model can be obtained by performing
independent optimizations in any two (or more) ``physics windows''
that the modeling provides or assumes relations between.
Moreover, with the recent advent of
automated tuning tools, it is now becoming possible to explore many
such independent optimizations with only modest investments
of computing and manpower.
In regions in which consistent parameter sets are obtained, with
predictions that are acceptably close to the data,
the model can be considered as interpolating well, i.e., it is universal.
If not, a breakdown in the ability of the model ability to span different
physical regimes has been identified, and can be addressed.
For simplicity, we concentrate on one particular model here, the
`new' interleaved multiple-interactions model
\cite{Sjostrand:2004pf,Sjostrand:2004ef}
implemented in the \textsc{Pythia 6} event generator \cite{Sjostrand:2006za}.
All parameters not explicitly subjected to optimization in this study
are those of the `Perugia 0' tune
\cite{Bartalini:2010su,Skands:2010ak}. We note that this model has
significant similarities with, but is not identical to, the one
implemented in \textsc{Pythia} 8 \cite{Sjostrand:2007gs}.
In section \ref{sec:setup}, we give brief overviews of the theory model, the
\textsc{Professor}\xspace/Rivet\xspace tuning framework, and the data sets we have used to do
the tuning. Section \ref{sec:energyscaling} contains our main
study of the energy scaling of three main model parameters at
currently existing colliders.
We round off with conclusions and outlook in section \ref{sec:conc}.
\section{Setup \label{sec:setup}}
In this section, we briefly describe the theoretical models we take as
starting points, emphasizing in particular
the parameters relevant to this study (section \ref{sec:model}), the tuning
framework we will be
using (section \ref{sec:professor}), and the data
sets that have been included (section \ref{sec:datasets}).
\subsection{Theoretical Model \label{sec:model}}
As mentioned above,
we consider the interleaved model of $p_\perp$-ordered showers and
multiple parton-parton interactions of \cite{Sjostrand:2004pf,Sjostrand:2004ef},
as implemented in \textsc{Pythia} 6.4.23 \cite{Sjostrand:2006za},
specifically the `Perugia 0' tune of that model
\cite{Skands:2010ak} unless otherwise specified.
In this class of models, pioneered by \cite{Sjostrand:1987su}, a
unified approach is taken to the modeling of all inelastic
non-diffractive events, in which dijet production and its associated
underlying event (UE) is viewed merely as the hard high-$p_\perp$ tail of
minimum-bias (MB), without any sharp modeling distinction
between the two. The fundamental building block for the model is the dijet
cross section, computed at leading order in perturbative QCD,
whose dominant component is simple low-$p_\perp$ Rutherford scattering
($t$-channel gluon exchange).
For $p_\perp\to 0$, i.e., when two partons scatter via the exchange
of a very soft gluon, this cross section
exhibits a divergence whose ultimate origin is similar to that of
initial-state bremsstrahlung. And similarly to what is done for
bremsstrahlung, the model recasts this divergence in terms of a
unitarized (Sudakov-suppressed) finite cross section for the hardest
scattering accompanied by a divergent \emph{number} of successively softer
multiple parton interactions (MPI), a number which is ultimately
regulated by the introduction of an infrared regularization scale. We do
not intend to give a full account of unitarization here, but instead
refer the reader to
\cite{Sjostrand:1987su,Sjostrand:2004pf,Sjostrand:2006za} for details
on it in the context of the MPI models here discussed. We also note
that an interesting exploration of the relation between
this effective scale and perturbative low-$x$ physics was recently
carried out
in \cite{Ryskin:2011qe}. Finally, a pedagogical
and more general discussion of underlying-event models in
general-purpose Monte Carlo generators can be found in the recent
review \cite{Buckley:2011ms}.
In this study, we focus on three main parameters of the resulting type
of model: the infrared
regularization scale, the proton transverse mass distribution, and the
color-reconnection strength, as follows:
\paragraph{Infrared Regularization Scale:} The fact that long-wavelength gluons
only see a coherent sum of the color charges in the hadronic
substructure --- color screening --- is assumed to ultimately regulate
the divergent number of parton-parton scatterings, similarly to how
the non-perturbative cutoff in parton showers regulates the
number of parton shower emissions. In the model we consider here,
a smooth regulator is introduced, by modyfing
the divergent parts of the cross section (including the strong
coupling since we use the standard MC scale choice
$\alpha_s(p_\perp^2)$) as follows,
\begin{equation}
\alpha_s(p_\perp^2) \ \frac{\mathrm{d}p_\perp^2}{p_\perp^4} \ \to \
\alpha_s(p_{\perp0}^2 + p_\perp^2) \
\frac{\mathrm{d}p_\perp^2}{(p_{\perp0}^2 + p_\perp^2)^2} ~,
\end{equation}
where $p_{\perp 0}$ physically expresses the scale at which the color screening
effect is supposed to become active. This parameter, which we call the
{\bf infrared regularization scale}, constitutes the main free
parameter for all models of this type, with low values yielding more
soft MPI activity (in the limit that it is taken to zero, the original
unregulated behaviour would be reobtained). In the \textsc{Pythia} model,
it is assumed to have a power-law scaling with the CM energy,
\begin{equation}
\label{eqn:ptmin}
p_{\perp 0}(\ensuremath{\sqrt{s}}\xspace) = \parp{82}\cdot
\left(\frac{\ensuremath{\sqrt{s}}\xspace}{\parp{89}}\right)^\parp{90}~,
\end{equation}
where \parp{82}, \parp{89}, and \parp{90} are tunable parameters.
Roughly speaking, \parp{82} gives the value of $p_{\perp 0}$
(in GeV) at a fixed reference CM energy = \parp{89} (also in GeV), and \parp{90}
determines the scaling behaviour of $p_{\perp0}$ away from that
energy. Below, instead of assuming the form, eq.~(\ref{eqn:ptmin}), we
shall fit for $p_{\perp0}$ independently at several different values
of $\sqrt{s}$. (Technically, we do this by fixing \parp{89} to the
energy of the relevant collider and fitting for \parp{82} which can
then be interpreted directly as $p_{\perp0}$ at that energy.) We can then
check whether the resulting points lie on a curve that is
consistent with the functional form of eq.~(\ref{eqn:ptmin}) or not.
To give the reader a more concrete idea of the dependence of the overall
event activity on the assumed scaling form, the left-hand pane of
Fig.~\ref{fig:th-pt0} illustrates the scaling
behaviour of
charged-particle multiplicities in non-diffractive minimum-bias
events\footnote{Specifically, the generated
events correspond to running \textsc{Pythia} in its ``minimum-bias''
mode with diffraction switched off. We permit ourselves this somewhat
unphysical definition here, since the illustration is intended for
qualitative purposes only. (For the numerical
studies later in this report, we use a full inelastic sample that
includes diffraction.)},
for two different
assumptions of the energy scaling of the $p_{\perp 0}$ parameter which we consider
comparatively extreme:
\begin{enumerate} \item
Solid lines: constant $p_{\perp 0}$, i.e., \parp{90} = 0.0, resulting
in a very fast growth of multiplicity with energy.
\item Dashed lines: \parp{90} = 0.32, i.e., $p_{\perp 0}$ varying
as $(\ensuremath{\sqrt{s}}\xspace)^{0.32}$, resulting in a
multiplicity growth with energy which is comparable to the
case without MPI, shown with dotted lines.
\end{enumerate}
For completeness, both
of the two \textsc{Pythia} min-bias models are included here, represented
by Tune A and Perugia 0, respectively. (For reference, the default for Tune A is a
scaling power of 0.25. For Perugia 0, it is 0.26.) In both cases, three
phase space regions are shown: inclusive (top), central (middle), and
central hard (bottom), with phase space cuts as indicated in the
grey shaded boxes.
\begin{figure}[t]
\vspace*{-0.1cm}\hspace*{-0.33cm}\includegraphics[scale=0.42]{pt0.pdf}\hspace*{-1.0cm}\includegraphics[scale=0.42]{pdf.pdf}\vspace*{-0.7cm}
\caption{Energy scaling of charged-particle multiplicities in
pp in three different phase space
regions (top: inclusive, middle: central, bottom: central hard).
{\sl Left:} Dependence on the scaling of the $p_{\perp 0}$ parameter for
two different \textsc{Pythia} models, represented by Tune A and
Perugia 0, respectively. The solid vertical line represents the reference energy,
1800 GeV, at which \ttt{PARP(82)} is defined for both models.
{\sl Right:} Dependence on the PDF set, for
the Perugia 0 model.
For reference, Tune A without MPI is also shown (dotted
lines).
\label{fig:th-pt0}}
\end{figure}
Comparing model curves with equal scaling assumptions (solid with
solid and dashed with dashed), it is evident that the two models have
somewhat different intrinsic scaling properties,
even with the same assumption for the
scaling of $p_{\perp 0}$. Thus, the value and scaling of $p_{\perp 0}$
alone is not sufficient to fix the energy scaling completely.
It is also worth noting that both models require \emph{at least one}
partonic scattering per hadron-hadron collision, hence it is
not possible for the average multiplicity to drop below the
``no-MPI'' case. This causes a rather abrupt change
in the scaling behaviour at low energies for those curves that
intersect the no-MPI (dotted) one.
On the right-hand pane of Fig.~\ref{fig:th-pt0}, we illustrate another
important dependence, on the PDF set used. Both Tune A and Perugia 0
were made using the CTEQ5L set \cite{Lai:1999wy} (solid
lines). Changing to either
CTEQ6L1 \cite{Pumplin:2002vw} (dashed lines) or MRST
LO** \cite{Sherstnev:2007nd} (dot-dashed lines) affects both the
value of the average
multiplicities as well as their scaling behaviour and distribution in
phase space. In particular, the LO** set generates a
significantly larger activity than its CTEQ cousins. It is therefore
not generally possible to separate the choice of PDF set from the
$p_{\perp 0}$ choice.
\paragraph{Transverse Mass Distribution:} A further important aspect
of the model is the shape of the assumed
proton matter distribution. In MPI models,
the probability for
additional parton-parton interactions to occur in a given collision is
proportional to the amount of matter overlap between the colliding beam
particles in that collision, which in turn depends on their impact
parameter, $b$. If the proton structure is very uniform (e.g., a
featureless pion/gluon cloud), the differences between peripheral and central
collisions will be quite small, while a strongly peaked distribution
(e.g., valence lumps / hot spots) can make the activity in
central collisions much higher than in peripheral ones. Thus,
while we may think of the infrared regularization scale above as
determining the \emph{average} number of multiple parton interactions,
the $b$ profile affects how much this number can \emph{deviate}
from the mean in peripheral vs.\ central events.
In the overlap model used for the
Perugia tunes, the overlap function (the time-integrated convolution
of two proton mass distributions, see
\cite{Sjostrand:1987su,Sjostrand:2004pf}) is cast as
\begin{equation}
{\cal O}(b) \propto \exp\left(-b^d\right)
\end{equation}
with the power $d$ a free parameter whose range is normally taken to
be from $d=1$ (exponential, representing a very peaked structure)
to $d=2$ (Gaussian, representing a smooth structure). Note that the
normalization of this distribution is fixed to unity. Note also that
$b$ is given in an arbitrary unit; since the only dimensionful
quantity is the total cross section, which is fixed by a
Donnachie-Landshoff formula \cite{Donnachie:1992ny}, the $b$ shape
does not affect the total cross section at all in this type of model,
and only the dimensionless ratio $b/\left<b\right>$ appears in the
explicit calculations\footnote{
For completeness, we note that,
while there is thus formally a dependence on the
overall proton-proton impact parameter $b$ in
the model, there is no actual space-time representation of the
collision, and hence no dependence on the direction of $b$ nor on the
individual parton-parton impact parameters. }.
The power, $d$, appears as the parameter \parp{83} in
\textsc{Pythia}\footnote{Strictly speaking, this form of the matter
profile is only selected for
\ttt{MSTP(82)=5}. See the \textsc{Pythia} documentation on
\ttt{MSTP(82)} for how to select other matter profiles,
such as the double-Gaussian one \cite{Sjostrand:2006za}.}.
It is
not assumed to change with energy, i.e.,
\begin{equation}
d(\sqrt{s}) = \parp{83} ~.
\end{equation}
By making separate tunes at each energy individually, we will obtain a
data-driven test of the validity of this assumption.
Since all expressions are cast in terms of the
dimensionless ratio of the impact parameter relative to its average,
the assumed shape also does not greatly affect the \emph{average} event
activity. The main consequence of different $b$ profiles thus lies in the
\emph{shape} of distributions, with a smooth matter profile generating
narrower ones than more lumpy profiles, a consequence of the latter
allowing for larger
event-to-event fluctuations. The fact that the
average multiplicity is not greatly affected by the choice of
impact parameter profile is illustrated on the left-hand pane of
Fig.~\ref{fig:th-b},
\begin{figure}[t]
\vspace*{-0.1cm}\hspace*{-0.33cm}\includegraphics[scale=0.42]{bdep.pdf}\hspace*{-1.0cm}\includegraphics[scale=0.42]{cr.pdf}\vspace*{-0.7cm}
\caption{Energy scaling of charged-particle multiplicities in
pp in three different phase space
regions (top: inclusive, middle: central, bottom: central hard).
{\sl Left:} two different impact parameter profiles. {\sl Right:}
three different color-reconnection strengths.
For reference, Tune A without MPI is also shown (dotted
lines). For all other curves, the parameters of Perugia 0 were used,
except for the modifications indicated on the plots.
\label{fig:th-b}}
\end{figure}
where the $b$ dependence of the Perugia 0 tune has been varied between a
Gaussian and an Exponential overlap distribution without substantially
altering neither the average values nor their scaling with energy.
\paragraph{Color Reconnection Strength:} The last main aspect of the
modeling we shall be concerned with here is
the strength of the color reconnections (CR) that are used to model the
collapse of the color wave function in the final state. The
so-called `color annealing' models employed by the Perugia tunes
were described in detail in
\cite{Sandhoff:2005jh,Buttar:2006zd,Skands:2007zg,Skands:2010ak} and are
qualitatively similar to the Generalized-Area-Law (GAL) models
developed earlier by the Uppsala group \cite{Rathsman:1998tp}. Briefly summarized,
each MPI corresponds to one or two color exchanges between the
beams (depending on whether a quark or a gluon is exchanged,
respectively).
In the $N_C\to\infty$ limit used to represent color topologies
in MC generators, every such exchange must be neutralized at
the hadronization stage. In \textsc{Pythia}'s case this is modeled
by the formation of
strings, which subsequently hadronize to produce observable
particles.
In the most naive $N_C\to\infty$ treatment, each such string would be
completely independent of the others. However, since the number of
real-world colors is finite, $N_C=3$, and since
the strings generated by MPI all traverse the same rapidity region
(between the remnants) there is some reason to suppose that the collapse of the
color wavefunction is instead more complicated and/or that
the strings after formation interact to fuse or cut each other
up. In the string picture, such effects should be driven
by a minimization of the total space-time area spanned by the strings
(the so-called area law for classical strings, as measured, e.g.,
by the $\lambda$ measure \cite{Andersson:1983jt,Andersson:1985qr}). Even without
understanding the dynamics in detail, we may therefore reasonably suspect that
the end result will be shorter string pieces, which in turn
will produce
fewer, but more energetic, particles, i.e., a harder fragmentation
spectrum.
At least within the $p_\perp$-ordered \textsc{Pythia} 6
modeling, some such mechanism does appear to be empirically
necessary in order to properly describe the observed increase of the
mean $p_\perp$ of
charged tracks with track multiplicity in min-bias events
\cite{Aaltonen:2009ne,Aad:2010rd}.
The annealing models developed in
\cite{Sandhoff:2005jh,Buttar:2006zd,Skands:2007zg,Skands:2010ak} are all formulated
in terms of one main
parameter: the basic color-reconnection/string-interaction
strength, $\xi_R$, given by \parp{78} in the code. The larger this parameter
is, the stronger the reconnection effect, and the faster the rise
of $\left<p_\perp\right>(N_{ch})$. However, since these models were only
intended as crude toy models, nothing has so far been said as to their
possible dependence on the energies of the colliding beams. The only
scaling built into the models is thus a rough scaling with the number
of MPI in an event, or in the most detailed variant (so far used only for the
Perugia 2010 and Perugia K tunes \cite{Skands:2010ak}) the number of
overlapping string pieces in each rapidity region. The fundamental
reconnection probability is assumed constant, i.e.,
\begin{equation}
\xi_R(\sqrt{s}) = \parp{78}~.
\end{equation}
Again, by making separate tunes at each energy individually, we will obtain a
data-driven test of the validity of this assumption.
The consequence of varying \parp{78} from zero to one is illustrated
in the right-hand pane of Fig.~\ref{fig:th-b}. We observe that the
average multiplicity at each energy can be modified by up to a factor
of 2 by this particular colour-reconnection model,
but note also that the relative scaling \emph{between} energies stays
virtually independent of $\xi_R$. Nonetheless, since any model with,
for instance, an energy-dependent $\xi_R$ would interpolate between our
curves --- leading to a different effective energy dependence ---
we must still conclude that the energy scaling of other models
could be qualitatively different from this.
Finally, we note
that the CR model employed by the Perugia 0 tune actually depends on one
more parameter, \parp{77}, which acts to suppress reconnections among
high-$p_\perp$ string pieces. Since this parameter is tightly
correlated with \parp{78} and since it only affects details of the
high-$p_\perp$ tail of the $p_\perp$ distribution, we have kept it
fixed to its Perugia-0 value in this study. However, we did check that
allowing its value to change or even fixing it to zero did not
qualitatively alter any of our conclusions.
\paragraph{Remarks on Diffraction:}
Finally, we should emphasize a point that is especially relevant for
the modeling of low-multiplicity minimum-bias collisions. As
mentioned above, the model we have outlined here attempts primarily
to describe
inelastic, non-diffractive events.
It would therefore have to be complemented by a separate
modeling of the diffractive parts of the cross section, to the extent that the
measurements are sensitive to these, as, e.g., in low-multiplicity
minimum-bias events. In this study, we use the diffractive modeling
provided by the \textsc{Pythia} 6 generator itself. We note that this
modeling does not include diffractive jet production and is presumably
not reliable for high-$p_\perp$ and/or high-mass diffractive particle
production. We therefore investigate the energy scaling of the
resulting combined min-bias model for two subsamples of events --- one that
includes the low-multiplicity region, and one that excludes it, as
will be discussed in the description of the data sets below.
\subsection{Tuning Framework \label{sec:professor}}
The \textsc{Professor}\xspace tuning framework~\cite{Buckley:2009bj} relies on the construction of a
fast analytic model of the generator by bin-wise parameterizations of the
generator's response to shifts in parameter-space. These
parameterizations are performed within a hypercube of user-specified
volume inside the generator parameter space. (It is up to the user to
make sure that the boundaries of this hypercube correspond to
meaningful generator settings.) The comparison to experimental
analyzes is performed via the Rivet\xspace \cite{Waugh:2006ip,rivet} analysis tool,
which in turn relies on the \textsc{HepMC} \cite{Dobbs:2001ck} event
record format and on the \textsc{HepData} repository \cite{Buckley:2010jn}.
An important point is to choose analyzes in Rivet\xspace that contain
observables that are sensitive to the tuning-parameters. The \textsc{Professor}\xspace system
contains tools that help to identify those observables that are not sensitive
to the parameters in question (\ttt{prof-sensitivities}) and to confirm that
the sampling-hypercube is chosen in such a way that the Monte-Carlo-generator
runs enclose the experimental data (\ttt{prof-envelopes}).
The actual tuning stage consists of a numerical minimization of a
goodness-of-fit (GOF) measure
constructed from the parameterizations $f^{(b)}(\ensuremath{\vec{p}}\xspace)$, experimental data $\mathcal{R}_b$
and most important a weight \ensuremath{w_{b}}\xspace for each bin of a set of observables to tune to:
\begin{align}
\ensuremath{\chi^2}\xspace(\ensuremath{\vec{p}}\xspace) =
\sum_{\mathcal{O}} \sum_{b \, \in \, \mathcal{O}}
\ensuremath{w_{b}}\xspace\cdot\frac{ (f^{(b)}(\ensuremath{\vec{p}}\xspace) - \mathcal{R}_b)^2 }{ \Delta^2_b },
\label{eq:chi2}
\end{align}%
The bin-weights \ensuremath{w_{b}}\xspace can be seen as the main user-input to the tuning-stage; they
help to emphasize or exclude certain regions of an observable in the
numerical fitting procedure. The return value of the fit will be a point \ensuremath{\vec{p}}\xspace that
minimizes the GOF given in \EqRef{eq:chi2}. This parameter set may then
be subjected to explicit validation in terms of comparison of the
observable as predicted from the parameterizations at \ensuremath{\vec{p}}\xspace
with the experimental data. Finding a set of weights that leads to a
satisfying description of observables usually requires some iteration.
For each collider energy, we perform a ``local'' tune that includes
only data from that particular energy. For these, we manually set
\parp{89} equal to the given energy and let \textsc{Professor} optimize for
\parp{82}, which can then be interpreted directly as a $p_{\perp 0}$
value at that energy. We used the Perugia 0 parameter settings as center of our
parameter sampling hypercube, and used the Perugia 0 energy scaling to
define the sampling range for \parp{82} at each energy.
An overview of the parameter sampling-ranges is given in
\TabRef{tab:ranges}.
\begin{table}[t]
\centering
\scalebox{0.9}{
\begin{tabular}{l*{5}{c}|c}
\toprule
\ensuremath{\sqrt{s}}\xspace & 630 \,\mbox{Ge\kern-0.2exV}\xspace & 900 \,\mbox{Ge\kern-0.2exV}\xspace & 1.8 \,\mbox{Te\kern-0.2exV} & 1.96 \,\mbox{Te\kern-0.2exV} & 7 \,\mbox{Te\kern-0.2exV} & Global fit\\
\midrule
\parp{82} & 1.066 \ldots 1.979 & 1.169 \ldots 2.171 & 1.4 \ldots 2.6 & 1.431 \ldots 2.658 & 1.993 \ldots 3.702 & 1.0 \ldots 3.0\\
\midrule
\parp{78} & \multicolumn{6}{c}{0.0 \ldots 0.7} \\
\parp{83} & \multicolumn{6}{c}{1.0 \ldots 2.0} \\
\midrule
\parp{90} & &&&& & 0.16 \ldots 0.34 \\
\bottomrule
\end{tabular}}
\caption{Parameter sampling-ranges used for the MC generator runs at each energy.}
\label{tab:ranges}
\end{table}
We also perform a ``global'' tune, using data from all energies
simultaneously, with an approximate relative weighting that attempts
to take into account that the amount of data --- and hence the
statistical power --- at each energy varies. The number of events
contained in each data sample we used is listed in
\TabRef{tab:nevt}. (These data sets will be discussed in more detail in
\SecRef{sec:datasets}.) In order to take into account the possibility
that several measurements of the same observable may have been
performed at closely spaced energies (e.g., Tevatron Run I and II),
we define an effective total number of events for each observable, for each
collider energy, as follows:
\begin{equation}
N_i^{\mrm{eff}} = \sum_j N_j \, e^{-r_{ij}^2/(2\sigma^2_E)}~,
~\label{eq:Neff}
\end{equation}
where $j$ runs over all included measurements of the given observable
at all energies, $r_{ij}=\log_2(E_i/E_j)$ provides a
logarithmic measure of the distance between two energies,
and we have chosen an ``energy resolution parameter'' of $\sigma_E =
1/3$ so that energies spaced a factor of 2 or more apart correspond to
being $3\,\sigma$ away from each other and will therefore effectively
contribute independently, while measurements closer than a
factor $2^{1/3}\sim1.25$ in energy will blend into each
other, being resolved by less than $1\,\sigma$.
(We note that this parameter could be
varied to help estimate the uncertainty on the tuning, but the
question of more rigorous uncertainties is not a simple one and
reaches beyond the scope we aim to address here.)
The effective event numbers computed in this way are
given, for each observable, in the two rightmost columns of
\TabRef{tab:nevt}, with the figures below illustrating the effective
contributions of each sample to the total at each energy.
The most naive weight normalization would be to simply let all the
samples enter with unit weights. This would unavoidably
bias the fit towards the energies at which most statistics has been
collected. Using $N_i^\mrm{eff}$, we may instead attempt to normalize the
statistical power in such a way as to force each energy to enter with
approximately equal weight, regardless of the amount of statistics
collected. This could be achieved by weighting each
sample by
\begin{equation}
w_i^{\mrm{eff}} = \frac{\max(N_i^\mrm{eff})}{N_i^{\mrm{eff}}}~,
\end{equation}
which we refer to as ``linear'' reweighting. In this scheme, two
event samples at widely spaced energies, containing for instance
10k and 1M events, respectively, would receive weights 100 and 1,
respectively. I.e., the power of the
measurement with the \emph{lowest} statistics would be artificially
enhanced, in order for the global fit not to be totally dominated by
the higher-statistics one. This strategy should be considered
extreme, however, since it grants no benefit at all to measurements
performed with superior statistics, and one therefore risks being overly
sensitive to noise in the poorly measured ones. As an intermediate
compromise, we propose to let the samples enter with relative weights
\begin{equation}
w_i^{\mrm{eff}} = \sqrt{\frac{\max(N_i^\mrm{eff})}{N_i^{\mrm{eff}}}}~.
\end{equation}
such that the samples in the example above would enter with weights
10 and 1, rather than 100 and 1. We refer to this as ``square root''
reweighting.
Obviously, we do not intend this to define a rigorous procedure,
but view it as a first attempt at highlighting and addressing the disparate
statistical powers available in the various sets.
\begin{table}[t]
\centering
\begin{tabular}{lcr|rr}\toprule
& & & \multicolumn{2}{c}{ --------------------- \
$N^\mrm{eff}_i$ \ ---------------------} \\
Analysis & $\sqrt{s}$~[GeV] & $N_\text{evt}$ &$P(N_\mrm{ch})$ \&
$\left<p_T\right>(N_\mrm{ch})$ & $dN_\mrm{ch}/dp_\perp$ \\
\midrule
CDF 1988\cite{Abe:1988yu} & 630 & 9400 & & 47k\\
CDF 2002\cite{Acosta:2001rm} & 630 & 1963157 & 2.0M\\
UA5 1989\cite{Ansorge:1988kn} & 900 & 1189 &
0.8M\\
ATLAS 2010\cite{ATLAS:1266235} & 900 & 124782 & 0.8M & 162k\\
CDF 1988\cite{Abe:1988yu} & 1800 & 55700 & & 9.2M \\
CDF 2002\cite{Acosta:2001rm} & 1800 & 2079558 & 11.2M\\
CDF Run-\textsc{II}\cite{Aaltonen:2009ne,Run2:MBnote} & 1960 & 9788000 & 11.7M & 9.8M\\
ATLAS 2010\cite{ATLAS:1266235} & 7000 & 5395000 & 5.4M & 5.4M\\
\midrule
$\max(N_i^\mrm{eff})$ & & & \multicolumn{2}{c}{11.7M} \\
\bottomrule
\end{tabular}\\[3mm]
{\footnotesize
\begin{tabular}{rr}
\multicolumn{1}{l}{\hspace*{0.75cm}$N_i^\mrm{eff}$ for $P(N_\mrm{ch})$
\& $\left<p_T\right>(N_\mrm{ch})$} &
\multicolumn{1}{l}{\hspace*{0.75cm}$N_i^\mrm{eff}$ for $dN_\mrm{ch}/dp_\perp$} \\[-0.4cm]
\includegraphics[scale=0.84]{neff-nch.pdf}&
\includegraphics[scale=0.84]{neff-pt.pdf}\\
$\sqrt{s}$ [GeV] & $\sqrt{s}$ [GeV]
\end{tabular}}
\caption{{\sl Top Row:} Number of events contained in each data sample used, and
the effective event numbers, $N_i^\mrm{eff}$,
for each observable, for each sample used for that
observable. {\sl Bottom Row:} illustration of the
relative sizes of each of the samples and their
contributions to $N_i^\mrm{eff}$, according to
eq.~(\ref{eq:Neff}), for {\sl Bottom Left:} $P(N_\mrm{ch})$ \&
$\left<p_T\right>(N_\mrm{ch})$ and {\sl Bottom Right:} $dN_\mrm{ch}/dp_\perp$.}
\label{tab:nevt}
\end{table}
As a final comment, we note that
a certain freedom exists for the parameterization of the generator response
which allows for systematic checks of the validity of obtained tuning results
which can also be turned into typical spreads which we use as rough uncertainty
estimates in this study. These estimates and their properties are
described in greater
detail in \cite{Buckley:2010nk}. We include them as light shaded
(cyan) bands on the plots in the following subsections, added in
quadrature with the ordinary $\chi^2$-based fit uncertainties computed
by Minuit, which are shown as darker shaded bands.
\subsection{Data Sets \label{sec:datasets}}
To study the energy dependence of parameters properly, we need to make
sure that the experimental data at each energy has either
been corrected for detector effects in a comprehensible way or that
the uncorrected data is presented with enough information on
efficiencies so that it can be
used with Rivet\xspace. Further, the distributions we choose must be both sensitive to
the parameters we want to tune and, simultaneously, consistently available
at all collider energies with phase-space regions and trigger
conditions not too different from one to the next.
A minimal set of minimum-bias distributions that matches these requirements
for our selection of tuning parameters and for energies ranging from
630 \,\mbox{Ge\kern-0.2exV}\xspace\ to 7\,\mbox{Te\kern-0.2exV}\ is:
\begin{itemize}
\item The charged-particle multiplicity-distribution
($N_\text{ch}$)
\item The charged-particle \ensuremath{p_\perp}\xspace-distribution
\item The charged-particle $\langle \ensuremath{p_\perp}\xspace \rangle$
vs. $N_\text{ch}$-distribution
\end{itemize}
We would have liked to extend the reach of our study by including data
from the RHIC experiments but were unable to find any published
(corrected) data for the $\langle \ensuremath{p_\perp}\xspace \rangle$ vs. $N_\text{ch}$
distribution at 200 GeV. The same is true for older experiments such as
SFM, where multiplicities and \ensuremath{p_\perp}\xspace -distributions were measured in pp-reactions
at 62 \,\mbox{Ge\kern-0.2exV}\xspace and lower. As has recently been emphasized \cite{Wraight:2011ej},
it would also have been interesting to add forward-backward
correlations to the list of variables, since these are particularly
sensitive to the mix of short- vs.\ long-distance processes, but
in the energy range we consider, only UA5 has so far reported
measurements \cite{Ansorge:1988fg}, with numbers at Tevatron and LHC
energies not (yet) available.
\begin{table}[t!]
\centering
\scalebox{1.0}{\small
\begin{tabular}{lrrlccc|crr}
\toprule
& $\sqrt{s}$ & & & &
$p_{T\mrm{min}}$ & \multicolumn{2}{c}{Fitrange} &\multicolumn{2}{c}{Weight} \\
Exp/Trig & \hspace*{-3mm}[GeV] & Ref & Observable & \hspace*{-1mm}$|\eta|_{\mrm{max}}$\hspace*{-1mm} &
[GeV] & $N_{\mrm{ch}}\ge 1$ & $N_{\mrm{ch}}\ge 6$ & \hspace*{-2mm}local & global \\
\midrule
\textbf{CDF}/TC1 & 630 & {\cite{Abe:1988yu}} &
$dN/dp_\perp$ & $1.0$ & $0.4$ & $0.4$ -- 3 & $0.4$ -- 3 &1.0 & 15.8\\
\textbf{CDF}/TC1 & 630 & \cite{Acosta:2001rm} &
$P(N_{\mrm{ch}})$ & $1.0$ & $0.4$ & 1 -- 16 & 6 -- 16 & 1.0 & 2.4\\
&&&$P(N_{\mrm{ch}})$ & $1.0$ & $0.4$ & 16 -- 30 & 16 -- 30 & 5.0 & 12.1\\
\textbf{CDF}/TC1 & 630 & \cite{Acosta:2001rm} &
$\langle p_\perp \rangle$ vs.\ $N_{\mrm{ch}}$ & $ 1.0$ & $0.4$
& 1 -- 16 & 6 -- 16 & 1.0 & 2.4\\
&&&$\langle p_\perp \rangle$ vs.\ $N_{\mrm{ch}}$ & $ 1.0$ & $0.4$
& 16 -- 30 & 16 -- 30 & 5.0 & 12.1\\
\midrule
\textbf{UA5}/TU1 & 900 & \cite{Ansorge:1988kn} &
$P(N_{\mrm{ch}})$& $0.5$ & $0.0$ & 1 -- 19 & 6 -- 19 & 1.0 & 3.8\\
\textbf{UA5}/TU1 & 900 & \cite{Ansorge:1988kn} &
$P(N_{\mrm{ch}})$ & $0.5$ & $0.0$ & 19 -- 30 & 19 -- 30 & 5.0 & 19.1\\
\midrule
\textbf{ATL}/TA$^*$ & 900 & \cite{ATLAS:1266235} &
$dN/dp_\perp$ & $2.5$ & $0.5$ & 0.5 -- $10.0$ & 0.5 -- 10.0 & 1.0 & 8.5\\
\textbf{ATL}/TA$^*$ & 900 & \cite{ATLAS:1266235} &
$P(N_{\mrm{ch}})$ & $2.5$ & $0.5$ & 1
-- 25 & 6-- 25 & 1.0 & 3.8\\
\textbf{ATL}/TA$^*$ & 900 & \cite{ATLAS:1266235} &
$P(N_{\mrm{ch}})$& $2.5$ & $0.5$ & 25 -- 60 & 25 -- 60 & 5.0 & 19.1\\
\textbf{ATL}/TA$^*$ & 900 & \cite{ATLAS:1266235} &
$\langle p_\perp \rangle$ vs.\ $N_{\mrm{ch}}$ & $ 2.5$ & $0.5$
& 1 -- 25 & 6 -- 25 & 1.0 & 3.8\\
&&&$\langle p_\perp \rangle$ vs.\ $N_{\mrm{ch}}$ & $ 2.5$ & $0.5$
& 25 -- 60 & 25 -- 60 & 5.0 & 19.1\\
\midrule
\textbf{CDF}/TC1 & 1800 & {\cite{Abe:1988yu}} &
$dN/dp_\perp$ & $1.0$ & $0.4$ & $0.4$ -- $10.0$ & $0.4$ -- $10.0$ & 1.0 & 1.1\\
\textbf{CDF}/TC1 & 1800 & \cite{Acosta:2001rm} &
$P(N_{\mrm{ch}})$ & $1.0$ & $0.4$ & 1 -- 17 & 6 -- 17 & 1.0 & 1.0\\
\textbf{CDF}/TC1 & 1800 & \cite{Acosta:2001rm} &
$P(N_{\mrm{ch}})$ & $1.0$ & $0.4$ & 17 -- 40 & 17 -- 40 & 5.0 & 5.0\\
\textbf{CDF}/TC1 & 1800 & \cite{Acosta:2001rm} &
$\langle p_\perp \rangle$ vs.\ $N_{\mrm{ch}}$ & $ 1.0$ & $0.4$
& 1 -- 17 & 6 -- 17 & 1.0 & 1.0\\
\textbf{CDF}/TC1 & 1800 & \cite{Acosta:2001rm} &
$\langle p_\perp \rangle$ vs.\ $N_{\mrm{ch}}$ & $ 1.0$ & $0.4$
& 17 -- 40 & 17 -- 40 & 5.0 & 5.0\\
\midrule
\textbf{CDF}/TC2 & 1960 & \cite{Aaltonen:2009ne} &
$dN/dp_\perp$ & $1.0$ & $0.4$ & $0.4$ -- $15.0$ & $0.4$ -- $15.0$ &1.0 & 1.0\\
\textbf{CDF}/TC2 & 1960 & \cite{Run2:MBnote} &
$P(N_{\mrm{ch}})$ & $1.0$ & $0.4$ & 1 -- 18 & 6 -- 18 & 1.0 & 1.0 \\
\textbf{CDF}/TC2 & 1960 & \cite{Run2:MBnote} &
$P(N_{\mrm{ch}})$ & $1.0$ & $0.4$ & 18 -- 30 & 18 -- 30 & 5.0 & 5.0\\
\textbf{CDF}/TC2 & 1960 & \cite{Aaltonen:2009ne} &
$\langle p_\perp \rangle$ vs.\ $N_{\mrm{ch}}$ & $ 1.0$ & $0.4$
& 1 -- 18 & 6 -- 18 & 1.0 & 1.0\\
\textbf{CDF}/TC2 & 1960 & \cite{Aaltonen:2009ne} &
$\langle p_\perp \rangle$ vs.\ $N_{\mrm{ch}}$ & $ 1.0$ & $0.4$
& 18 -- 30 & 18 -- 30 & 5.0 & 5.0 \\
\midrule
\textbf{ATL}/TA$^*$ & 7000 & \cite{ATLAS:1266235} &
$dN/dp_\perp$ & $2.5$ & $0.5$ & $0.5$ -- $40$ & $0.5$ -- $40.0$ & $1.0$ & 1.5 \\
\textbf{ATL}/TA$^*$ & 7000 & \cite{ATLAS:1266235} &
$P(N_{\mrm{ch}})$ & $2.5$ & $0.5$ & 1 -- 49 & 6 -- 49 & 1.0 & 1.5\\
\textbf{ATL}/TA$^*$ & 7000 & \cite{ATLAS:1266235} &
$P(N_{\mrm{ch}})$ & $2.5$ & $0.5$ & 49 -- 70 & $49$ -- $70$ & $5.0$ & 7.4 \\
\textbf{ATL}/TA$^*$ & 7000 & \cite{ATLAS:1266235} &
$\langle p_\perp \rangle$ vs.\ $N_{\mrm{ch}}$ & $ 2.5$ & $0.5$
& 1 -- 49 & 6 -- 49 & 1.0 & 1.5\\
\textbf{ATL}/TA$^*$ & 7000 & \cite{ATLAS:1266235} &
$\langle p_\perp \rangle$ vs.\ $N_{\mrm{ch}}$ & $ 2.5$ & $0.5$
& 49 -- 70 & 49 -- 70 & 5.0 & 7.4\\
\bottomrule
\end{tabular}}
\caption{Observables and ranges included in the study. The trigger
(``Trig'') conditions are as follows: {\sl TC1:} CDF Run
I MB \cite{Abe:1988yu}, {\sl TC2:} CDF Run II MB\cite{Aaltonen:2009ne}, {\sl TU1}: UA5 MB \cite{Abe:1988yu},
{\sl TA$^*$:} ATLAS trigger requiring
$\ge 1$ ($\ge 6$) charged particles within $|\eta|<2.5$ and $p_\perp >
0.5$~GeV for the $N_{\mrm{ch}}\ge 1$ ($N_{\mrm{ch}}\ge 6$) sample.
\label{tab:nch}. }
\end{table}
We consider two separate event samples: one where all events with at
least one charged track are included, corresponding to a conventional
min-bias definition (though excluding the ``zero bin'' when comparing
to experiments that include it in their measurements),
and one where only events with
$N_\mrm{ch}\ge 6$ are included, corresponding to a sample in which
diffractive contributions are expected to be strongly suppressed. Due
to the ambiguities discussed earlier concerning the treatment of
diffraction, we
shall base our conclusions mainly on the $N_\mrm{ch}\ge 6$ sample,
using the other as a further counter-check into the diffractive
region. Since measurements with an explicit $N_\mrm{ch}\ge 6$ definition
have so far only been carried out by ATLAS, the closest we can get
for other experiments is to suppress the five first bins in
$P(N_{\mrm{ch}})$ and $\langle \ensuremath{p_\perp}\xspace \rangle$ vs. $N_\text{ch}$, and
adjusting the normalization of the former such that the
remaining bins sum to unity.
A further complication concerns the phase-space (``fiducial'')
regions measured by the different experiments. Although no two
experiments have exactly the same coverage, it is here of great
help that all of the experiments we consider have taken data at at
least two energies, and in the best cases also in several different phase
space regions. This effectively allows us to construct a kind of
bootstrapped path among the different regions. In fact, without such
counter-checks, the method we propose here would be badly compromised ---
one could then never be certain whether a deviation in the optimized
parameters is caused by energy dependence or by the difference in
phase space regions. We therefore encourage the RHIC,
Tevatron, and LHC experiments in their efforts to make measurements
using several different combinations of trigger conditions, phase
space regions, and collider energies. Ultimately, it is by such
comprehensive and systematic sets of measurements, and by the
comprehensive tests that they enable, that we may establish a truly
reliable modeling of collider final states.
The observables, ranges, and weights used for both the $N_{\mrm{ch}}\ge 1$
and $N_{\mrm{ch}}\ge 6$ samples are given in table
\ref{tab:nch}. Larger weights are given to the high-multiplicity tails
of the $P(N_\mrm{ch})$ and $\left<p_\perp\right>(N_\mrm{ch})$
distributions to emphasize their asymptotic slopes. As our
definition for ``high multiplicity'', we took the $N_\mrm{ch}$ value
that came closest to separating out the 1\% highest-multiplicity
events, for each measurement.
\section{Consistency of Energy Scaling \label{sec:energyscaling}}
In this section, we study the degree to which the parameters obtained
for a best-fit tune across all included data sets and collider
energies are consistent with those obtained when we include only
specific subsets of the data. In particular, we focus on the
consistency on the assumed energy scaling by comparing the results of
the global fit to results obtained at each energy separately.
We study three specific questions
\begin{itemize}
\item Is the assumed scaling law governing the infrared regularization
scale for multiple parton interactions consistent with what one
finds when optimizing the tuning at each energy separately?
\item Is the transverse mass distribution of the proton
(assumed unchanging with energy) consistent with what one finds when
optimizing the tuning at each energy separately?
\item Is the assumed color reconnection strength
(assumed unchanging with energy) consistent with what one finds when
optimizing the tuning at each energy separately?
\end{itemize}
The evolution of the infrared regularization scale with energy is
depicted in \FigRef{fig:parp82}, with the left-hand panes showing the
results for the $N_{\mrm{ch}}\ge1$ sample
and the right-hand panes those for $N_{\mrm{ch}}\ge 6$. The scaling of
Perugia 0 (red dashed lines) is compared to the
global fit (red solid line in light shaded band) and to the
independent optimizations (blue horizontal lines inside cyan bands).
As mentioned in \SecRef{sec:professor}, the inner (darker blue) bands
correspond to the fit parameter uncertainties calculated by Minuit and
the outer (lighter cyan) bands include an estimate of
\textsc{Professor}'s interpolation uncertainty \cite{Buckley:2010nk}
as well, with the two
added in quadrature.
We make four conclusions concerning \parp{82}. One,
that the results of the independent optimizations are consistent with
the functional form represented by \EqRef{eqn:ptmin}
and hence we find no evidence for a need for any
significant departure from the model assumptions in this
energy range. Two, that the different individual
data sets appear to be consistent and reconcilable within this modeling
context --- the two different CDF measurements, the left- and
right-hand sample definitions, and the ATLAS and UA5 measurements at
900 GeV, all appear to give consistent parameters.
Three, that the light shaded (cyan) bands are very small, and that the
tune result can therefore be considered technically stable. And
finally, that the global fit \emph{does not} coincide with the
independent optimizations. Although not huge, this deviation hints
that one or more of the other parameters must be exhibiting a
non-universal behaviour.
Turning now to the scaling of the transverse shape parameter, \parp{83},
this is particularly interesting since it
could reveal whether minimum-bias collisions at different energies
effectively probe a different ``average proton shape''. Such a
variation could, e.g., be generated by correlations between $b$ and $x$
(see, e.g., \cite{Pancheri:2010dg,Blok:2010ge,Frankfurt:2010ea,Corke:2011yy}),
folded with the different $x$ ranges that are accessible at each
energy. Roughly speaking, we might then expect to see a slightly
more lumpy average proton at lower energies, consistent with a higher average
$x$ at those energies, and a smoother proton at higher energies / lower
average $x$. Results of the local and global tunes for \parp{83}
are shown in the middle panes of \FigRef{fig:parp82}.
\begin{figure}[tp]
\subfigure[\label{fig:p82nch1} \parp{82} vs \ensuremath{\sqrt{s}}\xspace, $N_\text{ch}\geq 1$ ]{
\includegraphics[width=0.5\textwidth]{fabulousplots-PARP82-nch1.pdf}
}
\subfigure[\label{fig:p82nch6} \parp{82} vs \ensuremath{\sqrt{s}}\xspace, $N_\text{ch}\geq 6$ ]{
\includegraphics[width=0.5\textwidth]{fabulousplots-PARP82-nch6.pdf}
}\\
\subfigure[\label{fig:p83nch1} \parp{83} vs \ensuremath{\sqrt{s}}\xspace, $N_\text{ch}\geq 1$ ]{
\includegraphics[width=0.5\textwidth]{fabulousplots-PARP83-nch1.pdf}
}
\subfigure[\label{fig:p83nch6} \parp{83} vs \ensuremath{\sqrt{s}}\xspace, $N_\text{ch}\geq 6$ ]{
\includegraphics[width=0.5\textwidth]{fabulousplots-PARP83-nch6.pdf}
}
\subfigure[\label{fig:p78nch1} \parp{78} vs \ensuremath{\sqrt{s}}\xspace, $N_\text{ch}\geq 1$ ]{
\includegraphics[width=0.5\textwidth]{fabulousplots-PARP78-nch1.pdf}
}
\subfigure[\label{fig:p78nch6} \parp{78} vs \ensuremath{\sqrt{s}}\xspace, $N_\text{ch}\geq 6$ ]{
\includegraphics[width=0.5\textwidth]{fabulousplots-PARP78-nch6.pdf}
}
\caption{Energy dependence of the the three tune parameters, from top
to bottom: \parp{82}, \parp{83}, and \parp{78}.
Independent optimizations (blue/shaded lines)
compared to global fit curve (red solid curves). {\sl Left:}
$N_{\mrm{ch}}\ge 1$ sample. {\sl Right:} $N_{\mrm{ch}}\ge 6$ sample.
\label{fig:parp82}}
\end{figure}
In our main $N_\mrm{ch}\ge 6$ sample, shown in the right-hand pane of
\FigRef{fig:parp82}, there may be some weak
evidence for such a trend, with a close-to-Gaussian proton (\parp{83}=2) favoured
by the high-energy data and a slightly more peaked distribution
favoured by the 630-GeV data. However, note that the shaded bands are here
larger, indicating that the fit is less well constrained than
it was for \parp{82}. Also, when we include the lowest-multiplicity
bins, in the left-hand pane, the trend disappears.
Our tentative conclusion is therefore that the uncertainties are
too large to make any firm conclusions, but that the model at least
appears to be self-consistent within those uncertainties.
Further studies at lower energies (e.g., including $pp$
data from RHIC and/or further Tevatron studies at 630 GeV)
and/or attempting to isolate different effective $x$ ranges at higher
energies, e.g., by using different rapidity and/or trigger regions, could
contribute significantly to probing this question further.
More theoretical work to improve the understanding of the relationship
between the language used here and that of other phenomenological
models, as was done, e.g., by \cite{Rogers:2009ke}, would also be
valuable and could open the possibility for a
consistent ``importation'' of constraints
from related physical models into the Monte Carlo context.
Returning to the present study, note also that \textsc{Professor}\xspace furnishes us with one
additional key piece of information. When optimizing several
parameters simultaneously, we not only get the optimized values for
each parameter separately; we also get a correlation matrix between
them. When interpreting our results, one should therefore be aware
that there is a strong correlation between \parp{82} and \parp{83}.
Hence, there is still a possibility that \parp{83} could
have a more significant energy dependence, to be traded off against that of
\parp{82}. However, since the fit result for \parp{82} was, itself,
quite stable, we consider this possibility something of a minority
report, not favored by the central fits; but also
not completely excluded.
Finally, the Color Reconnection strength, \parp{78}, is --- perhaps
not surprisingly --- the least well constrained parameter, with the
individual fit results showing a preferred scaling with energy that is not
accounted for by the underlying model. Given the large uncertainties
surrounding color correlations and final-state interactions in hadron
collisions, this can be considered a reminder that, although the
models do make attempts at incorporating this kind of phenomena, our
understanding is still very far from complete or reliable. For
the time being, any global tuning relying on the CR models considered
in this study would be forced to make a compromise
between the high- and low-energy data. Pragmatic alternatives for
physics studies at specific colliders would
range from giving that particular energy a larger weight in the global
fit to simply abandoning a global fit altogether. Although clearly not
theoretically satisfactory, the latter may be a useful strategy for
applications in which the Monte Carlo modeling is only used as a
sophisticated differential ``parameterization'' of the behavior of the
data. In particular at 7 TeV, large data samples are now
becoming available that probe many different and
complementary phase space regions in detail, allowing a fairly
complete set of constraints to be obtained for that particular
collider energy.
\begin{table}[t]
\centering
\begin{tabular}{l*{6}{c}}
\toprule
\ensuremath{\sqrt{s}}\xspace & 630 \,\mbox{Ge\kern-0.2exV}\xspace & 900 \,\mbox{Ge\kern-0.2exV}\xspace & 1.8 \,\mbox{Te\kern-0.2exV} & 1.96 \,\mbox{Te\kern-0.2exV} & 7 \,\mbox{Te\kern-0.2exV} & Global fit\\
\parp{89} & 630.0 & 900.0 & 1800.0 & 1960.0 & 7000.0 & 1800.0\\
\midrule
\multicolumn{6}{l}{Tuning to observables with $N_\text{ch}\geq1$}\\
\parp{78} & $0.53\pm0.10$ & $0.33\pm0.10$ & $0.49\pm0.09$ & $0.31\pm0.03$ & $0.22\pm0.05$\\
\parp{82} & $1.52\pm0.02$ & $1.68\pm0.06$ & $1.92\pm0.03$ & $2.00\pm0.04$ & $2.68\pm0.10$\\
\parp{83} & $1.93\pm0.07$ & $1.94\pm0.25$ & $1.82\pm0.14$ & $1.99\pm0.17$ & $1.85\pm0.20$\\
\midrule
\multicolumn{6}{l}{Tuning to observables with $N_\text{ch}\geq6$}\\
\parp{78} & $0.47\pm0.07$ & $0.33\pm0.09$ & $0.35\pm0.05$ & $0.31\pm0.02$ & $0.20\pm0.04$ & $0.40\pm0.11$\\
\parp{82} & $1.61\pm0.04$ & $1.69\pm0.07$ & $1.91\pm0.04$ & $1.97\pm0.05$ & $2.75\pm0.01$ & $2.19\pm0.06$ \\
\parp{83} & $1.50\pm0.14$ & $1.92\pm0.03$ & $2.02\pm0.09$ & $1.98\pm0.01$ & $1.73\pm0.02$ & $1.45\pm0.14$\\
\parp{90} & & & & & & $0.27\pm 0.02$ \\
\bottomrule
\end{tabular}
\caption{Tuning results obtained with the maximum information
parameterisations. The errors quoted are those calculated by Minuit.}
\label{tab:tunedparams}
\end{table}
\section{Conclusions \label{sec:conc}}
We have argued that the capabilities of modern tuning tools
can and should be used for more than just making ``best
fits'' to a collection of data. For example,
by making independent optimizations of the
MC generator parameters for several different collider energies, we
have here obtained a data-driven test of the universality of the generator
modeling. Three of the most important generator parameters controlling
the underlying-event and minimum-bias physics in the \textsc{Pythia 6}
generator were included in the study, corresponding to: the infrared
regularization scale for multiple parton interactions (MPI), the proton
transverse shape, and the strength of color reconnections (CR). A
brief discussion of each of these parameters was given in
\SecRef{sec:model}. The \textsc{Professor}\xspace tool used for the tunings as well
as a weighting strategy that attempts to take the size of
disparate statistical samples and measurements at closely spaced
energies formally into account were described in
\SecRef{sec:professor}. The data sets consisted of minimum-bias
measurements of charged
particle multiplicities, $p_\perp$ spectra, and the average of
$p_{\perp}$ vs.\ multiplicity, as described in \SecRef{sec:datasets}.
Our numerical results were presented and discussed
in \SecRef{sec:energyscaling}. We find
that the result of independent optimizations of the IR regularization
scale, energy by energy, are consistent with the power-law behaviour assumed
by the model, at least within the energy range we were able to probe,
from 630 to 7000 GeV. The transverse matter distribution may exhibit mild
deviations from universality, a question which data in particular from
minimum-bias measurements at RHIC could help shed further light
on. Finally, the optimal value of the color-reconnection strength
appears to vary significantly with energy, with lower values preferred
at higher energies, in contrast to the intrinsic assumption of a constant
strength in the model. This confirms the theoretical evaluation, that
the CR modeling is currently the largest source of theoretical
ambiguity, and
emphasizes that at least the models investigated here cannot be
considered truly universal over the studied energy range.
The parameter values corresponding to each of our ``local'' (i.e.,
energy-by-energy) tunes as well as those of a ``global'' one are collected
in \TabRef{tab:tunedparams}. All other parameter values were taken to
be those of the Perugia 0 tune \cite{Skands:2010ak}. These new tunes
have been included in \textsc{Pythia} starting from version 6.4.25,
with tune numbers 360 --- 365 (see the \textsc{Pythia} update notes
for details).
Finally, we argue that procedures similar to the one followed here
can be used to advantage also in other contexts, to give a clearer picture
of which regions the modeling is able to describe with approximately
universal parameters, which in turn helps isolate the
genuinely problematic areas more easily. The trend of the optimized parameters to deviate in one or another
direction in the problematic regions may also give clues as to the root of
the problem, though such interpretations should be made in conjunction
with a good understanding of the correlations between the parameters
and a careful evaluation of the possible missing physics components in the
modeling.
One clear possibility to apply this type of strategy to present
measurements would be
to perform independent optimizations using different complementary phase
space regions at each energy.
\section*{Acknowledgments}
This work was supported in part by the Marie Curie research training
network ``MCnet'' (contract number MRTN-CT-2006-035606).
HS acknowledges support from the German Research Foundation (DFG) in
the framework of Graduate School 1504.
\bibliographystyle{h-physrev3}
|
\section{Introduction}
Smooth minimal surfaces of general type with $p_g=q=0$ have been studied by several authors in the last years, but a classification is still missing.
We refer the surveys \cite{MP2} and \cite{BCP} for information on these surfaces.
There is, to my knowledge, only one example of a smooth minimal surface $S$ of general type with $p_g=0$ and $K^2=7$ (\cite{In}). This surface has an alternative description as a bidouble cover of a rational surface (\cite{MP1}). Its bicanonical map is of degree two onto a rational surface and is not composed with the other two involutions $i_1,i_2$ associated with the covering.
Moreover, $S/{i_1}$ and $S/{i_2}$ are also rational (cf. \cite{LS}).
In this paper we consider the case where $S$ has an involution $i$ such that the bicanonical map is not composed with $i$ and $S/i$ is not a rational surface.
We construct examples with $K^2=4,\ldots,7$ such that $S/{i}$ is birational to an Enriques surface.
This answers a question of Lee and Shin (\cite{LS}) about the existence of the cases with $K^2=5,6,7$ and $S/i$ birational to an Enriques surface.
In all cases $S$ has another involution $j$ such that $S/j$ is rational and the bicanonical map of $S$ is composed with $j$.
The paper is organized as follows.
First we recall some facts on involutions. Secondly we note that minor modifications to \cite[Theorems $7,$ $8$ and $9$]{Ri1} give a list of possibilities for the branch curve in the quotient surface $S/i$. Then we construct some examples as double covers of an Enriques surface obtained as a quotient of a Kummer surface. In Section \ref{Enr33} we describe these surfaces as bidouble covers of the plane. Finally we give some other bidouble plane examples.
\bigskip
\noindent{\bf Notation}
We work over the complex numbers; all varieties are assumed to be projective algebraic.
An {\em involution} of a surface $S$ is an
automorphism of $S$ of order 2. We say that a map is {\em composed with
an involution} $i$ of $S$ if it factors through the double cover $S\rightarrow
S/i.$
A {\em $(-2)$-curve} or {\em nodal curve} $N$ on a surface is a curve isomorphic to $\mathbb P^1$ such that $N^2=-2$.
An $(m_1,m_2,\ldots)$-point of a curve, or point of type $(m_1,m_2,\ldots),$ is a singular point of multiplicity $m_1,$ which resolves to a point of multiplicity $m_2$ after one blow-up, etc.
The rest of the notation is standard in Algebraic Geometry.\\
\bigskip
\noindent{\bf Acknowledgements}
The author wishes to thank Margarida Mendes Lopes for all the support.
He is a member of the Mathematics Center of the Universidade de Tr\'as-os-Montes e Alto Douro and is a collaborator of the Center for Mathematical Analysis, Geometry and Dynamical Systems of Instituto Superior T\'ecnico, Universidade T\' ecnica de Lisboa.
This research was partially supported by the Funda\c c\~ao para a Ci\^encia e a Tecnologia (Portugal) through Projects PEst-OE/MAT/UI4080/2011 and\newline PTDC/MAT/099275/2008.
\section{General facts on involutions}\label{GenFacts}
The following is according to \cite{CM}.\\
Let $S$ be a smooth minimal surface of general type with an
involution $i.$ Since $S$ is minimal of general type, this
involution is biregular. The fixed locus of $i$ is the union of a
smooth curve $R''$ (possibly empty) and of $t\geq 0$ isolated points
$P_1,\ldots,P_t.$ Let $S/i$ be the quotient of $S$ by $i$ and
$p:S\rightarrow S/i$ be the projection onto the quotient. The
surface $S/i$ has nodes at the points $Q_i:=p(P_i),$ $i=1,\ldots,t,$
and is smooth elsewhere. If $R''\not=\emptyset,$ the image via $p$
of $R''$ is a smooth curve $B''$ not containing the singular points
$Q_i,$ $i=1,\ldots,t.$ Let now $h:V\rightarrow S$ be the blow-up of
$S$ at $P_1,\ldots,P_t$ and set $R'=h^*(R'').$ The involution $i$
induces a biregular involution $\widetilde{i}$ on $V$ whose fixed
locus is $R:=R'+\sum_1^t h^{-1}(P_i).$ The quotient
$W:=V/\widetilde{i}$ is smooth and one has a commutative diagram:
$$
\begin{CD}\ V@>h>>S\\ @V\pi VV @VV p V\\ W@>g >> S/i
\end{CD}
$$
where $\pi:V\rightarrow W$ is the projection onto the quotient and
$g:W\rightarrow S/i$ is the minimal desingularization map. Notice
that $$A_i:=g^{-1}(Q_i),\ \ i=1,\ldots,t,$$ are $(-2)$-curves and
$\pi^*(A_i)=2\cdot h^{-1}(P_i).$
Set $B':=g^*(B'').$ Since $\pi$ is
a double cover with branch locus $B'+\sum_1^t A_i,$ it is determined
by a line bundle $L$ on $W$ such that $$2L\equiv B:=B'+\sum_1^t A_i.$$
\begin{proposition}[\cite{CM}, \cite{CCM}]
The bicanonical map of $S$ (given by $|2K_S|$) is composed with $i$ if and only if $h^0(W,\mathcal{O}_W(2K_W+L))=0$.
\end{proposition}
\section{List of possibilities}
Let $P$ be a minimal model of the resolution $W$ of $S/i,$ let $\rho:W\rightarrow P$ be the corresponding projection and denote by $\overline{B}$ the projection $\rho(B).$
\begin{theorem}\label{thm}
{\em ({\rm cf.} \cite{Ri1})}
Let $S$ be a smooth minimal surface of general type with $p_g=0$ having an involution $i$ such that the bicanonical map of $S$ is not composed with $i$ and $S/i$ is not rational.
With the previous notation, one of the following holds:
\begin{description}
\item[a)] $P$ is an Enriques surface and:
$\cdot$ $\overline B^2=0,$ $t-2=K_S^2\in\{2,\ldots,7\},$ $\overline B$ has a $(3,3)$-point or a $4$-uple point and at most one double point.
\item[b)] ${\rm Kod}(P)=1$ and:
$\cdot$ $K_P\overline B=2,$ $\overline B^2=-12,$ $t-2=K_S^2\in\{2,\ldots,8\},$ $\overline B$ has at most two double points, or
$\cdot$ $K_P\overline B=4,$ $\overline B^2=-16,$ $t=K_S^2\in\{4,\ldots,8\},$ $\overline B$ is smooth.
\item[c)] ${\rm Kod}(P)=2$ and:
$\cdot$ $K_S^2=2K_P^2,$ $K_P^2=1,\ldots,4,$ $\overline B$ is a disjoint union of four $(-2)$-curves, or
$\cdot$ $K_P\overline B=2,$ $K_P^2=1,$ $\overline B^2=-12,$ $t=K_S^2\in\{4,\ldots,8\},$ $\overline B$ has at most one double point, or
$\cdot$ $K_P\overline B=2,$ $K_P^2=2,$ $\overline B^2=-12,$ $t+2=K_S^2\in\{6,7,8\},$ $\overline B$ is smooth.
\end{description}
Moreover there are examples for {\rm a), b)} and {\rm c)}.
\end{theorem}
{\bf Proof :} This follows from the proof of \cite[Theorems $7,$ $8$ and $9$]{Ri1} taking in account that:
\begin{enumerate}
\item[$\cdot$] $p_g(P)=q(P)=0$ (because $p_g(P)\leq p_g(S),$ $q(P)\leq q(S)$);
\item[$\cdot$] $h^0(W,\mathcal{O}_W(2K_W+L))\leq\frac{1}{2}K_W^2+2$ (see \cite[Proposition $4$, b)]{Ri1});
\item[$\cdot$] $K_S^2\ne 9$ (see \cite[Theorem $4.3$]{DMP});
\item[$\cdot$] We can have $\overline {B'}^2>0$ (unlike the case $p_g=q=1$).
Examples for a) and b) are given below. Rebecca Barlow (\cite{Ba}) has constructed a surface of general type with $p_g=0$ and $K^2=1$ containing an even set of four disjoint $(-2)$-curves. This gives an example for c).
\end{enumerate}
\section{Examples}
\subsection{$S/i$ birational to an Enriques surface}\label{Enr}
Consider the involution of $\mathbb P^1\times \mathbb P^1$ $$j:[x:y\ ,\ a:b]\mapsto [y:x\ ,\ b:a]$$ and denote by $f,g$ the projections onto the first and second factors, respectively.
Let $F_1,\ldots,F_4,$ $G_1,\ldots,G_4$ be fibres of $f,$ $g$ such that $$C:=F_1+\cdots F_4+G_1+\cdots+G_4$$ is preserved by $j$ and does not contain the fixed points $[1:\pm 1,1:\pm 1]$ of $j.$
Let $$\pi:Q\rightarrow \mathbb P^1\times \mathbb P^1$$ be the double cover with branch locus $C$ and let $k$ be the corresponding involution.
It is well known that $Q$ is a Kummer surface and $$E:=Q/{k\circ j}$$ is an Enriques surface with $8$ nodes.
\subsubsection{$\overline B$ with a $4$-uple point}\label{Enr4}
Let $D\subset\mathbb P^1\times \mathbb P^1$ be a generic curve of bi-degree $(1,2)$ tangent to $C$ at smooth points $p_1, p_2$ of $C$ such that $p_2=j(p_1).$ The pullback $\pi^*(D+j(D))\subset Q$ is a reduced curve with two $4$-uple points, corresponding to the $(2,2)$-points of $D+j(D)$ (which are tangent to the branch curve $C$).
These points are identified by the involution $k\circ j,$ thus the projection of $\pi^*(D+j(D))$ into $E$ is a reduced curve $\overline{\overline {B'}}$ with one $4$-uple point.
Now let $\widetilde{E}$ be the minimal smooth resolution of the Enriques surface $E,$ $A_1,\ldots,A_8\subset\widetilde{E}$ be the nodal curves corresponding to the nodes of $E$ and $\overline {B'}\subset\widetilde E$ be the strict transform of $\overline{\overline {B'}}.$
If $D$ does not contain one of the $16$ double points of $C,$
the divisor $$\overline B:=\overline{B'}+\sum_1^8A_i$$ is reduced, divisible by $2$ in the Picard group and satisfies $\overline B^2=0.$
Let $S$ be the smooth minimal model of the double cover of $\widetilde E$ ramified over $\overline B$.
One can show that $S$ is a surface of general type with $p_g=0$ and $K_S^2=6.$
Moreover, $D$ can be chosen through one or two double points of $C.$ This provides examples with $K_S^2=5$ or $4$,
corresponding to branch curves $$\overline{B'}+\sum_1^7A_i\ \ \ \ {\rm or}\ \ \ \ \overline{B'}+\sum_1^6A_i.$$
\subsubsection{$\overline B$ with a $(3,3)$-point}
Let $D_1$ be a curve of bi-degree $(0,1)$ through $p,$ $D_2$ be a general curve of bi-degree $(1,1)$ through $p$ and $j(p)$ and set $D:=D_1+D_2.$ Then $D+j(D)$ is a reduced curve with triple points at $p$ and $j(p).$ Now we proceed as in Section \ref{Enr4}. In this case the branch curve $\overline B\subset\widetilde E$ has a $(3,3)$-point instead of a $4$-uple point. This gives an example of a surface of general type $S$ with $p_g=0$ and $K^2=7$ (notice that the resolution of the $(3,3)$-point gives rise to an additional nodal curve in the branch locus). As above, $D$ can be chosen containing one or two double points of $C,$ providing examples with $K_S^2=6$ or $5$.
\subsection{Bidouble plane description}\label{Enr33}
Here we obtain the examples of Section \ref{Enr} as bidouble covers of the plane.
\subsubsection{Construction}\label{construction}
Let $T_1,\ldots,T_4\subset\mathbb P^2$ be distinct lines through a point $p$ and $C_1,$ $C_2$ be distinct smooth conics tangent to $T_1,$ $T_2$ at points $p_1,p_2\ne p,$ respectively.
The smooth minimal model $\widetilde E$ of the double cover of $\mathbb P^2$ with branch locus $T_1+\ldots+T_4+C_1+C_2$ is an Enriques surface with $8$ disjoint nodal curves $A_1,\ldots,A_8,$ which correspond to the $8$ double points of $$G:=T_3+T_4+C_1+C_2.$$
Now let $p_3$ be a generic point in $T_3$ and consider the pencil $l$ generated by $2H_i+T_i,$ $i=1,2,3,$ where $H_i$ is a conic through $p_i$ tangent to $T_j,T_k$ at $p_j,p_k,$ for each permutation $(i,j,k)$ of $(1,2,3).$
Let $L$ be a generic element of $l.$
Notice that the quintic curve $L$ contains $p,$ it has a $(2,2)$-point at $p_i$ and the intersection number of $L$ and $T_i$ at $p_i$ is $4,$ $i=1,2,3.$
The strict transform of $L$ in $\widetilde E$ is a reduced curve $\overline{B'}$ with a $4$-uple point (at the pullback of $p_3$) such that the divisor $$\overline B:=\overline{B'}+\sum_1^8A_i$$
is reduced, satisfies $\overline B^2=0$ and is divisible by $2$ in the Picard group (because $L+T_1$ is divisible by $2$).
Let $S$ be the smooth minimal model of the double cover of $\widetilde E$ ramified over $\overline B$.
One can verify that $K_S^2=6.$
As in Section \ref{Enr4}, choosing $L$ through $1$ or $2$ double points of $T_3+T_4+C_1+C_2$ one obtains examples with $K_S^2=5$ or $4,$ respectively.
To obtain a branch curve $\overline B\subset\widetilde E$ with a $(3,3)$-point as in Section \ref{Enr33}, it suffices to change the $(2,2)$-point of the quintic $L$ at $p_3$ to an ordinary triple point. In this case $L$ is the union of a conic through $p_3$ with a cubic having a double point at $p_3.$ Choosing $C_1$ and $C_2$ so that $L$ passes through $0,$ $1$ or $2$ double points of $T_3+T_4+C_1+C_2$ one obtains examples with $K_S^2=7,$ $6$ or $5.$
\subsubsection{Involutions on $S$}
We refer \cite{Ca} or \cite{Pa} for information on bidouble covers.
Each surface $S$ constructed in Section \ref{construction} is the
smooth minimal model of the bidouble cover of $\mathbb P^2$
determined by the divisors
$$
\begin{array}{l}
D_1:=L,\\
D_2:=T_1+C_1+C_2,\\
D_3:=T_2+T_3+T_4.
\end{array}
$$
Let $i_g$ be the involution of $S$ corresponding to $D_j+D_k,$ for
each permutation $(g,j,k)$ of $(1,2,3).$ We have
that $S/{i_1}$ is birational to an Enriques surface, $S/{i_3}$ is a rational surface and
the bicanonical map of $S$ is not composed with $i_1, i_2$ and is composed with $i_3.$
Moreover, $S$ has an hyperelliptic fibration of genus $3$.
We omit the proof for these facts: it is similar to the one given in \cite{Ri3} for an example with $K_S^2=3.$
\subsection{More bidouble planes}
In the examples above, $S/{i_1}$ is birational to an Enriques
surface with $8$ disjoint $(-2)$-curves, corresponding to the $8$
nodes of the sextic $G=T_3+T_4+C_1+C_2,$ which contains $2$ lines.
Now we give examples with $G$ containing only one line and with $G$ without lines.
\subsubsection{$G$ with one line, $K_S^2=4,5,6$}
Let $T_1,T_2,T_3$ and $L$ be as in Section \ref{construction} and $p_4$ be a smooth point of $L.$ There exists a plane curve $J$ of degree $5$ through $p$ with $(2,2)$-points tangent to $T_1,T_2,L$ at $p_1,p_2,p_4,$ respectively (notice that we are imposing $19$ conditions to a linear system of dimension $20$; such a curve can be easily computed using the Magma function LinSys given in \cite{Ri2}).
Let $S$ be the smooth minimal model of the bidouble cover of
$\mathbb P^2$ determined by the divisors
$$
\begin{array}{l}
D_1:=L,\\
D_2:=T_3,\\
D_3:=T_1+T_2+J.
\end{array}
$$
Notice that the double plane with branch locus $D_2+D_3$ is an
Enriques surface $E$ with $6$ disjoint nodal curves
$A_1,\ldots,A_6$ (two of them are contained in the pullback of
$p_4$) and that the strict transform $\widehat L$ of $L$ in $E$
has a $4$-uple point at the pullback of $p_3.$ Moreover, the
divisor $\overline B:=\widehat L+\sum_1^6 A_i$ satisfies
$\overline B^2=0$ and is even (because $L+T_3$ is even).
This gives an example for Theorem \ref{thm}, a) with $K_S^2=4.$
To obtain an example with $K_S^2=5$ it suffices to choose the quintic $J$ with a triple point at $p_4$ instead of a $(2,2)$-point. In this case $J$ is the union of a conic with a singular cubic. Here the Enriques surface contains $7$ disjoint nodal curves, three of them contained in the pullback of $p_4.$
Finally, choosing $L$ with a triple point at $p_3$ one obtains $\widehat L\subset E$ with a $(3,3)$-point.
This gives examples for Theorem \ref{thm}, a) with $K_S^2=5,6.$
\subsubsection{$G$ without lines, $K_S^2=4$}
Consider, in affine plane, the points $p_0,\ldots,p_5$ with coordinates $(0,0),$ $(1,1),$ $(-1,1),$ $(2,3),$ $(-2,3),$ $(0,5),$ respectively, and let $T_{ij}$ be the line through $p_i,p_j.$
Let $C_1$ be the conic tangent to $T_{01},T_{02}$ at $p_1,p_2$ which contains $p_5$ and let $C_2$ be the conic tangent to $T_{01},T_{02}$ at $p_1,p_2$ which contains $p_3,p_4.$
Let $l$ be the linear system generated by $T_{01}+T_{02}+2T_{34}$ and $T_{03}+T_{04}+C_2.$
The element $Q$ of $l$ through $p_5$ is an irreducible quartic curve with double points at $p_0,p_3,p_4$ and tangent to $T_{01},T_{02}$ at $p_1,p_2$. Moreover, because of the symmetry with respect to $T_{05},$ the line $H$ tangent to $Q$ at $p_5$ is horizontal.
There is a cubic $F$ through $p_0,p_3,p_4$ tangent to $T_{01},$ $T_{02},$ $H$ at $p_1,$ $p_2,$ $p_5,$ respectively (notice that we are imposing $9$ conditions to a linear system of dimension $9$).
One can verify that $F$ contains no line, thus it is irreducible.
The surface $S$ is the smooth minimal model of the bidouble cover of $\mathbb P^2$ determined by the divisors
$$
\begin{array}{l}
D_1:=C_1+F,\\
D_2:=T_{01},\\
D_3:=T_{02}+C_2+Q.
\end{array}
$$
Notice that the double plane with branch locus $D_2+D_3$ is an
Enriques surface $E$ with $6$ disjoint nodal curves (contained in
the pullback of the triple points $p_3,p_4$ of $D_3$) and that the
strict transform of $D_1$ in $E$ has a $4$-uple point (at the
pullback of $p_5$).
The double plane with branch locus $D_1+D_3$ is a surface with Kodaira dimension 1. This gives an example for Theorem \ref{thm}, a), b) with $K_S^2=4.$
|
\section{Introduction}
Global nonlinear dynamics of various classes of closed loop attitude control systems have been studied in recent years. An overview of results on attitude control of a rotating rigid body is given in ~\cite{ChSaMcCSM11}. Closely related results on attitude control of a spherical pendulum (with attitude an element of the two-sphere $\ensuremath{\mathsf{S}}^2$) and of a 3D pendulum (with attitude an element of the special orthogonal group $\ensuremath{\mathsf{SO(3)}}$) are given in ~\cite{ChMcIJRNC07,ChMcBeAut08,ChaMcCITAC09}. These and other similar publications address the global closed dynamics of smooth vector fields. Assuming that the closed loop vector field has an asymptotically stable equilibrium, as desired in attitude stabilization problems, additional hyperbolic equilibria necessarily exist. The domain of attraction of the asymptotically stable equilibrium is contained in the complement of the union of the stable manifolds of the hyperbolic equilibria. These geometric factors motivate the current paper, in which new analytical and computational results on the stable manifolds of the hyperbolic equilibria are obtained.
To make the development concrete, the presentation is built around two specific closed loop vector fields: one for the attitude dynamics of a spherical pendulum and one for the attitude dynamics of a 3D pendulum. In analyzing these two cases, we introduce new analytical and computational tools that are broadly applicable to studying the geometry of more general attitude control systems.
\section{Spherical Pendulum}
A spherical pendulum is composed of a mass $m$ connected to a frictionless pivot by a massless link of length $l$.
It is acts under uniform gravity, and it is subject to a control moment $u$. The configuration of a spherical pendulum is described by a unit-vector $q\in\Re^3$, representing the direction of the link with respect to a reference frame.
Therefore, the configuration space is the two-sphere $\ensuremath{\mathsf{S}}^2=\{q\in\Re^3\,|\, q\cdot q =1\}$. The tangent space of the two-sphere at $q$, namely $\ensuremath{\mathsf{T}}_q\ensuremath{\mathsf{S}}^2$, is the two-dimensional plane tangent to the unit sphere at $q$, and it is identified with $\ensuremath{\mathsf{T}}_q\ensuremath{\mathsf{S}}^2\simeq\{\omega\in\Re^3\,|\, q\cdot \omega =0\}$, using the following kinematics equation:
\begin{align*}
\dot q= \omega\times q,
\end{align*}
where the vector $\omega\in\Re^3$ represents the angular velocity of the link. The equation of motion is given by
\begin{gather*}
\dot\omega = \frac{g}{l}q\times e_3 + \frac{1}{ml^2} u,
\end{gather*}
where the constant $g$ is the gravitational acceleration, and the vector $e_3=[0,0,1]\in\Re^3$ denotes the unit vector along the direction of gravity. The control moment at the pivot is denoted by $u\in\Re^3$.
\subsection{Control System}
Several proportional-derivative (PD) type control systems have been developed on $\ensuremath{\mathsf{S}}^2$ in a coordinate-free fashion~\cite{BulMurN95,BulLew05}. Here, we summarize a control system that stabilizes a spherical pendulum to a fixed desired direction $q_d\in\ensuremath{\mathsf{S}}^2$.
Consider an error function on $\ensuremath{\mathsf{S}}^2$, representing the projected distance from the direction $q$ to the desired direction $q_d$, given by
\begin{align*}
\Psi(q,q_d)=1-q\cdot q_d.
\end{align*}
The derivative of $\Psi$ with respect to $q$ along the direction $\delta q =\xi\times q$, where $\xi\in\Re^3$ and $\xi\cdot q=0$, is given by
\begin{align*}
\ensuremath{\mathbf{D}}_q \Psi(q,q_d)\cdot\delta q = -(\xi\times q)\cdot q_d = (q_d\times q)\cdot \xi.
\end{align*}
For positive constants $k_q,k_\omega$, the control input is chosen as:
\begin{align*}
u = ml^2(- k_\omega \omega - k_q q_d\times q -\frac{g}{l}q\times e_3).
\end{align*}
The corresponding closed loop dynamics are given by
\begin{gather}
\dot \omega = -k_\omega\omega - k_q q_d\times q,\label{eqn:dotw}\\
\dot q= \omega\times q.\label{eqn:dotq}
\end{gather}
This yields two equilibrium solutions: (i) the desired equilibrium $(q,\omega)=(q_d,0)$; (ii) additionally, there exists another equilibrium $(-q_d,0)$ at the antipodal point on the two-sphere.
It can be shown that the desired equilibrium is asymptotically stable by using the following Lyapunov function:
\begin{align*}
\mathcal{V} = \frac{1}{2} \omega\cdot\omega + k_q \Psi(q,q_d).
\end{align*}
In this paper, we analyze the local stability of each equilibrium by linearizing the closed loop dynamics to study the equilibrium structures more explicitly. In particular, we develop a coordinate-free form of the linearized dynamics of \refeqn{dotw}, \refeqn{dotq}, in the following section.
\subsection{Linearization}
A variation of a curve $q(t)$ on $\ensuremath{\mathsf{S}}^2$ is a family of curves $q^\epsilon(t)$ parameterized by $\epsilon\in\Re$, satisfying several properties~\cite{BulLew05}. It cannot be simply written as $q^\epsilon(t)=q(t)+\epsilon\delta q(t)$ for $\delta q(t)$ in $\Re^3$, since in general, this does not guarantee that $q^\epsilon(t)$ lies in $\ensuremath{\mathsf{S}}^2$. In~\cite{LeeLeoIJNME08}, an expression for a variation on $\ensuremath{\mathsf{S}}^2$ is given in terms of the exponential map as follows:
\begin{align}
q^\epsilon(t) = \exp(\epsilon \hat\xi(t))q(t),\label{eqn:qe}
\end{align}
for a curve $\xi(t)$ in $\Re^3$ satisfying $\xi(t)\cdot q(t)=0$ for all $t$. The \textit{hat map} $\hat\cdot:\Re^3\rightarrow\ensuremath{\mathfrak{so}(3)}$ is defined by the condition that $\hat x y =x\times y$ for any $x,y\in\Re^3$. The resulting infinitesimal variation is given by
\begin{align}
\delta q (t) = \frac{d}{d\epsilon}\bigg|_{\epsilon=0} q^\epsilon(t) = \xi(t)\times q(t). \label{eqn:delq}
\end{align}
The variation of the angular velocity can be written as
\begin{align}
\omega^\epsilon(t) = \omega(t) + \epsilon \delta \omega(t),\label{eqn:we}
\end{align}
for a curve $\delta w(t)$ in $\Re^3$ satisfying $q(t)\cdot w(t)=0$ for all $t$. Hereafter, we do not write the dependency on time $t$ explicitly.
The time-derivative of $\delta q$ can be obtained either from \refeqn{delq} or by substituting \refeqn{qe}, \refeqn{we} into \refeqn{dotq}, and considering the first order terms of $\epsilon$. In either case, we have
\begin{align*}
\delta\dot q= \dot\xi\times q + \xi\times(\omega\times q) = \delta\omega\times q + \omega\times(\xi\times q).
\end{align*}
Using the vector cross product identity $a\times (b\times c)=(a\cdot c)b-(a\cdot b) c$ for any $a,b,c\in\Re^3$, this can be written as
\begin{gather*}
\dot\xi\times q + (\xi\cdot q)w -(\xi\cdot\omega) q = \delta\omega\times q +(\omega\cdot q)\xi -(\omega\cdot\xi)q.
\end{gather*}
Since $\xi\cdot q=0$, $\omega\cdot q=0$, this reduces to
\begin{gather*}
\dot\xi\times q = \delta\omega\times q.
\end{gather*}
Since both sides of the above equation are perpendicular to $q$, this is equivalent to $q\times(\dot\xi\times q) = q\times(\delta\omega\times q)$, which yields
\begin{gather*}
\dot \xi - (q\cdot\dot\xi) q = q\times(\delta\omega\times q).
\end{gather*}
Since $\xi\cdot q =0$, we have $\dot\xi\cdot q +\xi\cdot\dot q=0$. Using this, the above equation can be rewritten as
\begin{align}
\dot \xi & = -(\xi\cdot(\omega\times q))q+ q\times(\delta\omega\times q)\nonumber\\
& = (qq^T\hat\omega) \xi +(I-qq^T)\delta\omega.\label{eqn:dotxi}
\end{align}
This corresponds to the linearized equation of motion for \refeqn{dotq}. Similarly, by substituting \refeqn{delq}, \refeqn{we} into \refeqn{dotw}, we obtain
\begin{align}
\delta\dot\omega &= -k_\omega\delta\omega -k_qq_d\times(\xi\times q)\nonumber\\
&=-k_\omega\omega +k_q\hat q_d \hat q\,\xi,\label{eqn:dotdelw}
\end{align}
which is the linearized equation for \refeqn{dotw}.
Equations \refeqn{dotxi}, \refeqn{dotdelw} can be written in a matrix form as
\begin{align}
\dot x =
\begin{bmatrix}\dot\xi \\ \delta\dot\omega \end{bmatrix}
=\begin{bmatrix} qq^T\hat\omega & I-qq^T\\k_q\hat q_d\hat q & -k_wI\end{bmatrix}
\begin{bmatrix}\xi \\ \delta\omega \end{bmatrix}=Ax,\label{eqn:xdot}
\end{align}
where the state vector of the linearized controlled system is $x=[\xi;\delta\omega]\in\Re^6$. A spherical pendulum has two degrees of freedom, but this linearized equation of motion evolves in $\Re^6$ instead of $\Re^4$. Since $q\cdot\omega=0$ and $q\cdot\xi=0$, we have the following two additional constraints on $\xi,\delta\omega$:
\begin{align}
Cx=
\begin{bmatrix} q^T & 0 \\ -\omega^T\hat q & q^T\end{bmatrix}
\begin{bmatrix}\xi \\ \delta\omega \end{bmatrix}
=\begin{bmatrix} 0 \\ 0 \end{bmatrix}.\label{eqn:con}
\end{align}
Therefore, the state vector $x$ should lie in the null space of the matrix $C\in\Re^{2\times 4}$. However, this is not an extra constraint that should be imposed when solving \refeqn{xdot}. As long as the initial condition $x(0)$ satisfies \refeqn{con}, the structure of \refeqn{dotw}, \refeqn{dotq}, and \refeqn{xdot}, guarantees that the state vector $x(t)$ satisfies \refeqn{con} for all $t$, i.e. $\frac{d}{dt}C(t)x(t) =0$ for all $t\geq 0$ when $C(0)x(0)=0$. This means that the null space of $C$ is a flow-invariant subspace.
\subsection{Equilibrium Solutions}
We choose the desired direction as $q_d=e_3$. The equilibrium solution $(q_d,0)=(e_3,0)$ is referred to as the hanging equilibrium, and the additional equilibrium solution $(-q_d,0)=(-e_3,0)$ is referred to as the inverted equilibrium. We study the eigen-structure of each equilibrium using the linearized equation \refeqn{xdot}. To illustrate the ideas, the controller gains are selected as $k_q=k_\omega=1$.
\subsubsection{Hanging Equilibrium}
The eigenvalues $\lambda_i$, and the eigenvectors $v_i$ of the matrix $A$ at the hanging equilibrium $(e_3,0)$ are given by
\begin{gather*}
\lambda_{1,2}=(-1\pm\sqrt{3}i)/2,\;
\lambda_{3,4}=\lambda_{1,2},\;\lambda_5=0,\;\lambda_6=-1,\\
v_{1,2}= e_1 + (-1\pm\sqrt{3}i)e_4/2,\;
v_{3,4}= e_2 + (-1\pm\sqrt{3}i)e_5/2,\\
v_5=e_3,\quad v_6=e_6,
\end{gather*}
where $e_i\in\Re^6$ denotes the unit-vector whose $i$-th element is one, and other elements are zeros. Note that there are repeated eigenvalues, but we obtain six linearly independent eigenvectors, i.e., the geometric multiplicities are equal to the algebraic multiplicities.
The basis of the null space of the matrix $C$, namely $\mathcal{N}(C)$ is $\{e_1,e_2,e_4,e_5\}$. The solution of the linearized equation can be written as $x(t)=\sum_{i=1}^6 c_i \exp(\lambda_it) v_i$ for constants $c_i$ that are determined by the initial condition: $x(0)=\sum_{i=1}^6 c_i v_i$. But, the eigenvectors $v_5,v_6$ do not satisfy the constraint given by \refeqn{con}, since they do not lie in $\mathcal{N}(C)$. Therefore, the constants $c_5,c_6$ are zero for initial conditions that are compatible with \refeqn{con}. We have $\mathrm{Re}[\lambda_i]<0$ for $1\leq i\leq 4$. Therefore, the equilibrium $(q,\omega)=(e_3,0)$ is asymptotically stable.
\subsubsection{Inverted Equilibrium}
The eigenvalues $\lambda_i$, and the eigenvectors $v_i$ of the matrix $A$ at the inverted equilibrium $(-e_3,0)$ are given by
\begin{gather}
\lambda_{1,2}=-(\sqrt{5}+1)/2,\lambda_{3,4}=(\sqrt{5}-1)/2,\lambda_5=0,\lambda_6=-1,\nonumber\\
v_{1}= e_1 -(\sqrt{5}+1)e_4/2,\, v_2=e_2-(\sqrt{5}+1)e_5/2,\label{eqn:v1v2}\\
v_3=(\sqrt{5}+1)e_1/2 +e_4,\, v_4=(\sqrt{5}+1)e_2/2+e_5,\nonumber\\
v_5=e_3,\, v_6=e_6.\nonumber
\end{gather}
The basis of $\mathcal{N}(C)$ is $\{e_1,e_2,e_4,e_5\}$. Hence, the eigenvectors $v_5,v_6$ do not lie in $\mathcal{N}(C)$. Therefore, the solution can be written as $x(t)=\sum_{i=1}^4 c_i \exp(\lambda_it) v_i$ for constants $c_i$ that are determined by the initial condition.
We have $\mathrm{Re}[\lambda_{1,2}]<0$, and $\mathrm{Re}[\lambda_{3,4}]>0$. Therefore, the inverted equilibrium $(q,\omega)=(-e_3,0)$ is a hyperbolic equilibrium, and in particular, a saddle point.
\subsection{Stable Manifold for the Inverted Equilibrium}\label{sec:SM}
\subsubsection{Stable Manifold}
The saddle point $(-e_3,0)$ has a stable manifold $W^s$, which is defined to be
\begin{align*}
W^s(-e_3,0) &= \{ (q,\omega)\in \ensuremath{\mathsf{T}}\ensuremath{\mathsf{S}}^2\,|\, \lim_{t\rightarrow\infty} \mathcal{F}^t(q,\omega) = (-e_3,0)\},
\end{align*}
where $\mathcal{F}^t:(q(0),\omega(0))\rightarrow(q(t),\omega(t))$ denotes the flow map along the solution of \refeqn{dotw}, \refeqn{dotq}. The existence of $W^s(-e_3,0)$ has nontrivial effects on the overall dynamics of the controlled system. Trajectories in $W^s(-e_3,0)$ converge to the antipodal point of the desired equilibrium $(e_3,0)$, and it takes a long time period for any trajectory near $W^s(-e_3,0)$ to asymptotically converge to the desired equilibrium $(e_3,0)$.
According to the stable and unstable manifold theorem~\cite{Kuz98}, a local stable manifold $W^s_{loc}(-e_3,0)$ exists in the neighborhood of $(-e_3,0)$, and it is tangent to the stable eigenspace $E^s(-e_3,0)$ spanned by the eigenvectors $v_1$ and $v_2$ of the stable eigenvalues $\lambda_{1,2}$. The (global) stable manifold can be written as
\begin{align}
W^s(-e_3,0) & = \bigcup_{t>0} \mathcal{F}^{-t} ( W^s_{loc}(-e_3,0)),\label{eqn:Ws}
\end{align}
which states that the stable manifold $W^s$ can be obtained by globalizing the local stable manifold $W^s_{loc}$ by the backward flow map.
This yields a method to compute $W^s(-e_3,0)$~\cite{KraOsiIJBC05}. We choose a small ball $B_\delta\subset W^s_{loc}(-e_3,0)$ with a radius $\delta$ around $(-e_3,0)$, and we grow the manifold $W^s(-e_3,0)$ by evolving $B_\delta$ under the flow $\mathcal{F}^{-t}$. More explicitly, the stable manifold can be parameterized by $t$ as follows:
\begin{align}
W^s(-e_3,0) =\{ \mathcal{F}^{-t} (B_\delta)\}_{t>0}.\label{eqn:Wc}
\end{align}
We construct a ball in the stable eigenspace of $(-e_3,0)$ with sufficiently small radius $\delta$, i.e. $B_\delta\subset E^s_{loc}(-e_3,0)$. From the stable eigenvectors $v_1,v_2$ at \refeqn{v1v2}, $E^s_{loc}(-e_3,0)$ can be written as
\begin{align}
E^s_{loc}& (-e_3,0) = \{ (q,\omega)\in \ensuremath{\mathsf{T}}\ensuremath{\mathsf{S}}^2\,|\,
q=\exp(\alpha_1\hat e_1+\alpha_2\hat e_2)(-e_3),\nonumber\\
& \omega=-\hat q^2(-(\sqrt{5}+1)/2)(\alpha_1 e_1+\alpha_2e_2)\text{ for $\alpha_1,\alpha_2\in\Re$}\},\label{eqn:Esloc}
\end{align}
where $-\hat q^2$ in the expression for $\omega$ corresponds to the orthogonal projection onto the plane normal to $q$, as required due to the constraint $q\cdot\omega=0$.
We define a distance on $\ensuremath{\mathsf{T}}\ensuremath{\mathsf{S}}^2$ as follows:
\begin{align}
d_{\ensuremath{\mathsf{T}}\ensuremath{\mathsf{S}}^2} ((q_1,\omega_1),(q_2,\omega_2)) = \sqrt{\Psi(q_1,q_2)} + \|\omega_1-\omega_2\|.\label{eqn:dis}
\end{align}
For $\delta>0$, the subset $B_\delta$ of $E^s_{loc}(-e_3,0)$ is parameterized by $\theta\in\ensuremath{\mathsf{S}}^1$ as
\begin{align}
B_\delta & = \{ (q,\omega)\in \ensuremath{\mathsf{T}}\ensuremath{\mathsf{S}}^2\,|\,
q=\exp(\alpha_1\hat e_1+\alpha_2\hat e_2)(-e_3),\nonumber\\
& \omega=-\hat q^2(-(\sqrt{5}+1)/2)(\alpha_1 e_1+\alpha_2e_2),\text{ where}\nonumber\\
& \text{$\alpha_1=\frac{\delta}{1/\sqrt{2}+(\sqrt{5}+1)/2}\cos\theta,$\;}\nonumber\\
& \text{$\alpha_2=\frac{\delta}{1/\sqrt{2}+(\sqrt{5}+1)/2}\sin\theta$,} \text{ for $\theta\in\ensuremath{\mathsf{S}}^1$} \}.\label{eqn:Bdelta}
\end{align}
The given choice of the constants $\alpha_1,\alpha_2$ guarantees that any point in $B_\delta$ has a distance $\delta$ to $(-e_3,0)$ according to the distance metric \refeqn{dis}.
\subsubsection{Variational Integrators}
The parameterization of the stable manifold $W_s$ in \refeqn{Wc} requires the computation of the backward flow map $\mathcal{F}^{-t}$. However, general purpose numerical integrators may not preserve the structure of the two-sphere or the underlying dynamic characteristics, such as energy dissipation rate, accurately, and they may yield qualitatively incorrect numerical results in simulating a complex trajectory over a long-time period~\cite{HaiLub00}.
Geometric numerical integration is concerned with developing numerical integrators that preserve geometric features of a system, such as invariants, symmetry, and reversibility. In particular, variational integrators are geometric numerical integrators for Lagrangian or Hamiltonian systems, constructed according to Hamilton's principle. They have desirable computational properties of preserving symplecticity and momentum maps, and they exhibit good energy behavior~\cite{MarWesAN01}. A variational integrator is developed for Lagrangian or Hamiltonian systems evolving on the two-sphere in~\cite{LeeLeoIJNME08}. It preserves both the underlying symplectic properties and the structures of the two-sphere concurrently.
A variational integrator on $\ensuremath{\mathsf{S}}^2$ for the controlled dynamics of a spherical pendulum can be written in a backward-time integration form as follows:
\begin{align}
q_{{k}} & = -\parenth{h\omega_{k+1} - \frac{h^2}{2ml^2} M_{k+1}}\times q_{k+1}\nonumber\\
&\quad + \parenth{1-\norm{h\omega_{k+1} - \frac{h^2}{2ml^2} M_{k+1}}^2}^{1/2} q_{k+1},\label{eqn:qk}\\
\omega_{{k}} & = \omega_{k+1} - \frac{h}{2ml^2} M_k - \frac{h}{2ml^2} M_{k+1},\label{eqn:wk}
\end{align}
where the constant $h>0$ is time step, the subscript $k$ denotes the value of a variable at the time $t_k=kh$, and $M_k = ml^2(-k_\omega \omega_k - k_q q_d\times q_k)$. For given $(q_{k+1},\omega_{k+1})$, we first compute $M_{k+1}$. Then, $q_k$ is obtained by \refeqn{qk}, followed by $M_k$, and $\omega_k$ is computed by \refeqn{wk}. This yields an explicit, discrete inverse flow map $\mathcal{F}_d^{-h}((q_{k+1},\omega_{k+1}))=(q_{k},\omega_{k})$.
\subsubsection{Visualization}
\setlength{\unitlength}{0.1\columnwidth}
\begin{figure}
\footnotesize\selectfont
\centerline{
\subfigure[$t=7\,(\mathrm{sec}$), $\|\omega\|_{\max}=0.05\,(\mathrm{rad/s})$]{
\begin{picture}(4.5,4.5)(0,0)
\put(0,0){\includegraphics[width=0.45\columnwidth]{SphRA_k3501}}
\put(4.0,2.2){$e_1$}\put(2.1,0.15){$e_2$}\put(1.2,2.7){$-e_3$}
\end{picture}}
\hspace*{.0cm}
\subfigure[$t=8\,(\mathrm{sec}$), $\|\omega\|_{\max}=0.29\,(\mathrm{rad/s})$]{
\begin{picture}(4.5,4.5)(0,0)
\put(0,0){\includegraphics[width=0.45\columnwidth]{SphRA_k4001}}
\put(4.0,2.2){$e_1$}\put(2.1,0.15){$e_2$}\put(1.2,2.7){$-e_3$}
\end{picture}}
}
\centerline{
\subfigure[$t=8.5\,(\mathrm{sec}$), $\|\omega\|_{\max}=0.65\,(\mathrm{rad/s})$]{
\begin{picture}(4.5,4.5)(0,0)
\put(0,0){\includegraphics[width=0.45\columnwidth]{SphRA_k4251}}
\put(4.0,2.2){$e_1$}\put(2.1,0.15){$e_2$}\put(1.2,2.7){$-e_3$}
\end{picture}}
\hspace*{.0cm}
\subfigure[$t=9\,(\mathrm{sec}$), $\|\omega\|_{\max}=1.43\,(\mathrm{rad/s})$]{
\begin{picture}(4.5,4.5)(0,0)
\put(0,0){\includegraphics[width=0.45\columnwidth]{SphRA_k4501}}
\put(4.0,2.2){$e_1$}\put(2.1,0.15){$e_2$}\put(1.2,2.7){$-e_3$}
\end{picture}}
}
\centerline{
\subfigure[$t=8.5\,(\mathrm{sec}$), $\|\omega\|_{\max}=2.96\,(\mathrm{rad/s})$]{
\begin{picture}(4.5,4.5)(0,0)
\put(0,0){\includegraphics[width=0.45\columnwidth]{SphRA_k4751}}
\put(4.0,2.2){$e_1$}\put(2.1,0.15){$e_2$}\put(1.2,2.7){$-e_3$}
\end{picture}}
\hspace*{.0cm}
\subfigure[$t=9\,(\mathrm{sec}$), $\|\omega\|_{\max}=8.02\,(\mathrm{rad/s})$]{
\begin{picture}(4.5,4.5)(0,0)
\put(0,0){\includegraphics[width=0.45\columnwidth]{SphRA_k5251}}
\put(4.0,2.2){$e_1$}\put(2.1,0.15){$e_2$}\put(1.2,2.7){$-e_3$}
\end{picture}\label{fig:Ws9}}
}
\caption{Stable manifold to $(q,\omega)=(-e_3,0)$ represented by $\{ \mathcal{F}^{-t} (B_\delta)\}_{t>0}$ for several values of $t$. One hundred points of $B_\delta$ in the stable eigenspace to $(-e_3,0)$ are chosen with $\delta=10^{-6}$, and they are integrated backward in time. Each trajectory is illustrated on a sphere, where the magnitude of angular velocity at each point is denoted by color shading (red: $\|\omega\|_{\max}$, blue: $\|\omega\|_{\min}\simeq 0$).}\label{fig:Ws}
\end{figure}
We choose 100 points on the surface of $B_\delta$ with $\delta=10^{-6}$, and each point is integrated backward using \refeqn{qk}, \refeqn{wk} with timestep $h=0.002$. The resulting trajectories are illustrated in \reffig{Ws} for several values of $t$. Each colored curve on the sphere represents a trajectory on $\ensuremath{\mathsf{T}}\ensuremath{\mathsf{S}}^2$, since at any point $q$ on the curve, the direction of $\dot q=\omega\times q$ is tangent to the curve at $q$, and the magnitude of $\dot q$ is indirectly represented by color shading.
We observe the following characteristics of the stable manifold $W_s(-e_3,0)$ of the inverted equilibrium:
\begin{itemize}
\item The boundary of the stable manifold $W_s(-e_3,0)\subset \ensuremath{\mathsf{T}}\ensuremath{\mathsf{S}}^2$ parameterized by $t$ is circular when projected onto $\ensuremath{\mathsf{S}}^2$.
\item Each trajectory in $W_s(-e_3,0)$ is on a great circle, when projected onto $\ensuremath{\mathsf{S}}^2$. According to the closed loop dynamics \refeqn{dotw}, and the given initial condition at the surface of $B_\delta$, the direction of $\dot\omega$ is always parallel to $\omega$. Therefore, the direction of $\omega$ is fixed, and the resulting trajectory of $q$ is on a great circle. This also corresponds to the fact that the eigenvalue $\lambda_1$ for the first mode representing the rotations about the first axis is equal to the eigenvalue $\lambda_2$ for the second mode representing the rotations about the second axis at \refeqn{v1v2}, i.e. the convergence rates of these two rotations are identical.
\item The angular velocity decreases to zero as the direction of the pendulum $q$ converges to $-e_3$.
\item The stable manifold $W_s(-e_3,0)$ may cover $\ensuremath{\mathsf{S}}^2$ multiple times if $t$ is sufficiently large, as illustrated at \reffig{Ws9}. Therefore, at any point $q\in\ensuremath{\mathsf{S}}^2$, we can choose $\omega$ such that $(q,\omega)$ lies in the stable manifold $W^s(-e_3,0)$ (the corresponding value of $\omega$ is not unique, since if it is sufficiently large, $q$ can traverse the sphere several times before converging to $-e_3$). This is similar to \textit{kicking} a damped spherical pendulum carefully such that it converges to the inverted equilibrium.
\end{itemize}
\section{3D Pendulum}
A 3D pendulum is a rigid body supported by a frictionless pivot acting under a gravitational potential. This is a generalization of a planar pendulum or a spherical pendulum, as it has three rotational degrees of freedom. It has been shown that a 3D pendulum may exhibit irregular maneuvers~\cite{ChaLeeJNS11}.
We choose a reference frame, and a body-fixed frame. The origin of the body-fixed frame is located at the pivot point. The attitude of a 3D pendulum is the orientation of the body-fixed frame with respect to the reference frame, and it is described by a rotation matrix representing the linear transformation from the body-fixed frame to the reference frame. The configuration manifold of a 3D pendulum is the special orthogonal group, $\ensuremath{\mathsf{SO(3)}}=\{R\in\Re^{3\times 3}\,|\, R^T R=I,\mathrm{det}[R]=1\}$.
The equations of motion for a 3D pendulum are given by
\begin{gather*}
J\dot\Omega + \Omega\times J\Omega = mg \rho\times R^T e_3 + u,\label{eqn:Wdot2}\\
\dot R = R\hat\Omega,\label{eqn:Rdot}
\end{gather*}
where the matrix $J\in\Re^{3\times 3}$ is the inertia matrix of the pendulum about the pivot, and $\rho\in\Re^3$ is the vector from the pivot to the center of mass of the pendulum represented in the body-fixed frame. The control moment at the pivot is denoted by $u\in\Re^3$.
\subsection{Control System}
Several control systems have been developed on $\ensuremath{\mathsf{SO(3)}}$~\cite{BulLew05,ChaMcCITAC09,LeePACC11}. Here, we summarize a control system to stabilize a 3D pendulum to a fixed desired attitude $R_d\in\ensuremath{\mathsf{SO(3)}}$. Consider an attitude error function given by
\begin{align*}
\Psi(R,R_d)=\frac{1}{2}\tr{(I-R_d^TR)G},
\end{align*}
for a diagonal matrix $G=\mathrm{diag}[g_1,g_2,g_3]\in\Re^{3\times 3}$ with $g_1,g_2,g_3>0$. The derivative of this attitude error function with respect to $R$ along the direction of $\delta R= R\hat\eta$ for $\eta\in\Re^3$ is given by
\begin{align*}
\ensuremath{\mathbf{D}}_R & \Psi(R,R_d)\cdot\delta R = -\frac{1}{2}\tr{R_d^TR\hat\eta G} \\
& = \frac{1}{2}(GR_d^TR -R^TR_d G)^\vee \cdot\eta\equiv e_R\cdot \eta,
\end{align*}
where we use the property that $\mathrm{tr}[\hat x A]=-x\cdot(A-A^T)^\vee$ for any $x\in\Re^3,A\in\Re^{3\times 3}$. The \textit{vee map}, $\vee:\ensuremath{\mathfrak{so}(3)}\rightarrow\Re^3$, denotes the inverse of the hat map. An attitude error vector is defined as $e_R = \frac{1}{2}(GR_d^TR -R^TR_d G)\in\Re^3$. For positive constants $k_\Omega,k_R$, we choose the following control input:
\begin{align*}
u = -k_R e_R -k_\Omega \Omega -mg \rho\times R^T e_3.
\end{align*}
The corresponding closed loop dynamics are given by
\begin{gather}
J\dot\Omega =- \Omega\times J\Omega -k_R e_R -k_\Omega \Omega,\label{eqn:Wdot2}\\
\dot R = R\hat\Omega.\label{eqn:Rdot}
\end{gather}
This system has four equilibria: in addition to the desired equilibrium $(R_d,0)$, there exist three other equilibria at $(R_d\exp (\pi\hat e_i,0),0)$ for $i\in\{1,2,3\}$, which correspond to the rotation of the desired attitude by $180^\circ$ about each body-fixed axis.
The existence of additional, undesirable equilibria is due to the nonlinear topological structure of $\ensuremath{\mathsf{SO(3)}}$, and it cannot be avoided by constructing a different control system (as long as it is continuous). It has been shown that it is not possible to design a continuous feedback control stabilizing an attitude globally on \ensuremath{\mathsf{SO(3)}}~\cite{BhaBerSCL00,KodPICDC98}.
The stability of the desired equilibrium can be studied by using the following Lyapunov function,
\begin{align*}
\mathcal{V} =\frac{1}{2}\Omega\cdot J\Omega + k_R \Psi(R,R_d).
\end{align*}
In this paper, we analyze the stability of each equilibrium by linearizing the closed loop dynamics to study the equilibrium structures more explicitly.
\subsection{Linearization}
A variation in $\ensuremath{\mathsf{SO(3)}}$ can be expressed as ~\cite{LeeLeoPICCA05}:
\begin{align}
R^\epsilon=R\exp(\epsilon\hat\eta),\quad \Omega^\epsilon=\Omega +\epsilon\delta\Omega,\label{eqn:delRdelW}
\end{align}
for $\eta,\delta\Omega\in\Re^3$. The corresponding infinitesimal variation of $R$ is given by $\delta R = R\hat\eta$. Substituting this into \refeqn{Rdot},
\begin{align*}
R\hat\Omega\hat\eta + R\hat{\dot\eta}= R\hat\eta\hat\Omega + R\delta\hat\Omega.
\end{align*}
Using the property $\hat x \hat y -\hat y\hat x=\widehat{x\times y}$ for any $x,y,\in\Re^3$, this can be rewritten as
\begin{align}
\dot\eta = \delta\Omega -\hat\Omega\eta.\label{eqn:etadot}
\end{align}
Similarly, by substituting \refeqn{delRdelW} into \refeqn{Wdot2}, we obtain
\begin{align}
J\delta\dot\Omega & = -\delta\Omega\times J\Omega -\Omega\times J\delta\Omega\nonumber\\
&\quad -\frac{1}{2}k_R (GR_d^T R\hat\eta +\hat\eta R^T R_d G) -k_\Omega\delta\Omega,\nonumber\\
& = (\widehat{J\Omega}-\hat\Omega J -k_\Omega I) \delta\Omega -\frac{1}{2}k_RH\eta,\label{eqn:delWdot}
\end{align}
where $H=\mathrm{tr}[R^T R_d G]I-R^T R_d G\in\Re^{3\times 3}$, and we used the property, $\hat x A + A^T\hat x = \mathrm{tr}[A]I-A$ for any $x\in\Re^3, A\in\Re^{3\times 3}$. Equations \refeqn{etadot},\refeqn{delWdot} can be written in matrix form as
\begin{align}
\dot x & = \begin{bmatrix}\dot\eta \\ \delta\dot\Omega \end{bmatrix}
=\begin{bmatrix}
-\hat\Omega & I\\
-\frac{1}{2}k_R J^{-1}H & J^{-1}(\widehat{J\Omega}-\hat\Omega J -k_\Omega I)
\end{bmatrix}
\begin{bmatrix}\eta \\ \delta\Omega \end{bmatrix}\nonumber\\
& = Ax.\label{eqn:xdotSO3}
\end{align}
This corresponds to the linearized equation of motion of \refeqn{Wdot2}, \refeqn{Rdot}.
\subsection{Equilibrium Solutions}
We choose the desired attitude as $R_d=I$. In addition to the desired equilibrium $(I,0)$, there are three additional equilibria, namely $(\exp(\pi\hat e_1),0)$, $(\exp(\pi\hat e_2),0)$, $(\exp(\pi\hat e_3),0)$. We study the eigen-structure of each equilibrium using the linearized equation \refeqn{xdotSO3}. We assume that
\begin{align*}
J=\mathrm{diag}[3,2,1]\,\mathrm{kgm^2},\; G=\mathrm{diag}[0.9,1,1.1],\; k_R=k_\Omega=1.
\end{align*}
\subsubsection{Equilibrium $(I,0)$}
The eigenvalues of the matrix $A$ at the desired equilibrium $(I,0)$ are given by
\begin{gather*}
\lambda_{1,2}=-0.1667\pm0.5676i,\\
\lambda_{3,4}=-0.25\pm 0.6614i,\\
\lambda_{5,6}=-0.5\pm 0.8367i.
\end{gather*}
This equilibrium is an asymptotically stable focus.
\subsubsection{Equilibrium $(\exp(\pi\hat e_1),0)$}
At this equilibrium, the eigenvalues and the eigenvectors of $A$ are given by
\begin{gather}
\lambda_1=-0.7813,\quad v_1= e_1-0.7813 e_4,\nonumber\\
\lambda_2=-0.5854,\quad v_2=e_2-0.5854e_5,\nonumber\\
\lambda_3=-1.0477,\quad v_3=e_3-1.0477e_6,\label{eqn:v3_SO31}\\
\lambda_4=0.4480,\quad v_4=e_1+0.4480e_4,\nonumber\\
\lambda_5=0.0854,\quad v_5=e_2+0.0854e_5,\nonumber\\
\lambda_6=0.0477,\quad v_6=e_3+0.0477e_6.\nonumber
\end{gather}
Therefore, this equilibrium is a saddle point, where three modes are stable, and three modes are unstable.
\subsubsection{Equilibrium $(\exp(\pi\hat e_2),0)$}
At this equilibrium, the eigenvalues and the eigenvectors of $A$ are given by
\begin{gather}
\lambda_1=-0.3775,\quad v_1= e_1-0.3775 e_4,\nonumber\\
\lambda_2=-1,\quad v_2=e_2-e_5,\label{eqn:v2_SO32}\\
\lambda_3=-0.9472,\quad v_3=e_3-0.9472e_6,\nonumber\\
\lambda_4=-0.0528,\quad v_4=e_3-0.0528e_6,\nonumber\\
\lambda_5=0.0442,\quad v_5=e_1+0.0442e_4,\nonumber\\
\lambda_6=0.5,\quad v_6=e_2+5e_5.\nonumber
\end{gather}
Therefore, this equilibrium is a saddle point, where four modes are stable, and two modes are unstable.
\subsubsection{Equilibrium $(\exp(\pi\hat e_3),0)$}
At this equilibrium, the eigenvalues and the eigenvectors of $A$ are given by
\begin{gather}
\lambda_1=-0.0613,\quad v_1= e_1-0.0613 e_4,\nonumber\\
\lambda_2=-0.2721,\quad v_2= e_1-0.2721 e_4,\nonumber\\
\lambda_3=-0.1382,\quad v_3= e_2-0.1382 e_5,\nonumber\\
\lambda_4=-0.3618,\quad v_4= e_2-0.3618 e_5,\nonumber\\
\lambda_5=-1.5954,\quad v_5= e_3-1.5954 e_6,\label{eqn:v5_SO33}\\
\lambda_6= 0.5954,\quad v_6= e_2+0.5954 e_6.\nonumber
\end{gather}
Therefore, this equilibrium is a saddle point, where five modes are stable, and one mode is unstable.
\subsection{Stable Manifolds for the Saddle Points}
The eigen-structure analysis shows that there exist multi-dimensional stable manifolds for each saddle point. They have zero measure as the dimension of stable manifold is less than the dimension of $\ensuremath{\mathsf{T}}\ensuremath{\mathsf{SO(3)}}$. But, the existence of these stable manifolds may have nontrivial effects on the attitude dynamics.
We numerically characterize these stable manifolds using backward time integration, as discussed in Section \ref{sec:SM}.
The stable eigenspace for each saddle point can be written as
\begin{align*}
&E^s_{loc} (\exp(\pi\hat e_1),0) = \{ (R,\Omega)\in \ensuremath{\mathsf{T}}\ensuremath{\mathsf{SO(3)}}\,|\,\nonumber\\
&\; R=\exp(\pi\hat e_1)\exp(\alpha_1\hat e_1+\alpha_2\hat e_2+\alpha_3\hat e_3),\\
&\; \Omega=-0.7813\alpha_1e_1-0.5854\alpha_2e_2-1.0477\alpha_3e_3\text{ for $\alpha_i\in\Re$}\},\\
&E^s_{loc} (\exp(\pi\hat e_2),0) = \{ (R,\Omega)\in \ensuremath{\mathsf{T}}\ensuremath{\mathsf{SO(3)}}\,|\,\nonumber\\
&\; R=\exp(\pi\hat e_2)\exp(\alpha_1\hat e_1+\alpha_2\hat e_2+(\alpha_3+\alpha_4)\hat e_3),\\
&\; \Omega=-0.37\alpha_1e_1-\alpha_2e_2-(0.94\alpha_3+0.05\alpha_4)e_3\text{ for $\alpha_i\in\Re$}\},\\
&E^s_{loc} (\exp(\pi\hat e_3),0) = \{ (R,\Omega)\in \ensuremath{\mathsf{T}}\ensuremath{\mathsf{SO(3)}}\,|\,\nonumber\\
&\; R=\exp(\pi\hat e_3)\exp((\alpha_1+\alpha_2)\hat e_1+(\alpha_3+\alpha_4)\hat e_2+\alpha_5\hat e_3),\\
&\; \Omega=-(0.06\alpha_1+0.27\alpha_2)e_1-(0.13\alpha_3+0.36\alpha_4)e_2\\
&\;\quad -1.59\alpha_5e_3\text{ for $\alpha_i\in\Re$}\},
\end{align*}
We define a distance on $\ensuremath{\mathsf{T}}\ensuremath{\mathsf{SO(3)}}$ as follows:
\begin{align*}
d_{\ensuremath{\mathsf{T}}\ensuremath{\mathsf{SO(3)}}} ((R_1,\Omega_1),(R_2,\Omega_2)) = \sqrt{\Psi(R_1,R_2)} + \|\Omega_1-\Omega_2\|.
\end{align*}
A variational integrator for the attitude dynamics of a rigid body on $\ensuremath{\mathsf{SO(3)}}$ is developed in~\cite{LeeLeoPICCA05,LeeLeoCMDA07}. It can be rewritten in a backward integration form as follows:
\begin{gather}
h (\Pi_{k+1}-\frac{h}{2}M_{k+1})^\wedge = J_dF_k - F_k^TJ_d,\label{eqn:Fk}\\
R_k = R_{k+1}F_k^T,\label{eqn:Rk}\\
\Pi_k = F_k\Pi_{k+1}-\frac{h}{2}F_kM_{k+1}-\frac{h}{2} M_k,\label{eqn:Pik}
\end{gather}
where $M_{k}=u_k+mg\rho\times R^T e_3\in\Re^3$ is the external moment, $\Pi_k=J\Omega_k\in\Re^3$ is the angular momentum. The matrix $J_d\in\Re^{3\times 3}$ denotes a non-standard inertia matrix given by $J_d = \frac{1}{2}\mathrm{tr}[J]I-J$, and the rotation matrix $F_k\in\ensuremath{\mathsf{SO(3)}}$ represent the relative attitude update between two integration time steps. For given $(R_{k+1},\Pi_{k+1})$, we first compute $M_{k+1}$, and solve \refeqn{Fk} for $F_k$. Then, $R_k$ is obtained by \refeqn{Rk}, and $\Pi_k$ is computed by \refeqn{Pik}. This yields a discrete inverse flow map, $\mathcal{F}^{-h}_d(R_{k+1},\Pi_{k+1})\rightarrow(R_{k},\Pi_{k})$.
\setlength{\unitlength}{0.1\columnwidth}
\begin{figure*}
\footnotesize\selectfont
\centerline{
\subfigure[$t=11\,(\mathrm{sec}$), $\|\Omega\|_{\max}=0.06\,(\mathrm{rad/s})$]{
\begin{picture}(4.5,4.5)(0,0)
\put(0,0){\includegraphics[width=0.45\columnwidth]{SO3RA_k5501}}
\put(3.9,0.9){$e_2$}\put(3.9,3.4){$e_1$}\put(2.3,4.25){$e_3$}
\put(0.1,3.4){$-e_2$}\put(2.3,0.1){$-e_3$}
\end{picture}}
\hspace*{0.2cm}
\subfigure[$t=12\,(\mathrm{sec}$), $\|\Omega\|_{\max}=0.17\,(\mathrm{rad/s})$]{
\begin{picture}(4.5,4.5)(0,0)
\put(0,0){\includegraphics[width=0.45\columnwidth]{SO3RA_k6001}}
\put(3.9,0.9){$e_2$}\put(3.9,3.4){$e_1$}\put(2.3,4.25){$e_3$}
\put(0.1,3.4){$-e_2$}\put(2.3,0.1){$-e_3$}
\end{picture}}
\hspace*{0.2cm}
\subfigure[$t=13\,(\mathrm{sec}$), $\|\Omega\|_{\max}=0.50\,(\mathrm{rad/s})$]{
\begin{picture}(4.5,4.5)(0,0)
\put(0,0){\includegraphics[width=0.45\columnwidth]{SO3RA_k6501}}
\put(3.9,0.9){$e_2$}\put(3.9,3.4){$e_1$}\put(2.3,4.25){$e_3$}
\put(0.1,3.4){$-e_2$}\put(2.3,0.1){$-e_3$}
\end{picture}}
\hspace*{0.2cm}
\subfigure[$t=14\,(\mathrm{sec}$), $\|\Omega\|_{\max}=1.42\,(\mathrm{rad/s})$]{
\begin{picture}(4.5,4.5)(0,0)
\put(0,0){\includegraphics[width=0.45\columnwidth]{SO3RA_k7001}}
\put(3.9,0.9){$e_2$}\put(3.9,3.4){$e_1$}\put(2.3,4.25){$e_3$}
\put(0.1,3.4){$-e_2$}\put(2.3,0.1){$-e_3$}
\end{picture}}
}
\centerline{
\subfigure[$t=15\,(\mathrm{sec}$), $\|\Omega\|_{\max}=3.93\,(\mathrm{rad/s})$]{
\begin{picture}(4.5,4.5)(0,0)
\put(0,0){\includegraphics[width=0.45\columnwidth]{SO3RA_k7501}}
\put(3.9,0.9){$e_2$}\put(3.9,3.4){$e_1$}\put(2.3,4.25){$e_3$}
\put(0.1,3.4){$-e_2$}\put(2.3,0.1){$-e_3$}
\end{picture}}
\hspace*{0.2cm}
\subfigure[$t=16\,(\mathrm{sec}$), $\|\Omega\|_{\max}=10.67\,(\mathrm{rad/s})$]{
\begin{picture}(4.5,4.5)(0,0)
\put(0,0){\includegraphics[width=0.45\columnwidth]{SO3RA_k8001}}
\put(3.9,0.9){$e_2$}\put(3.9,3.4){$e_1$}\put(2.3,4.25){$e_3$}
\put(0.1,3.4){$-e_2$}\put(2.3,0.1){$-e_3$}
\end{picture}}
\hspace*{0.2cm}
\subfigure[$t=17\,(\mathrm{sec}$), $\|\Omega\|_{\max}=29.00\,(\mathrm{rad/s})$]{
\begin{picture}(4.5,4.5)(0,0)
\put(0,0){\includegraphics[width=0.45\columnwidth]{SO3RA_k8501}}
\put(3.9,0.9){$e_2$}\put(3.9,3.4){$e_1$}\put(2.3,4.25){$e_3$}
\put(0.1,3.4){$-e_2$}\put(2.3,0.1){$-e_3$}
\end{picture}}
\hspace*{0.2cm}
\subfigure[$t=18\,(\mathrm{sec}$), $\|\Omega\|_{\max}=78.84\,(\mathrm{rad/s})$]{
\begin{picture}(4.5,4.5)(0,0)
\put(0,0){\includegraphics[width=0.45\columnwidth]{SO3RA_k9001}}
\put(3.9,0.9){$e_2$}\put(3.9,3.4){$e_1$}\put(2.3,4.25){$e_3$}
\put(0.1,3.4){$-e_2$}\put(2.3,0.1){$-e_3$}
\end{picture}}
}
\caption{Stable manifold to $(\exp(\pi\hat e_1),0)=([e_1,-e_2,-e_3],0)$ represented by $\{ \mathcal{F}^{-t} (B_\delta)\}_{t>0}$ with $\delta=10^{-6}$ for several values of $t$.}\label{fig:SM1}
\end{figure*}
\subsubsection{Visualization of $W_s(\exp(\pi\hat e_1),0)$}
In~\cite{LeeLeoPICDC08}, a method to visualize a function or a trajectory on $\ensuremath{\mathsf{SO(3)}}$ is proposed. Each column of a rotation matrix represents the direction of a body-fixed axis, and it evolves on $\ensuremath{\mathsf{S}}^2$. Therefore, a trajectory on $\ensuremath{\mathsf{SO(3)}}$ can be visualized by three curves on a sphere, representing the trajectory of three columns of a rotation matrix. The direction of the angular velocity should be chosen such that the corresponding time-derivative of the rotation matrix is tangent to the curve, and the magnitude of angular velocity can be illustrated by color shading. An example of visualizing a rotation about a single axis is illustrated in \reffig{visSO3_demo}.
\setlength{\unitlength}{0.1\columnwidth}
\begin{figure}
\centerline{
\subfigure[Visualization on sphere]{
\begin{picture}(4.5,4.5)(0,0)
\put(0,0){\includegraphics[width=0.45\columnwidth]{visSO3_Demo}}
\put(4.0,1.0){$e_2$}\put(0.3,0.6){$e_1$}\put(2.2,4.25){$e_3$}
\end{picture}}
\subfigure[Rotation angle $\beta$ (deg), and rotation rate $\dot\beta$ (rad/sec)]{
\includegraphics[width=0.50\columnwidth]{visSO3_Demo2}}
}
\caption{Visualization of an attitude maneuver: $R(t)=\exp(\beta(t)\hat e_3)$ for $0\leq t\leq 1$, where $\beta(t)=\frac{\pi}{6}(\sin\frac{\pi}{2}t -1)$. This maneuver corresponds to a rotation about the $e_3$ axis by $30^\circ$ to $R(1)=I$. The trajectory of the $i$-th column of $R(t)$ representing the direction of the $i$-th body-fixed axis is illustrated on a sphere for $i\in\{1,2,3\}$ (left). As the third body-fixed axis does not move during this maneuver, it is represented by a single point along the $e_3$ axis on the sphere. The direction of $\dot R(t)$ is tangent to these curves, and the magnitude of $\dot R(t)$ is denoted by color shading, according to the magnitude of the rotation rate (right).}\label{fig:visSO3_demo}
\end{figure}
We choose 112 points on the surface of $B_\delta\subset E^s_{loc} (\exp(\pi\hat e_1),0)$ with $\delta=10^{-6}$, and each point is integrated backward using \refeqn{qk}, \refeqn{wk} with timestep $h=0.002$. The resulting trajectories are illustrated in \reffig{SM1} for several values of $t$.
In each figure, three body-fixed axes of the desired attitude $R_d=[e_1,e_2,e_3]$, and three body-fixed axes of the additional equilibrium attitude $\exp(\pi\hat e_1)=[e_1,-e_2,-e_3]$ are shown. From these computational results, we observe the following characteristics on the stable manifold $W_s(\exp(\pi\hat e_1),0)$:
\begin{itemize}
\item When $t\leq 15$, the trajectories in $W_s(\exp(\pi\hat e_1),0)$ are close to rotations about the third body-fixed axis $e_3$ to $\exp(\pi\hat e_1)$. This is consistent with the linearized dynamics, where the eigenvalue of the third mode, corresponding to the rotations about $e_3$, has the fastest convergence rate, as seen in \refeqn{v3_SO31}.
\item When $t\geq 15$, the first mode representing the rotations about $e_1$ starts to appear, followed by the second mode representing the rotation about $e_2$. This corresponds to the fact that the first mode has a faster convergence rate than the second mode, i.e. $|\lambda_1|>|\lambda_2|$.
\item As $t$ is increased further, the third body-fixed axis leaves the neighborhood of $-e_3$, and it exhibit the following pattern:
\centerline{\includegraphics[width=0.32\columnwidth]{SO3RA_k9001_a}}
\item The stable manifold $W_s(\exp(\pi\hat e_1),0)$ covers a certain part of $\ensuremath{\mathsf{SO(3)}}$, when projected on to it. So, when an initial attitude is chosen such that its third body-fixed axis is sufficiently close to $-e_3$, there possibly exist multiple initial angular velocities such that the corresponding solution converges to $\exp(\pi\hat e_1)$ instead of the desired attitude $R_d=I$.
\end{itemize}
\subsubsection{Visualization of $W_s(\exp(\pi\hat e_2),0)$}
We choose 544 points on the surface of $B_\delta\subset E^s_{loc} (\exp(\pi\hat e_2),0)$ with $\delta=10^{-6}$, and each point is integrated backward using \refeqn{qk}, \refeqn{wk} with timestep $h=0.002$. The resulting trajectories are illustrated in \reffig{SM2} for several values of $t$.
In each figure, three body-fixed axes of the desired attitude $R_d=[e_1,e_2,e_3]$, and three body-fixed axes of the additional equilibrium attitude $\exp(\pi\hat e_2)=[-e_1,e_2,-e_3]$ are shown. From these computational results, we observe the following characteristics on the stable manifold $W_s(\exp(\pi\hat e_2),0)$:
\begin{itemize}
\item When $t\leq 12$, the trajectories in $W_s(\exp(\pi\hat e_2),0)$ is close to the rotations about the second body-fixed axis $e_2$. As $t$ increases, rotations about $e_3$ starts to appear. This corresponds to the linearized dynamics where the second mode representing rotations about $e_2$ has the fastest convergence rate, followed by the third mode at \refeqn{v2_SO32}.
\item As $t$ is increased further, nonlinear modes become dominant. The trajectories in $W_s(\exp(\pi\hat e_2),0)$ almost cover \ensuremath{\mathsf{SO(3)}}. This suggests that for any initial attitude, we can choose several initial angular velocities such that the corresponding solutions converges to $\exp(\pi\hat e_2)$.
\end{itemize}
\subsubsection{Visualization of $W_s(\exp(\pi\hat e_3),0)$}
Similarly, we choose 976 points on the surface of $B_\delta\subset E^s_{loc} (\exp(\pi\hat e_3),0)$ with $\delta=10^{-6}$, and each point is integrated backward using \refeqn{qk}, \refeqn{wk} with timestep $h=0.002$. The resulting trajectories are illustrated in \reffig{SM3} for several values of $t$.
At each figure, three body-fixed axes of the desired attitude $R_d=[e_1,e_2,e_3]$, and three body-fixed axes of the additional equilibrium attitude $\exp(\pi\hat e_3)=[-e_1,-e_2,e_3]$ are shown. From these computational results, we observe the following characteristics on the stable manifold $W_s(\exp(\pi\hat e_3),0)$:
\begin{itemize}
\item When $t\leq 8$, the trajectories in $W_s(\exp(\pi\hat e_3),0)$ are close to the rotations about the third body-fixed axis $e_3$. This corresponds to the linearized dynamics where the fifth mode representing rotations about $e_3$ has the fastest convergence rate given in \refeqn{v5_SO33}.
\item The rotations about $e_3$ are still dominant, even as $t$ is increased further. For the given simulation times, all trajectories in $W_s(\exp(\pi\hat e_3),0)$ are close to rotations about $e_3$.
\end{itemize}
\begin{figure*}
\footnotesize\selectfont
\centerline{
\subfigure[$t=11\,(\mathrm{sec}$), $\|\Omega\|_{\max}=0.03\,(\mathrm{rad/s})$]{
\begin{picture}(4.5,4.5)(0,0)
\put(0,0){\includegraphics[width=0.45\columnwidth]{SO3RA2_k5501}}
\put(0.2,1.1){$e_1$}\put(0.3,3.3){$e_2$}\put(2.35,4.25){$e_3$}
\put(3.8,3.2){$-e_1$}\put(2.35,0.1){$-e_3$}
\end{picture}}
\hspace*{0.2cm}
\subfigure[$t=12\,(\mathrm{sec}$), $\|\Omega\|_{\max}=0.09\,(\mathrm{rad/s})$]{
\begin{picture}(4.5,4.5)(0,0)
\put(0,0){\includegraphics[width=0.45\columnwidth]{SO3RA2_k6001}}
\put(0.2,1.1){$e_1$}\put(0.3,3.3){$e_2$}\put(2.35,4.25){$e_3$}
\put(3.8,3.2){$-e_1$}\put(2.35,0.1){$-e_3$}
\end{picture}}
\hspace*{0.2cm}
\subfigure[$t=13\,(\mathrm{sec}$), $\|\Omega\|_{\max}=0.25\,(\mathrm{rad/s})$]{
\begin{picture}(4.5,4.5)(0,0)
\put(0,0){\includegraphics[width=0.45\columnwidth]{SO3RA2_k6501}}
\put(0.2,1.1){$e_1$}\put(0.3,3.3){$e_2$}\put(2.35,4.25){$e_3$}
\put(3.8,3.2){$-e_1$}\put(2.35,0.1){$-e_3$}
\end{picture}}
\hspace*{0.2cm}
\subfigure[$t=14\,(\mathrm{sec}$), $\|\Omega\|_{\max}=0.69\,(\mathrm{rad/s})$]{
\begin{picture}(4.5,4.5)(0,0)
\put(0,0){\includegraphics[width=0.45\columnwidth]{SO3RA2_k7001}}
\put(0.2,1.1){$e_1$}\put(0.3,3.3){$e_2$}\put(2.35,4.25){$e_3$}
\put(3.8,3.2){$-e_1$}\put(2.35,0.1){$-e_3$}
\end{picture}}
}
\centerline{
\subfigure[$t=15\,(\mathrm{sec}$), $\|\Omega\|_{\max}=1.69\,(\mathrm{rad/s})$]{
\begin{picture}(4.5,4.5)(0,0)
\put(0,0){\includegraphics[width=0.45\columnwidth]{SO3RA2_k7501}}
\put(0.2,1.1){$e_1$}\put(0.3,3.3){$e_2$}\put(2.35,4.25){$e_3$}
\put(3.8,3.2){$-e_1$}\put(2.35,0.1){$-e_3$}
\end{picture}}
\hspace*{0.2cm}
\subfigure[$t=16\,(\mathrm{sec}$), $\|\Omega\|_{\max}=3.37\,(\mathrm{rad/s})$]{
\begin{picture}(4.5,4.5)(0,0)
\put(0,0){\includegraphics[width=0.45\columnwidth]{SO3RA2_k8001}}
\put(0.2,1.1){$e_1$}\put(0.3,3.3){$e_2$}\put(2.35,4.25){$e_3$}
\put(3.8,3.2){$-e_1$}\put(2.35,0.1){$-e_3$}
\end{picture}}
\hspace*{0.2cm}
\subfigure[$t=17\,(\mathrm{sec}$), $\|\Omega\|_{\max}=7.01\,(\mathrm{rad/s})$]{
\begin{picture}(4.5,4.5)(0,0)
\put(0,0){\includegraphics[width=0.45\columnwidth]{SO3RA2_k8501}}
\put(0.2,1.1){$e_1$}\put(0.3,3.3){$e_2$}\put(2.35,4.25){$e_3$}
\put(3.8,3.2){$-e_1$}\put(2.35,0.1){$-e_3$}
\end{picture}}
\hspace*{0.2cm}
\subfigure[$t=18\,(\mathrm{sec}$), $\|\Omega\|_{\max}=18.22\,(\mathrm{rad/s})$]{
\begin{picture}(4.5,4.5)(0,0)
\put(0,0){\includegraphics[width=0.45\columnwidth]{SO3RA2_k9001}}
\put(0.2,1.1){$e_1$}\put(0.3,3.3){$e_2$}\put(2.35,4.25){$e_3$}
\put(3.8,3.2){$-e_1$}\put(2.35,0.1){$-e_3$}
\end{picture}}
}
\caption{Stable manifold to $(\exp(\pi\hat e_2),0)=([-e_1,e_2,-e_3],0)$ represented by $\{ \mathcal{F}^{-t} (B_\delta)\}_{t>0}$ with $\delta=10^{-6}$ for several values of $t$.}\label{fig:SM2}
\end{figure*}
\begin{figure*}
\footnotesize\selectfont
\centerline{
\subfigure[$t=8\,(\mathrm{sec}$), $\|\Omega\|_{\max}=0.224\,(\mathrm{rad/s})$]{
\begin{picture}(4.5,4.5)(0,0)
\put(0,0){\includegraphics[width=0.43\columnwidth]{SO3RA3_k4001}}
\put(0.0,0.75){$-e_1$}\put(0.15,3.0){$e_2$}\put(2.25,4.15){$e_3$}
\put(3.95,3.0){$e_1$}\put(3.90,0.8){$-e_2$}
\end{picture}}
\hspace*{0.2cm}
\subfigure[$t=9\,(\mathrm{sec}$), $\|\Omega\|_{\max}=1.09\,(\mathrm{rad/s})$]{
\begin{picture}(4.5,4.5)(0,0)
\put(0,0){\includegraphics[width=0.43\columnwidth]{SO3RA3_k4501}}
\put(0.0,0.75){$-e_1$}\put(0.15,3.0){$e_2$}\put(2.25,4.15){$e_3$}
\put(3.95,3.0){$e_1$}\put(3.90,0.8){$-e_2$}
\end{picture}}
\hspace*{0.2cm}
\subfigure[$t=10\,(\mathrm{sec}$), $\|\Omega\|_{\max}=4.26\,(\mathrm{rad/s})$]{
\begin{picture}(4.5,4.5)(0,0)
\put(0,0){\includegraphics[width=0.43\columnwidth]{SO3RA3_k5001}}
\put(0.0,0.75){$-e_1$}\put(0.15,3.0){$e_2$}\put(2.25,4.15){$e_3$}
\put(3.95,3.0){$e_1$}\put(3.90,0.8){$-e_2$}
\end{picture}}
\hspace*{0.2cm}
\subfigure[$t=14\,(\mathrm{sec}$), $\|\Omega\|_{\max}=222.99\,(\mathrm{rad/s})$]{
\begin{picture}(4.5,4.5)(0,0)
\put(0,0){\includegraphics[width=0.43\columnwidth]{SO3RA3_k7001}}
\put(0.0,0.75){$-e_1$}\put(0.15,3.0){$e_2$}\put(2.25,4.15){$e_3$}
\put(3.95,3.0){$e_1$}\put(3.90,0.8){$-e_2$}
\end{picture}}
}
\caption{Stable manifold to $(\exp(\pi\hat e_3),0)=([-e_1,-e_2,e_3],0)$ represented by $\{ \mathcal{F}^{-t} (B_\delta)\}_{t>0}$ with $\delta=10^{-6}$ for several values of $t$.}\label{fig:SM3}
\end{figure*}
\section{Conclusions}
Stable manifolds of saddle points that arise in the closed-loop dynamics of two pendulum models are characterized numerically, and several properties are observed. Although the analytical and computational results have been presented for a spherical pendulum and a 3D pendulum, the methods presented naturally extend to any closed loop attitude control system with configurations in either $\ensuremath{\mathsf{S}}^2$ or $\ensuremath{\mathsf{SO(3)}}$.
|
\section{Introduction}
Much attention has been given to studying the transport of particles along narrow channels \cite{RMP09}. Such strong confinement occurs for example in the ion channels of biological membranes \cite{hille01}, in zeolites and other porous materials \cite{kaerger92} and in microfluidic devices \cite{squires05}. Experimental studies of colloidal particles confined within grooves etched on a surface \cite{wei00} have already addressed the case when confinement is so extreme that particles cannot pass one another and so single-file diffusion sets in \cite{lutz04}. In many cases, the motion of a file occurs in the presence of a periodic pinning potential. The latter can be induced, for instance, by defects, such as in the case of superconducting vortices moving in easy-flow channels \cite{besseling98}, by other particles, as in the case of a fluctuating quasi-1-dimensional (1D) channel \cite{coupier07}, by a periodic distribution of charges, as in the case of the motor proteins moving along a microtubulus \cite{ashkin90}, or, more generally, by a periodic corrugation of the channel walls \cite{wambaugh,taloni06}. When the left-right symmetry of the pinning potential is broken, an externally applied center-symmetric AC drive induces a net drift of the file in a certain direction. The efficiency of such a rectification mechanism strongly depends on the number of particles in the file, their size, and frequency of the drive \cite{vicsek95}.
In addition to the interaction with the channel, the colloids have an excluded volume interaction between them \cite{BarratHansen2003, vicsek95} and may also exhibit mutual attraction, either due to Van der Waals forces \cite{hansen2006tsl} or because of the presence of other passive molecules in the solution, such as in the case of colloid-polymer mixtures \cite{poon02, stradner04, savsav}. It has been recognized that attractive forces lead to the formation of particle clusters and, consequently, to a dramatic increase in particle diffusion and mobility \cite{sholl97}. Such enhancement is explained by the mismatch between the size of the particle clusters and the characteristic length scale of the corrugated potential induced by the walls of the channel. In the case of the diffusion of long alkane chains in zeolites, a similar phenomenon is called the ``window effect" \cite{dubbeldam03}, namely, the mobility of the alkane chain becomes enhanced, whenever its length is not commensurate with the zeolite cage. More generally, the incommensurability between the lateral dimensions of a biological molecule and the size of a catalyst is known as ``shape selectivity" \cite{smit08}, a recurrent scheme utilized by nature to control enzymatic reactions in living cells.
Recently, we have developed a theory \cite{PA10}, based on dynamical density functional theory (DDFT) \cite{marini99, marini00, archer04, archer04b}, that captures the essential features of the condensation process from the disordered to the condensed state in a randomly distributed single-file of interacting particles. Pair attraction can be used to enhance the transport of colloidal particles in 1D. For instance, entrained attracting particles can be effectively shuttled along an asymmetric corrugated channel by means of a low frequency AC field. Collective shuttling of entrained particles directly applies to the problem of diffusion of long molecular chains in zeolites and shape selective catalytic reactions in living cells. In particular, we stress that collective shuttles can be much more efficient than some other shuttle mechanisms, as they allow one to control the rate of transport by adding (removing) a single molecule to (from) the molecular chain.
The present paper is organized as follows: Our main goal is to analyze the effects of the pair attraction between the particles on the rectification current of single-files of colloidal particles which are confined within a narrow channel with corrugated walls and subjected to DC or AC drives. The general theoretical framework for our analysis of our model system, which is based on DDFT, is presented in Sec.\,\ref{ddft}. The one body density distribution $\rho(x,t)$ of the diffusing particles, which is a function of position $x$ and time $t$, obeys a nonlinear Fokker-Planck equation and DDFT \cite{marini99, marini00, archer04, archer04b} provides a closure approximation that allows us to solve for the dynamics of $\rho(x,t)$. The DDFT dynamical equation for the system takes the form of a conserved gradient dynamics, which requires as input a suitable approximation for the Helmholtz free energy functional for the system \cite{marini99, marini00, archer04, archer04b}. In the presence of a DC drive, the free energy contains a potential energy term proportional to $x$ that acts as a continuous external energy source, i.e., the system remains permanently out of thermodynamic equilibrium. In consequence, the free energy of a single-file of interacting particles is not necessarily a monotonically decreasing function of time. Therefore, the system can exhibit stable time-periodic density profiles and currents. In particular, in Secs\,\ref{cond} and \ref{DC} we discuss the relation of the onset of time-periodic density variations to the condensation of particles into compact clusters and the depinning of these clusters from the channel corrugations.
In Sec.\,\ref{zerolength} we examine the dynamics of point-like particles and we demonstrate that spontaneous symmetry breaking induced by attraction leads to the coexistence of stable time-periodic and stationary densities. Multistability of the long-time density distributions indicates that for the same combination of parameters in our model, the channel can operate in two different regimes, transporting the particles with either high or low efficiency. For finite-size particles the range of values of the system parameters that allows for time-periodic density profiles, is much broader than for point-like particles, as explained in Sec.\,\ref{finitelength}.
In Sec.\,\ref{acdrive} we discuss the low frequency rectification current for particles driven by an AC (square wave) drive through a channel with a spatially asymmetric potential. In Sec.\,\ref{currmax} we show that the effect of the spatial asymmetry is that the rectification current may be maximized, using the strength of the attraction between the particles as the control parameter. In Sec.\,\ref{strongatt} we derive an effective equation of motion for a condensate in the limit of infinitely strong attraction between the particles. We show that the low frequency transport efficiency can be increased by several orders of magnitude for strongly attracting particles as compared to non-interacting particles.
Finally, in Sec.\ \ref{conc} we close with a few concluding remarks.
\section{DDFT for interacting hard rods in a periodic external potential}
\label{ddft}
When finite sized colloids are confined within a long narrow channel, the particle motion becomes one dimensional, and the particles can be modeled as 1D hard rods of length $h$. We model the dynamics of $N$ hard rods using overdamped stochastic equations of motion. The particles move in a channel of total length $S$ and interact with the channel walls via a periodic corrugated potential $U(x)$ with spatial period $L$. In a system with periodic boundary conditions (e.g., a circular geometry), $S$ is an integer multiple of $L$, i.e. $S = M L$. The integer $M$ determines the average number of particles per unit cell of length $L$, to be $N/M$. To ensure that the combined length of the $N$ rods is smaller than the total system size, we require $N h \le S$.
The total instantaneous potential energy of the $N$ rods, which move in the periodic external potential $U(x)$ under the action of a time-dependent external drive $A(t)$, is:
\begin{eqnarray}
\label{potential}
\Phi(\{x_j\},t)=\sum_i[U(x_i)-A(t)x_i]+\frac{1}{2}\sum_{i,j \neq i} w(| x_i -x_j |),
\end{eqnarray}
where $w(x)$ denotes the interaction potential between a pair of particles $i$ and $j$, where $i,j=1,\dots,N$, which are located at positions $x_i$ and $x_j$, respectively, and are separated by the distance $x=|x_i -x_j|$. In general, the potential $w(x)$ can be decomposed into two terms. One accounts for the attraction (denoted by subscript ``at'') between the particles and the other for the hardcore repulsion (subscript ``hc''), i.e., $w(x) = w_{\rm at}(x) + w_{\rm hc}(x)$. The hard-core repulsive potential ensures that the rods are impenetrable, that is,
\begin{eqnarray}
\label{hr}
w_{\rm hc}(x) =
\left\{
\begin{array}{c}
\infty,\,\,\,x<h \\
0,\,\,\,\,\,x \geq h.
\end{array}
\right.
\end{eqnarray}
We assume that $w_{\rm at}(x)$ becomes negligible at distances $x$ much larger than a certain effective interaction range $l_{\rm int}$.
The overdamped dynamics of the particles is described by $N$ coupled Langevin equations
\begin{eqnarray}
\label{langevin}
\frac{1}{\Gamma}\frac{\partial x_i}{\partial t} = -\frac{\partial \Phi(\{x_j\},t)}{\partial x_i} + \sqrt{2T}\xi_i(t),
\end{eqnarray}
where $T$ is the temperature of the system and $\xi_i(t)$ are independent Gaussian white noises with the correlation functions $\langle \xi_i(t) \xi_j(t^\prime) \rangle = \delta_{ij}\delta(t-t^\prime)$. Henceforth we set the solvent friction constant $\Gamma=1$.
In the case of a circular geometry, Eqs.\,(\ref{langevin}) are supplemented with periodic boundary conditions (BC) with the period equal to the system size $S$. This requires that all functions in Eqs.\,(\ref{langevin}) are periodic with period $S$. This is not the case if the contribution to the interaction potential $w(x)$ from the potential $w_\mathrm{at}(x)$ is long ranged, as particles would interact then with themselves. However, $w(x)$ can be made compatible with the desired periodic BC by taking the interaction range $l_{\rm int}$ to be sufficiently small compared with the system size. Therefore, throughout we impose the condition $w_{\rm at}(S) \approx 0$.
The Fokker-Planck (Smoluchowski) equation for the time evolution of the one body density distribution $\rho(x,t)$ is \cite{marini99, marini00, archer04, archer04b}:
\begin{eqnarray}\notag
\frac{\partial \rho(x,t)}{\partial t} &=& T \frac{\partial^2 \rho(x,t)}{\partial x^2}
+\,\frac{\partial}{\partial x}\Bigg[ \rho(x,t) \frac{\partial U_{\rm eff}(x,t)}{\partial x}\Bigg] \notag \\
&\,&+\frac{\partial}{\partial x}\Bigg[ \int_{-\frac{S}{2}}^{\frac{S}{2}} {\rm d} x' \rho^{(2)}(x,x',t) \frac{\partial}{\partial x} w(|x-x'|)\Bigg],
\label{eq:FP}
\end{eqnarray}
where $\rho^{(2)}(x,x',t)$ is the non-equilibrium two-body distribution function for the particles in the system and $U_{\rm eff}(x,t)=U(x)-A(t) x$ is the effective external potential. Note that the density profile $\rho(x,t)$ is normalized, so that the spatial integral over $\rho(x,t)$ is equal to the total number of particles in the system; i.e.\ $\int_{-S/2}^{S/2}\rho(x,t)\,dx = N$.
In order to solve Eq.\ \eqref{eq:FP}, a suitable closure approximation for $\rho^{(2)}(x,x',t)$ is required. The approach taken in DDFT is to approximate $\rho^{(2)}(x,x',t)$ by the two-body distribution function of an equilibrium fluid with the same one-body density profile as the non-equilibrium system \cite{marini99, marini00, archer04, archer04b}. This closure relates the integral in the final term in Eq.\ \eqref{eq:FP} to the functional derivative of the excess part of the Helmholtz free energy functional $F[\rho]$, which is the central quantity of interest in equilibrium density functional theory \cite{hansen2006tsl, evans1992fif, evans}. Making this approximation yields the following equation for the dynamics of the one-particle density distribution $\rho(x,t)$:
\begin{eqnarray}
\label{eq3}
\frac{\partial \rho(x,t)}{\partial t} = \frac{\partial}{\partial x}\left[ \rho(x,t)\frac{\partial }{\partial x}\frac{\delta F[\rho(x,t)]}{\delta \rho(x,t)}\right].
\end{eqnarray}
For the case with periodic BC, the Helmholtz free energy functional $F[\rho]$ is of the form \cite{evans1992fif, hansen2006tsl}:
\begin{eqnarray}
\label{dft_energy}
F[\rho(x,t)] &=& T \int_{-\frac{S}{2}}^{\frac{S}{2}} dx\,\rho(x,t)[\ln{\rho(x,t)}-1] \nonumber \\
&+& \int_{-\frac{S}{2}}^{\frac{S}{2}} dx\,U_{\rm eff}(x,t)\rho(x,t) \nonumber \\
&+& F_{\rm hc}[\rho] + F_{\rm at}[\rho],
\end{eqnarray}
where the first term on the right hand side is the ideal-gas contribution to the free energy, $F_{\rm hc}$ is the excess contribution to the free energy due to the hardcore repulsion between the particles, and $F_{\rm at}$ represents the contribution due to the attractions between the particles. In a mean-field approximation \cite{evans1992fif}, $F_{\rm at}$ is given by
\begin{equation}
\label{eq:F_at}
F_{\rm at}[\rho] = \frac{1}{2}\int_{-\frac{S}{2}}^{\frac{S}{2}} dx \int_{x-\frac{S}{2}}^{x+\frac{S}{2}} dx'\,w_{\rm at}(\mid x-x' \mid)\rho(x)\rho(x').
\end{equation}
The exact expression for the equilibrium excess Helmholtz free energy for hard rods of length $h$, $F_{\rm hc}[\rho]$, was first presented in Ref.\,\cite{percus76}. The result is:
\begin{eqnarray}
\label{hr_energy}
F_{\rm hc}[\rho] = \frac{1}{2}\int_{-\frac{S}{2}}^{\frac{S}{2}} dx\, \phi[\rho(x)] \left\{\rho\left(x+\frac{h}{2}\right)+\rho\left(x-\frac{h}{2}\right)\right\},
\end{eqnarray}
where $\phi[\rho(x,t)] = -T\ln{[1-\eta(x,t)]}$ and $\eta(x,t) = \int_{x-h/2}^{x+h/2}dx'\,\rho(x',t)$. It should be emphasized that the functional in Eq.\,(\ref{hr_energy}), is strictly only exact for 1D equilibrium systems of hard-rods treated in the grand canonical ensemble \cite{percus76, evans1992fif}. However, as the (average) number of particles in a system is increased, the difference between results from treating a system canonically or grand canonically diminishes, so that the theory can safely be extended to describe single-files with a fixed but large numbers of rods.
In earlier work, the DDFT approach was used to study the dynamics of an ensemble of pure hard rods (i.e.\ with no attractive interactions) \cite{marini99, marini00, penna03}. More recently \cite{PA10}, we have applied the DDFT formalism to describe a file of hard rods interacting via a pair potential with an attractive contribution. In both cases, the free energy functional given in Eq.\,(\ref{hr_energy}) was shown to reproduce fairly closely the results from Brownian dynamics computer simulations [i.e.\ results from numerically integrating Eqs.\ \eqref{langevin}], even for relatively small numbers of particles, $N \gtrsim 10$.
Using Eqs.\,(\ref{dft_energy}) and (\ref{hr_energy}), Eq.\,(\ref{eq3}) can be rewritten in the form of a conservation law, i.e. in terms of the instantaneous current density $J(x,t)$,
\begin{eqnarray}
\label{eq:continuity}
\frac{\partial \rho(x,t)}{\partial t} &=& -\frac{\partial J(x,t)}{\partial x},
\end{eqnarray}
where
\begin{eqnarray}
\label{current}
J(x,t) &=& \rho(x,t)\left[ -T \frac{\partial }{\partial x}\ln \rho(x,t) -\frac{dU(x)}{dx}+A(t) \right.\nonumber \\
&-& T\left( \frac{\rho(x+h,t)}{1-\eta(x+h/2,t)} - \frac{\rho(x-h,t)}{1-\eta(x-h/2,t)}\right) \nonumber \\
&-& \left. \int_{x-\frac{S}{2}}^{x+\frac{S}{2}}dx^\prime\,\rho(x^\prime,t)\frac{\partial w_{\rm at}}{\partial x}(x-x^\prime) \right].
\end{eqnarray}
Before we discuss our results for solutions of Eqs.~(\ref{eq:continuity}) and (\ref{current}), we would like to point out that we see many parallels between the dynamics described by the partial integro-differential Eqs.~(\ref{eq3}) with (\ref{dft_energy}) and the dynamics of partially wetting drops and films on solid substrates described by so-called thin film equations (fourth order partial differential equations) \cite{KaTh07}. Formal similarities between thin film equations and some Fokker-Planck equations for interacting particles have recently been pointed out in the case of spatially-asymmetric ratchets with a temporal AC drive \cite{ThJo10}. In the present case, we find that some of our results are similar to results for the case of liquid drops on horizontal \cite{TBBB03} and inclined \cite{dep2} heterogenous solid substrates. These similarities arise due to (i) the similar gradient dynamics form of the evolution equation for the conserved field (here $\rho$), (ii) the presence of similar physical effects. The role of the attractive and repulsive force between particles is taken by the partial wettability of the liquid \cite{deGe85}; the stabilizing role of diffusion is played by surface tension; the channel corrugations are similar to substrate heterogeneities; the DC drive corresponds to constant driving parallel to the substrate (e.g., drop on an incline); and the present AC drive is similar to substrate vibrations \cite{JoTh10} or oscillating electric fields \cite{JoTh07}.
Here, we solve Eq.\,(\ref{eq:continuity}) imposing periodic BC on the domain $x\in [-S/2,S/2]$ which is centered at the origin, $x=0$. Due to the external driving force the system remains permanently out of thermodynamic equilibrium. Therefore, in an infinite domain or in a finite domain with periodic BC, the non-equilibrium dynamics of the system with DC drive is not relaxational, in contrast to the case of zero-flux BC that would correspond to a closed finite system \cite{marini99}. Note, however, that the channel corrugations may result in local equilibria. An AC drive keeps the system out of equilibrium for any BC. The non-relaxational character is reflected by the finding that with periodic BC, the free energy $F[\rho(x,t)]$ in Eq.~(\ref{dft_energy}) is not necessarily a monotonically decreasing function of time, i.e., it does not play the role of a Lyapunov functional for the gradient dynamics Eq.~(\ref{eq3}). This can best be seen by considering the total time derivative of the free energy
\begin{eqnarray}
\label{dfdt}
\frac{d F[\rho]}{d t} &=& \int_{-\frac{S}{2}}^{\frac{S}{2}}\frac{\delta F}{\delta \rho }\frac{\partial \rho(x,t)}{\partial t}\,dx \nonumber \\
&=& \int_{-\frac{S}{2}}^{\frac{S}{2}}\frac{\delta F}{\delta \rho }\frac{\partial}{\partial x}\left[\rho(x,t)\frac{\partial }{\partial x}\frac{\delta F}{\delta \rho}\right] \,dx \nonumber \\
&=&\left[\frac{\delta F}{\delta \rho}\rho \frac{\partial }{\partial x} \frac{\delta F}{\delta \rho}\right]_{-\frac{S}{2}}^{\frac{S}{2}}- \int_{-\frac{S}{2}}^{\frac{S}{2}}\rho \left[\frac{\partial }{\partial x}\frac{\delta F}{\delta \rho}\right]^2\,dx.
\end{eqnarray}
where we have used Eq.\,(\ref{eq3}) and integrated by parts. The functional derivative $\delta F/\delta \rho$ is not periodic in $x$, due to the ``tilted" effective potential $U_{\rm eff} = U(x)-A(t)x$, and so the boundary term in the last line of Eq.\,(\ref{dfdt}) does not normally vanish. It is equal to $AS J_{S/2}>0$, where $J_{S/2}$ is the current density on the boundary, and $J_{S/2}<0$ for $A<0$.
It then follows that for $A \not= 0$ the time derivative $d F[\rho(x,t)]/d t$ is not necessarily a negative quantity. This allows for time-oscillatory behavior of the system, even in the long-time limit. In other words, Eq.\,(\ref{eq:continuity}), with $A(t)=A$, may admit stable cyclo-stationary or time-periodic solutions, the existence of which will be discussed in the next section.
For stable time-periodic solutions, the average particle current $\bar{J}$ is obtained from Eq.\,(\ref{current}) by averaging $J(x,t)$ over position and over time, namely,
\begin{eqnarray}
\label{eq8}
\bar{J} = \frac{1}{\tau}\int_{-\frac{S}{2}}^{\frac{S}{2}}dx\,\int_{t^\prime}^{t^\prime + \tau}dt\,J(x,t),
\end{eqnarray}
where $\tau$ is the period of the oscillations. Throughout this paper we will use the average current per particle $J$, which is related to the total current in Eq.\,(\ref{eq8}) via $J = \bar{J}/N$.
\section{Spontaneous condensation
\label{cond}
In the absence of the channel potential and without AC or DC drive, i.e for $U(x)_{\rm eff}=0$, the DDFT equation (\ref{eq3}) admits a stationary homogeneous solution $\rho(x)=\rho_0$, with constant average density $\rho_0 = N/S$. Because of the competition between the destabilizing attractive forces and the stabilizing effect of the thermal motion of the particles (diffusion), the homogeneous density distribution is not always stable. In order to minimize their free energy, the attracting particles tend to condense into a set of clusters. This tendency is opposed by the stabilizing action of diffusion, which leads to a spreading of the particles away from one another. Depending on which process dominates, the homogeneous state $\rho(x)=\rho_0$ may either be (linearly) stable or unstable. Note that in this case the dynamics is relaxational and the functional (\ref{dft_energy}) is a Lyapunov functional [cf.~Eq.~(\ref{dfdt})].
One may alternatively take a thermodynamic rather than a dynamical point of view for understanding this instability: Recall that the Helmholtz free energy of the system $F={\cal U}-T{\cal S}$, where ${\cal U}=\langle\Phi\rangle$ is the internal energy, ${\cal S}$ is the entropy of the system and $\langle \cdots \rangle$ denotes a statistical average \cite{hansen2006tsl}.
At equilibrium, $F$ is minimal. At high temperatures $T$, this is achieved by maximizing ${\cal S}$ (i.e.\ by dispersing the particles throughout the system in a maximally disordered way), since the term $-T{\cal S}$ is the dominant contribution to the free energy at high temperatures $T$. However, at low temperatures, the internal energy contribution ${\cal U}$ dominates the free energy $F$ and so the attracting particles minimize the free energy by gathering together to minimize the internal energy ${\cal U}$.
We model the attractive contribution to the pair potential between the particles by a simple exponential function of the form
\begin{eqnarray}
\label{attractive_potential}
w_{\rm at}(x) =-\alpha \exp{(-\lambda x)},
\end{eqnarray}
where the parameters $\alpha$ and $\lambda$ characterize the strength of attraction and the attraction range $l_{\rm at} = 1/\lambda$, respectively. It should be noted that the mean field approximation for the contribution to the Helmholtz free energy due to the attractive interactions \eqref{eq:F_at} used here, is only quantitatively reliable when $\lambda h \lesssim 1$, i.e.\ when any given particle is interacting with several of its neighbors so that a mean-field approximation is appropriate. However, even when the attraction range is somewhat shorter than this, the mean-field approximation remains qualitatively correct.
In order to study the stability of the homogeneous state, we linearize Eq.\,(\ref{eq3}) using the standard plane wave anstatz $\rho(x,t) = \rho_0 + \varepsilon e^{\beta(k)t + ikx}$, where the sign of the growth rate $\beta(k)$ determines the stability of the homogeneous solution $\rho_0$. Note that for the stable fluid the dispersion relation $\beta(k)$ is closely related to the static structure factor $S(k)$, via $\beta(k)=-T k^2/S(k)$ \cite{archer04}. Neglecting exponentially small terms of the order of $\exp{(-\lambda S)}$, we obtain to leading order in $\varepsilon$
\begin{eqnarray}
\label{disp}
\beta(k) &=& -Tk^2 - \frac{2Tk\rho_0\sin{(kh)}}{1-\rho_0 h} -\frac{4 T\rho_0^2[\sin{(kh/2)}]^2}{(1-\rho_0 h)^2} \nonumber \\
&+& \frac{2\alpha \lambda \rho_0 k^2}{k^2+\lambda^2}.
\end{eqnarray}
\begin{figure}[t]
\centering
\includegraphics[width=0.48\textwidth]{newFig1.eps}
\caption{(Color online) (a) Typical dispersion relations $\beta(k)$ close to the onset of the spinodal mode (sp), as shown by dashed lines and the freezing mode (fr), as shown by solid lines. The $\beta(k)$ curves near the spinodal instability are for $\rho_0=0.2$ and $\alpha=10$, $11.8$ and $17$, which are close to the onset of the instability. The curves for the freezing instability are for $\rho_0=0.6$ and $\alpha=4$, $4.9$ and $6$. The other parameters are $T=1$, $\lambda=3$, $h=1$. (b-c) Stability diagram for the system with uniform constant density $\rho(x)=\rho_0$, in the plane spanned by $\widetilde{T}=Th\lambda/\alpha$ and $\tilde\rho=\rho_0 h$ for (b) $\xi=\lambda h=3$ and (c) $\xi=1.5$. The labels ``sp'' and ``fr'' denote (hatched) regions where the uniform density is only linearly unstable to the spinodal and freezing instability, respectively. In the cross-hatched region both modes are unstable.\label{F0}}
\end{figure}
Inspection of Eq.\,(\ref{disp}) shows that two different scenarios exist where the homogeneous density $\rho_0$ is linearly unstable. On the one hand, the solution $\rho(x,t)=\rho_0$ can become unstable via the standard spinodal phase separation mechanism where the material separates into regions of low density (gas) and hight density (liquid), which we call the ``spinodal mode'' (denoted by ``sp''). It is associated with an instability against harmonic perturbations with wave numbers $0<k<k_0$, where $k_0$ is an upper value obtained by solving the equation $\beta(k_0)=0$. The spinodal instability sets in when the leading coefficient of the expansion of $\beta(k)$ in powers of $k^2$ (i.e., the coefficient of the $k^2$ term) vanishes, i.e., the wave number at onset is zero. This corresponds to a type $II_s$ instability in the classification of Cross and Hohenberg \cite{CrHo93}. Based on the Taylor expansion of the relation (\ref{disp}) one obtains $\widetilde T=2\tilde\rho\,(1-\tilde\rho^2)$ as the condition for the onset of the spinodal mode [cf.~Fig.~\ref{F0}(b)]. Here, we have introduced the reduced temperature $\widetilde{T}=Th\lambda/\alpha$ and the reduced density $\tilde{\rho}=\rho_0 h$. The critical point for decomposition is found at $\widetilde{T}_c= 8/27$ and $\tilde{\rho}_c=1/3$.
The stability condition is equivalent to the thermodynamic stability criterion requiring that the isothermal compressibility be negative. This corresponds to the Helmholtz free energy per unit length of the system becoming concave, namely, the boundary of the spinodal instability is given by $\delta^2 F[\rho] / \delta \rho^2|_{\rho_0} =0$. When $\delta^2 F[\rho] / \delta \rho^2|_{\rho_0} >0$ the uniform solution is linearly stable and when $\delta^2 F[\rho] / \delta \rho^2|_{\rho_0} <0$ the uniform solution is linearly unstable. The dispersion relation $\beta(k)$ close to the onset of the spinodal mode is shown in Fig.\ref{F0}(a) by dashed lines, which correspond to three different values of the attraction strength $\alpha$, chosen close to the spinodal instability threshold as obtained from Eq.\,(\ref{disp}).
On the other hand, the solution $\rho(x,t)=\rho_0$ can become unstable via a freezing mode of the system, where the particles become localized and the density profile exhibits a series of sharp peaks separated by distances smaller than $L$. This instability corresponds to type $I_s$ in the classification of Ref.~\cite{CrHo93}. In the context of reaction-diffusion systems it is sometimes referred to as a Turing instability \cite{Nico95}. This mode (which we denote by ``fr'') sets in at a non-zero critical wave number $k_c$, as illustrated in Fig.\ref{F0}(a) by the solid lines for $\rho_0 =0.6$. Beyond onset, the freezing mode gives rise to the growth of periodic modulations in the density profile with a wavelength $ \approx 2\pi/k_c$.
In Figs.\,\ref{F0}(b) and (c) we plot the linear stability diagram of the system with uniform density $\rho_0$, in the plane spanned by $\widetilde T =Th\lambda/\alpha$ and $\tilde\rho=\rho_0 h$ for the interaction length ratio (b) $\xi\equiv\lambda h=3$ and (c) $\xi=1.5$. The labels ``sp'' and ``fr'' mark regions where the uniform system is linearly unstable to the spinodal and freezing mode, respectively. As discussed above, the onset of the spinodal mode only depends on the reduced temperature $\widetilde T$ and reduced density $\tilde\rho$. In contrast, the onset of the freezing mode also depends on the value of $\xi$. For relatively long rods with $\xi=3$, where the attraction range is short compared to the core size $h$, the region of the freezing instability extends down to moderate values of the reduced density $\tilde\rho\approx 0.5$. For a reduced temperature above $\widetilde T_c=8/27$ only the freezing instability exists at large and moderate $\tilde\rho$. For small values of $\xi$, corresponding to the attraction range $\lambda^{-1}$ being large compared to $h$, and for which the mean-field approximation for the free energy used is expected to be most reliable, the freezing mode is only found at extremely high packing fractions, $\tilde\rho \approx 1$. At such high densities, the critical wave number $k_c$ of the freezing mode is approximately $k_c \approx 2\pi/h$, giving rise to the formation of density peaks separated by a distance $\approx h$, as one would expect for a frozen system.
Below $\widetilde T_c$, the spinodal mode exists for a range of $\tilde\rho$ that with decreasing $\widetilde T$ extends on both sides of the critical value $\tilde\rho =1/3$. This implies that the spinodal mode sets in at smaller and smaller values of the density as $\widetilde T$ is decreased. At large $\tilde\rho$ there exists a region where both linear modes are unstable [cross-hatched in Figs.\,\ref{F0}(b) and (c)]. The region where only the freezing mode exists is shifted towards higher densities as $\widetilde T$ decreases. The definition of $\widetilde T$ implies that for any given physical temperature $T$, a decrease in the interaction strength $\alpha$ below the threshold value $\alpha_c=27Th\lambda/8$ stabilizes the spinodal mode.
For any finite temperature the system can be quenched into the freezing unstable region by increasing the average density in the system.
\section{DC drive}
\label{DC}
Having considered the stability of the uniform system, we now consider the non-uniform system that is subject to a periodic external potential $U(x)$ and the DC driving force $A$. We model the periodic potential, induced by the corrugated channel walls, by the standard bi-harmonic ratchet potential $U(x)=\sin{(2\pi x)}+0.25\sin{(4\pi x)}$ \cite{RMP09}. Recall that to drag a single particle over one of the barriers in $U(x)$, one must apply a force $A_R=3\pi$ to pull the particle over the barrier to the right and a force $A_L=3\pi/2$ to pull it to the left. $A_{L}$ and $A_R$ are termed the left and the right depinning thresholds, respectively \cite{RMP09}. Note that all the results reported in this section for a DC drive remain qualitatively valid even for a simpler symmetric periodic external potential, such as $U(x)=\sin{(2\pi x)}$. The case of a periodic external potential without and with DC drive shows some similarities to liquid drops/films on periodically heterogeneous substrates without \cite{TBBB03} and with \cite{dep1} a driving force parallel to the substrate, respectively. The former case will be explored elsewhere. Below we discuss similarities and differences for the case with DC driving.
\subsection{Zero rod length and finite interaction range}
\label{zerolength}
As a reference system we first consider a file of point-like particles, i.e., with $h=0$, interacting solely via the exponential soft core potential $w_{\rm at}(x)$ and driven by a DC external force. When the characteristic interaction range $l_{\rm at}$ between the point-like rods is small, i.e.\ when $l_{\rm at}=1/\lambda \rightarrow 0$, a local approximation can be made for the dynamical equations for the system, as shown in Refs.\,\cite{sav04,sav05}. In this limit, the integral involving $w_{\rm at}$ in Eq.\,(\ref{current}) can be reduced to a local function of the form $\sim g\rho(x,t)\partial \rho(x,t)/\partial x$, where the coefficient $g$ is a parameter determined by the strength of the interactions between the particles. In Refs.\,\cite{sav04,sav05} it was shown that when $g$ is increased beyond a certain critical value, the density distribution of the attracting particles exhibits a spontaneous symmetry breaking transition, where the stationary periodic density profile with period $L$ [the period of the modulations in $U(x)$] becomes unstable and evolves toward a stable stationary distribution with period $S$ (the total system length). We now go beyond the analysis of Refs.\,\cite{sav04,sav05} and consider such a symmetry breaking mechanism for particles interacting via a potential with a nonzero interaction range, $l_{\rm at} \not= 0$.
For convenience we fix the average particle density to be $\bar \rho L = 1$, corresponding to one particle per period $L$ of the channel, and we set the constant drive, $A=-1$. Using the numerical continuation package {\it AUTO} \cite{AUTO}, we follow the branch of solutions corresponding to a stationary density distribution $\rho_s(x)$, that originates from the stationary density profile for the case when $\alpha=0$ (i.e.\ a non-interacting ideal-gas of particles). Note that in the long-time limit, the density profile for the ideal-gas remains stable and stationary, regardless of the form of the channel potential, $U(x)$, the magnitude of the drive, $A$, or the temperature of the system, $T$ \cite{risken,reimann02}.
We determine the stationary solutions of Eq.\ (\ref{eq:continuity}) with the current given by Eq.\,(\ref{current}) that contains nonlocal terms, by means of the Fourier mode method described in Ref.\,\cite{bordyugov07}. The density profile is discretized over the domain $[-S/2,S/2]$, derivatives are obtained using finite difference approximations, and the non-local terms are calculated using a Fast Fourier transform. We start from the equation for the stationary solution of Eq.\,(\ref{eq:continuity}), $\partial J(x,t)/\partial x =0$, which is then written as a set of algebraic equations for the Fourier components of the current $J(x,t)$ and from this we obtain our solutions for the stationary density profile $\rho_s(x)$. Using the continuation package {\it AUTO} allows us to detect the presence of Hopf bifurcations as well as to trace the solution branches for both the stationary solutions and the time-periodic ones that emerge from them.
We begin by discussing the bifurcation diagrams of the stationary solutions of Eq.\,(\ref{eq:continuity}) on varying the interaction strength, $\alpha$, for a fixed value of the range parameter of the pair potential, $\lambda=5$, and the fixed $U(x)$. These are shown in Figs.\,\ref{F1}(a)-(c), for three different systems with lengths, $S=2L$, $3L$ and $4L$, respectively. The solid lines correspond to stable solutions and the dashed lines to unstable (saddle point) solutions. The labels ``HB'' and ``BP'' stand for Hopf bifurcation and branching point, respectively.
\begin{figure*}[t]
\centering
\includegraphics[width=0.98\textwidth]{newFig2.eps}
\caption{(Color online) Panels (a), (b) and (c) display bifurcation diagrams for the stationary density distributions in terms of the average current $J$ versus the interaction strength $\alpha$, for $\lambda=5$, $\bar \rho=1$, $A=-1$, $h=0$ and total system length (a) $S=2L$, (b) $3L$ and (c) $4L$. The solid and dashed lines correspond to stable and unstable solutions, respectively. The points labeled ``HB'' and ``BP'' denote Hopf bifurcation and branching points, respectively. Panels (d), (e) and (f) display selected corresponding density profiles. The solid lines are profiles with spatial periods $2L$, $3L$ and $4L$, respectively, for $\alpha=2.2$. The dashed lines represent the unstable solution with period $L$. In (f) the solid and dotted lines represent all four possible $4L$-periodic solutions. \label{F1}}
\end{figure*}
As the interaction strength is increased beyond a critical value $\alpha_c$, the $1L$-periodic solution that is stable for small $\alpha$ becomes unstable either via a (period-doubling) pitchfork bifurcation (for $S=2L$), or via a Hopf bifurcation (for $S>2L$). In the case of the pitchfork bifurcation, displayed in Fig.\,\ref{F1}(a), a double branch of stable solutions emerges at the bifurcation point. The two branches are related by the discrete translation symmetry $x\to x+L$ and can therefore not be distinguished in Fig.~\ref{F1}(a). This new branch corresponds to a solution with a larger spatial period, equal to the system size $S=2L$, and a smaller value for the particle current $J$. One may say that for $\alpha>\alpha_c$, the periodic potential is not strong enough to pin the clusters against their natural tendency to coarsen. For $S=2L$, we know on general grounds that [in a homogeneous system without driving $A=U(x)=0$] there are two possible coarsening modes: a translation mode, where the two clusters move towards each other, and a volume transfer mode, where material is transfered from one cluster to the other \cite{KaTh07}. It is known that both are stabilized by substrate heterogeneities
exerting a strong enough pinning influence \cite{TBBB03}. Not much is known, however, for driven systems ($A\neq0$). Our DDFT simulations show that in the present system, the instability is related to the volume transfer mode of coarsening. The corresponding stable (solid line) and unstable (dashed line) stationary density profiles are displayed in Fig.\,\ref{F1}(d) for the case when $\alpha=2.2$.
The bifurcation diagram for the system with length $S=3L$ is qualitatively different from the one for the $S=2L$ case, as can be seen in Fig.\,\ref{F1}(b). One observes that the $1L$-periodic solution becomes unstable via a Hopf bifurcation. There exist branches of stationary solutions where the $x\to x+L$ translational symmetry is broken. However, they do not touch the primary branch of the $1L$-periodic solutions but are generated through a saddle-node bifurcation at the point marked by ``LP''. Solutions on these branches have period $S=3L$ and are either stable (upper branch) or unstable (lower branch). An example of a stable $3L$-periodic solution for $\alpha=2.2$ is displayed in Fig.\,\ref{F1}(e).
The bifurcation scenario for $S=4L$, displayed in Fig.\,\ref{F1}(c), is substantially more complex. The $1L$-periodic solution becomes unstable through a Hopf bifurcation. Very close to the HB point, the unstable stationary solution undergoes a primary period-doubling pitchfork bifurcation BP. Note that this first BP point for $S=4L$ coincides as expected with the BP point for $S=2L$. The newly formed $2L$-periodic solution is unstable and undergoes a further period-doubling pitchfork bifurcation at a second BP, which lies very close to the first BP, as shown in the inset of Fig.\,\ref{F1}(c).
For interaction strengths significantly larger than the critical values corresponding to the BP and the HB points, the only stable stationary solution of the DDFT equation (\ref{eq3}) has a period equal to the total system size $S$. Note that the multiplicity of the branch of the solutions with broken $x\to x+L$ symmetry depends on the total system size. For instance, for $S=4L$ there exist four such branches with spatial period $S$, as can be seen in Fig.\,\ref{F1}(f). Each solution exhibits a prominent maximum centered around one of the four minima of the external potential $U(x)$. The solutions on the 4 branches are related by the symmetry $x\to x+L$. Therefore all of them correspond to the same value for the particle current $J$ and they can not be distinguished from one another in Fig.\,\ref{F1}(c).
From the results displayed in Fig.\,\ref{F1} we may draw two important conclusions: First, the detected Hopf bifurcations of the pinned stationary solutions signals the onset of time-periodic solutions of the DDFT equation (\ref{eq3}), even in the presence of a time independent drive. Second, for certain values of the interaction strength $\alpha$, two stable stationary solutions may coexist, giving rise to current multiplicity.
\begin{figure}[t]
\centering
\includegraphics[width=0.48\textwidth]{newFig3.eps}
\caption{(Color online) Magnification of the region of Fig.\,\ref{F1}(b) close to HB point. The line connecting HB and ``hm'' corresponds to stable time-periodic solutions of Eq.\,(\ref{eq3}). The label ``hm'' stands for homoclinic bifurcation point. (Inset) Temporal period $\tau$ of the time-periodic solutions as a function of $\alpha_\mathrm{hm}-\alpha$. \label{F2}}
\end{figure}
These two findings are illustrated in detail in Fig.\,\ref{F2}, where we display a magnification of the region close to the bifurcations in Fig.\,\ref{F1}(b). In addition to the Hopf bifurcation (HB) and the saddle-node bifurcation (LP) of the stationary $3L$-periodic solutions, we display the branch of time-periodic solutions of Eq.\,(\ref{eq3}). It emerges at the HB point and terminates in a homoclinic bifurcation (labeled by ``hm'') where the time-periodic solution (limit cycle) collides simultaneously with all three unstable $3L$-periodic solution (unstable equilibria) \cite{Stro94}. The inset of Fig.\,\ref{F2} gives the temporal period $\tau$ as a function of the distance to the homoclinic bifurcation $\alpha_\mathrm{hm}-\alpha$. It shows a logarithmic dependence as expected close to a homoclinic bifurcation. We emphasize that these time-periodic solutions are stable, i.e., the corresponding Floquet multipliers are always located within the unit circle (not shown). Note that for clarity we not only suppress the branch of time-periodic solutions in Fig.\ \ref{F1}(b) but also a similar branch in Fig.\ \ref{F1}(c), for the system with $S=4L$.
In nonequilibrium driven systems, the loss of stability of the stationary solutions and the appearance of time-periodic solutions with a larger mean flow is sometimes associated with the concept of depinning. For example, in the study of liquid droplets on an inclined heterogeneous solid substrate, the dynamics of drop depinning has been studied in great detail -- see for example Refs.\ \cite{dep1,BKHT11} and references therein. In this situation the depinning is generally a transition from a steady droplet, pinned by the heterogeneity of the substrate, to a moving droplet, sliding down the incline under the action of gravity (or other driving forces parallel to the substrate). The depinning is usually investigated by increasing the driving force with all other parameters kept fixed. In such a case the dominant depinning mechanism is often related to a Saddle Node Infinite PERiod (sniper) bifurcation, although depinning via a Hopf bifurcation may also be observed in certain parameter regions \cite{dep1,BKHT11,dep2}. The depinning exhibited by the present system is observed when increasing the particle attraction $\alpha$, for a fixed value of the external drive $A$ and potential $U(x)$. This would correspond to a decrease in wettability for a droplet depinning in the thin film model.
Note also that this collective depinning is very distinct from the $T=0$ transition that is also referred to as `depinning', when the drive on a single particle exceeds either $A_L$ or $A_R$, the left and right single particle depinning thresholds \cite{RMP09}.
\begin{figure}[t]
\centering
\includegraphics[width=0.48\textwidth]{newFig4.eps}
\caption{(Color online) (a) Snapshots of the time-periodic density, $\rho(x,t)$, for $\alpha=2.08$ at three times, $t_0$ (dashed line), $t_0+\tau/3$ (dot-dashed line), and $t_0+2\tau/3$ (solid line). The temporal period of the solution is $\tau=30$; the remaining numerical parameters are as in Fig.\,\ref{F1}(b). The channel potential $U(x)$ is also displayed (heavy solid line limiting the shaded area). (b) The time dependent current $J(t) = (1/N)\int_{-S/2}^{S/2}J(x,t)\,dx$, corresponding to the solution in (a). \label{F2a}}
\end{figure}
At the HB point, the newly formed stable time-periodic solution has a finite period, as it can be seen from Fig.\,\ref{F2}. In order to illustrate the dynamics of the depinning of the stationary solution, we set $\alpha=2.08$ and plot in Fig.\,\ref{F2a}(a) snapshots of the time-periodic solution $\rho(x,t)$ at three subsequent times, $t_0$, $t_0+\tau/3$, and $t_0+2\tau /3$, where $t_0$ was chosen as described below and $\tau$ is the temporal period of the solution. Inspection of the density profiles indicates that the depinned solution can be seen as a superposition of two parts: a stationary part with spatial period $L$ and a time-periodic part with spatial period $S=3L$, that slides `on top' of the stationary part. The time-periodic part corresponds to a wave traveling to the left, that is, in the direction of our negative constant drive. Here, at $t=t_0$, the absolute maximum of the density profile is located at the rightmost minimum of the channel potential, $U(x)$. After one third of the temporal period $\tau$, the absolute maximum has moved to the central well of the channel and after two thirds of $\tau$, the maximum has finally reached the leftmost well. After one full period $\tau$, the cycle is repeated.
The time-periodic solution changes its character along the branch in a continuous manner. With increasing attraction strength the amplitude of the time-periodic part becomes larger as compared to the steady part until finally most of the particles travel. They travel, however, not in the form of a translation of a compact cluster, but rather in the form of a volume transfer of the cluster from one potential well to the next. The temporal period becomes larger with increasing $\alpha$ and the overall flux oscillates between a low absolute value (when the cluster sits in a well) and a large absolute value (when the cluster is transferred to the next well). This is shown in Fig.~\ref{F2a}(b). With increasing $\alpha$ the dependence of the flux on time becomes increasingly non-harmonic as the cluster spends an increasing fraction of the time period around the three maxima of the channel potential. In the vicinity of the homoclinic bifurcation the density profiles for clusters mainly localised at one of the three maxima closely resemble the corresponding profiles on the three unstable stationary $3L$-periodic solutions. This also implies that at the homoclinic bifurcation the stable cycle collides with all three unstable equilibria at once.
Summarizing the results displayed in Figs.\,\ref{F2} and \ref{F2a}, we conclude that, for values of $\alpha$ between the points labeled by LP and HB, there exist two stable stationary solutions, with spatial period $L$ and $3L$, respectively. Moreover, between the points HB and hm, a stable time-periodic solution coexists with the stable stationary $3L$-periodic solutions.
By perturbing the time-periodic density profile with a finite amplitude disturbance, one can induce the transition to the stable stationary $3L$-periodic solution. To do so, one starts a simulation in time with a stable time-periodic solution and adds a finite (mass-conserving) perturbation. If the perturbation is large enough, the solution evolves after a short transient toward the stable 3L periodic stationary solution. Note also that we were not able to find the opposite transition: Perturbing the stable stationary $3L$-periodic solution by shifting it slightly in the direction of the drive will `depin' the cluster only for a short transient. It moves to the left and settles into the next potential well, i.e., it moves to the stable stationary $3L$-periodic branch that is related by the translation $x\to x-L$.
\subsection{Finite rod length and short range attraction}
\label{finitelength}
We discuss now the effects of having a finite rod length in addition to the attraction between the particles. In Fig.\,\ref{F4a}(a) we display the bifurcation diagram in terms of the stationary current $J$ as a function of $\alpha$ for rod lengths $h=0, 0.1$, and $0.2$ for a domain length $S=4L$. A relatively small change in the size of the rods is sufficient to cause a significant change in the current. First, one observes that the magnitude of the current at $\alpha=0$ increases with $h$; this also remains true for $\alpha>0$. Second, the critical interaction strength, $\alpha_c$, at which the $1L$-periodic solution looses its stability via a Hopf bifurcation, increases with $h$; i.e.\ as expected a system of finite length rods is more stable than the reference system with $h \to 0$. Third, the region in parameter space in which the stationary $1L$-periodic and the $4L$-periodic solutions coexist, shrinks as $h$ is increased.
\begin{figure*}[t]
\centering
\includegraphics[width=0.9\textwidth]{newFig5.eps}
\caption{(Color online) (a) Current $J$ versus $\alpha$ for $h=0$, 0.1 and 0.2, as indicated on the curves. The other system parameters are $T=1$, $\lambda=5$, $S=4L$, $A=-1$ and $\bar\rho=1$. The symbols denote the Hopf bifurcation (HB) and the branching points (BP). The primary bifurcation is always a HB. The solid and dotted lines for $h=0.1$ and $h=0.2$ represent stable and unstable solution branches, respectively. For $h=0.2$, the branch originating from the HB point is the branch of stable time-periodic solutions. Next to each stable branch is a label indicating the spatial periodicity of the corresponding solutions. (b) displays a magnification of the region in the vicinity of the HB point for $h=0.1$. The branch of stable time-periodic solutions starts at the HB point and terminates at the hm point. Panel (d) shows the average free energy $\langle F/S \rangle$ for $h=0.1$ and the same range of values of $\alpha$ as in (b). Panel (c) shows the free energy of the various solution branches, for $h=0.2$. Panel (e) shows a snapshot of a typical time-periodic solution, obtained for parameters as in (c) and $\alpha=8$. Panel (f) represents the temporal period of the stable time-periodic solutions in (b) as a function of the distance from the hm point, i.e. $(\alpha_\mathrm{hm}-\alpha)$. Dashed line is the decay law $\tau \sim \ln{(\alpha_\mathrm{hm}-\alpha)}$. \label{F4a}}
\end{figure*}
This can be explained as follows: For the $4L$-periodic solution to be stable at relatively small values of $\alpha$, one must squeeze all the particles (there are $4$ particles in the system with length $S=4L$ and $\bar\rho=1$) into a small part of the total system, not larger than half a ratchet period, $4h < L/2$. As a consequence, the critical rod length above which the $4L$-periodic and the $1L$-periodic solutions are unlikely to coexist, is approximately $h=0.125$, for $L=1$. A magnification of the bifurcation diagram for $h=0.1$ slightly below this critical value is displayed in Fig.\,\ref{F4a}(b).
There, the stationary $1L$-periodic solutions become unstable at the Hopf bifurcation (HB) and a stable branch of time-periodic density profiles emerges supercritically [heavy green solid line in Fig.\,\ref{F4a}(b)]. Slightly beyond the Hopf bifurcation, the unstable branch of stationary $1L$-periodic solutions undergoes a supercritical period-doubling pitchfork bifurcation (BP) $(\alpha\approx 3.35)$. The emerging branch of stationary $2L$-periodic solutions is unstable w.r.t.\ two modes. It becomes more unstable at a secondary period-doubling pitchfork bifurcation at $(\alpha\approx 3.45)$ (BP). The bifurcating branch consists of stationary unstable $4L$-periodic solutions with 2 unstable eigenmodes. One of them is stabilized at a first saddle-node bifurcation at $\alpha\approx 3.55$ where the branch turns back toward smaller $\alpha$. The branch of stationary $4L$-periodic solutions finally becomes stable at another saddle-node bifurcation at $\alpha\approx 3.47$, where it turns again towards larger $\alpha$. The
branch of stable time-periodic solutions terminates as in the case of $h=0$ length rods in a homoclinic bifurcation on the branch of unstable stationary $4L$-periodic solutions. The exact location of the homoclinic bifurcation (labeled ``hm'') is very close to (but numerically clearly distinguished from) the saddle-node bifurcation. The temporal period of the time-periodic solutions diverges logorithmically on approaching the ``hm'' point, as shown in Fig.\,\ref{F4a}(f).
To obtain some indication as to which stable solution might be selected in time evolutions of the DDFT, starting from various initial states, we compute the (time-averaged) Helmholtz free energy per unit length, $\langle F/S \rangle$, for all stable solutions. They are displayed in Fig.\,\ref{F4a}(c) and (d) for $h=0.2$ and $h=0.1$, respectively. In calculating these, we subtract the non-periodic potential energy term, $\int_{-S/2}^{S/2} A \rho(x)\,dx$, associated with the DC drive, from the full expression in Eq.\,(\ref{dft_energy}). This ensures that solutions on branches that are related by the discrete $x\to x+L$ translational symmetry have an identical value for the free energy under periodic BC.
For $h=0.2$, we observe in Fig.\,\ref{F4a}(a) that there exists no branch of stationary $4L$-periodic solutions; instead the stable branch of $2L$-periodic solutions continues toward large $\alpha$. Note that this branch is unstable when it bifurcates from the $1L$ solutions, but becomes stable as a result of a another Hopf bifurcation at $\alpha\approx 5.6$. The emerging time-periodic branch is unstable and will not be further considered here. The only stable solutions with spatial period equal to the system size, $S=4L$, are the time-periodic ones, which correspond along most of the branch to a single compact cluster of particles traveling in the direction of the drive. Close to the Hopf bifurcation, it resembles a small amplitude wave moving `on top' of the stationary $1L$ state. Further away from the bifurcation the behaviour resembles the one described above in connection with Fig.~\ref{F2a}: Most of the particles travel in the form of a volume transfer of the cluster from one potential well to the next. Increasing $\alpha$ further, at about $\alpha\approx7$ the flux increases by about 50\% over a very small $\alpha$-range. And the cluster morphology also changes from a compact ``drop-like'' shape to a multi-hump localized structure as depicted in Fig.\,\ref{F4a}(e), with an arrow indicating the direction of motion of the cluster. Each hump corresponds to a single particle. The particles in the cluster are strongly bound together and the distance between the particles remains almost constant as the cluster moves through the system as a single unit. This implies that at $\alpha\approx7$ the transport mode also changes from a volume transfer mode or to a translation mode.
In contrast to the case $h=0.1$, the time-periodic branch continues toward large $\alpha$. In other words, for $h=0.2$ all $4L$-periodic solutions are depinned. In Fig.\,\ref{F4a}(c) we see that time-periodic solutions have on average a lower free energy than the stable stationary $2L$-periodic solutions. However, as the system is permanently out of equilibrium, in general, the solution of lower free energy is not necessarily the one that the system converges to in the long time $t \to \infty$ limit.
Thus, for $h=0.2$ the onset of the time-periodic solutions of the DDFT equation is associated with a transition between two major transport modes: (i) At small values of the attraction strength $\alpha$ or, equivalently, for high temperatures, stationary density distributions exist, with the particles uniformly distributed among the wells of the channel potential. Under the action of the stochastic (thermal) noise, the particles jump occasionally either to the right or to the left, but with a higher probability for jumps in the direction of the applied drive. One may call this the ``stationary mode''. (ii) At larger $\alpha$ (or smaller temperature), time-periodic density profiles seem to dominate. They either correspond to transport from well to well by a volume transfer mode or by a translation mode. The latter
correspond to depinned compact clusters in which strongly attracting particles travel together. One may call this the ``condensed traveling mode''. Such a traveling cluster has a characteristic length $\sim hN$ and, in the limit where the attraction $\alpha$ is strong (i.e.\ when $\alpha \gg T$), it moves as a whole in the direction of the drive.
Fig.\,\ref{F4a}(a) shows that the magnitude of the average particle current $J$ is substantially larger (at the same $\alpha$) when transport occurs through the condensed traveling mode, than when in the stationary mode. This can be understood by noticing that for well separated particles, which are effectively not interacting, the average drifting motion of the particles is only resisted by the periodic channel potential. However, when $N$ particles are clustered (bonded) together, then the total pinning force exerted by the channel walls on the cluster is $f=-\sum_{i=1}^N dU(x_i)/d x_i$. As we show in detail below, the value of this net force is very sensitive to the cluster size, and when the length of the cluster $hN$ is an integer multiple of the period of the channel potential $L$, the total pinning force on the cluster vanishes, leading to a maximal drift velocity equal to $A$ \cite{PA10}.
\section{Low frequency AC drive}
\label{acdrive}
In this section we discuss the behavior of the system when driven by an unbiased AC (square-wave) drive $A(t)=A\,{\rm sgn}\left[ \cos{(\omega t)} \right]$ in the low frequency limit, i.e.\ in the limit $\omega \rightarrow 0$. We focus in particular on the behavior of the average rectification current $\langle J \rangle $. For vanishingly small frequencies, $\langle J \rangle$ is obtained as the arithmetic average of the two unidirectional currents $J^+$ and $J^-$, with $J^{\pm}$ denoting the average currents induced by the DC drives $\pm A$.
\subsection{Maximization of the rectification current}
\label{currmax}
\begin{figure}[t]
\centering
\includegraphics[width=0.48\textwidth]{newFig6.eps}
\caption{(Color online) (a) The unidirectional currents $J^{\pm}$ as functions of $\alpha$, calculated using the DDFT, for $S=5L$, $N=5$, $A=\pm 1$, $\lambda=3$, $T=0.5$, and for $h=0.2$ (solid line) and $h=0.16$ (dashed line). Symbols ``Solid squares'' mark the corresponding Hopf bifurcation of the stationary density distribution for $h=0.2$. Note that to left of the Hopf bifurcation we show the current for the stationary solutions and to the right for the time-periodic solution.
(b) The rectification current $\langle J \rangle$ versus $\alpha$, computed as the arithmetic mean of $J^+$ and $J^-$ in (a) (curves) and from direct Brownian dynamics simulations with $h_{\rm eff}=0.16$ (symbols). All other system parameters are as in (a). Panel (c) gives the unidirectional currents $J^{\pm}$ obtained from Brownian dynamics simulations as functions of $\alpha$ for $T=0.5$ (solid line) and $T=0.2$ (dashed line). The remaining parameters are $S=10$, $N=10$, $A=\pm 1$, $\lambda=3$, and $h_{\rm eff}=0.16$. (d) The rectification currents $\langle J \rangle$, computed as the arithmetic mean of the currents $J^+$ and $J^-$ in (c). \label{F5}}
\end{figure}
As shown in the previous section, for constant drive $A$, increasing the pair attraction strength $\alpha$, leads to the onset of a condensed traveling transport mode associated with the clustering of the particles traveling in the direction of the drive. As the condensation sets in, the opposite unidirectional currents $J^{\pm}$ increase in magnitude. However, due to the asymmetry of the channel potential, the condensation sets in at different values of $\alpha$, depending on the orientation of the drive. This phenomenon is illustrated in Fig.\,\ref{F5}(a), where the two relevant HB points are marked, for the case when $h=0.2$. Owing to the spatial asymmetry of $U(x)$, the depinning of the stationary density profile when the drive is $-A$, with current $J^{-}$, occurs at a lower value of $\alpha$ than when the drive is $+A$, with current $J^{+}$. Therefore, when $\alpha$ is gradually increased beyond the value at the HB point for negative drive $-A$, the cycle averaged rectification current $\langle J \rangle = (1/2)(J^{+}+J^{-})$, is negative and increases in absolute value, as shown in Fig.\,\ref{F5}(b). As $\alpha$ is further increased to the value at the HB point for positive drive $+A$, the magnitude of $\langle J \rangle$ reaches a local maximum as a function of $\alpha$. Increasing $\alpha$ even further results in a decrease in the magnitude of $\langle J \rangle$. This occurs because the particles are now transported as a condensed traveling mode in both directions.
A qualitatively similar behavior of $\langle J \rangle$ is found for a range of different values of the particle size $h$. However, on increasing $\alpha$ even further, so that it is well above the value at the HB points, the dependence of $\langle J \rangle$ on $\alpha$ becomes very sensitive to the value of $h$. For instance, in Fig.\,\ref{F5}(b) the rectification current attains a second minimum at around $\alpha=6$ for $h=0.2$, whereas for $h=0.16$ the second minimum disappears and $\langle J \rangle$ increases monotonically as a function of $\alpha$.
To confirm the validity of the (mean field) DDFT results, we performed Brownian dynamics computer simulations -- i.e.\ we numerically integrated the Langevin equations of motion (\ref{langevin}), in order to compare with our DDFT results. In order to make the simulations more convenient to implement, we replace the hard core potential, $w_{\rm hr}$, by an equivalent, more tractable soft core potential, $w_{\rm s}(x_{ij}) = \epsilon (h_*/x_{ij})^{19}$, where the constants $\epsilon$ and $h_*$ can be tuned to reproduce the desired effective hard-core length of the potential. For fixed $\epsilon$ and $h_*$ the effective hard core length $h_{\rm eff}$ of the particles becomes a function of $\alpha$, $T$ and, in general, also of the number of particles $N$ \cite{barker76}. In our simulations we set $\epsilon=0.01$ and $h_*=0.2$ which corresponds to an effective hard-core $h_{\rm eff}\approx 0.16$, for $\alpha=10$. Our numerical data suggests that the dependence of $h_{\rm eff}$ on $T$ and $N$ is rather weak, and can therefore be neglected.
In Fig.\,\ref{F5}(b) we compare the DDFT predictions for $h=0.16$ with the corresponding simulation results for $h_{\rm eff}\approx 0.16$. The first minimum in the current $\langle J \rangle$ as a function of $\alpha$ is clearly confirmed by the Brownian dynamics simulation results for $N=5$ particles. The simulation results displayed in Figs.\,\ref{F5}(c) and (d) also show that the overall structure of $\langle J \rangle$ as a function of $\alpha$ does not change much as $N$ is increased up to $10$. Moreover, as the temperature is decreased from $T=0.5$ down to $T=0.2$, the maximum in the magnitude of the current curve, $|\langle J \rangle|$, becomes even more pronounced, with the magnitude of the peak rectification current increasing by one order of magnitude. This effect, which is well established in the ratchet literature \cite{RMP09}, underlines the key role of noise in activating transport (in either direction) when the amplitude of the drive is smaller than both the depinning thresholds, $A_L$ and $A_R$, of the ratchet potential, $U(x)$.
\subsection{Strong attraction limit}
\label{strongatt}
\begin{figure}[t]
\centering
\includegraphics[width=0.48\textwidth]{newFig7.eps}
\caption{(Color online) (a) The unidirectional currents $J^{\pm}$ as functions of the number of particles $N$, in a system with total length $S=10L$. The heavy solid and dashed curves correspond to the currents in the limit of very strong attraction ($\alpha\to\infty$) for $h=0.1$ and $h=0.2$, respectively. The thin dashed curves represent the currents for non-attracting particles ($\alpha=0$) with $h=0.1$. In the inset we display the $T=0$ limiting values for the critical amplitude $A$ required for a current to flow, as a function of $hN$. In (b) we display the rectification currents $\langle J\rangle$ obtained from the currents displayed in (a). \label{F6}}
\end{figure}
In order to study the properties of the system when the attraction between the particles dominates over the thermal motion of the particles and the pinning by the external potential, we consider the limit $\alpha \to \infty$. This allows us to reduce the system of equations (\ref{langevin}) to a single equation of motion for the center of mass of the particle condensate, $y=(1/N)\sum_{i=1}^{N} x_i$. As noted above in Sec.\ \ref{finitelength}, the total force exerted by the channel potential on the condensate is $f=-\sum_{i=1}^{N}dU(x_i)/dx_i$. If we assume that the pair attraction is so strong that the rods are closely packed together in a single condensate with their ends touching, the total force $f$ can be rewritten as $f=-\sum_{i=1}^{N}dU(x+(i-1)h)/dx$, where $x$ denotes the coordinate of the center of the first particle in the file. Now, if we assume that $h$ is small compared to the period $L$ of the channel potential, then the sum can be replaced by an integral:
\begin{eqnarray}
f \approx -\frac{1}{hN}\int_{x}^{x+hN}\frac{dU(y)}{dy}dy \notag \\
= -\frac{U(x+hN)-U(x)}{hN},
\end{eqnarray}
leading to the following effective equation of motion for the center of mass:
\begin{eqnarray}
\label{langevin2}
\frac{d x}{d t} = -\frac{U(x+hN)-U(x)}{hN} +A(t)+ \sqrt{\frac{2T}{N}}\xi(t).
\end{eqnarray}
Here, $\xi(t)$ has the same statistics as $\xi_i(t)$ in Eq.\,(\ref{langevin}). To derive Eq.\,(\ref{langevin2}), we use the fact that the sum of $N$ independent sources of Gaussian white noise with variance $1$ is also a Gaussian noise, but with variance $1/N$.
Equation (\ref{langevin2}) corresponds to the equation of motion for a single Brownian particle diffusing in the effective external potential $V_{\rm eff}(x) = \frac{1}{hN} \int [U(x+hN)-U(x)]\,dx$, in contact with a thermal bath with temperature $T/N$. The first observation from Eq.\ \eqref{langevin2} is that for large condensates, diffusion becomes negligible, so that $J^{\pm}$ become sizable only if the drive amplitude, $A$, overcomes the pinning force induced by the effective potential $V_{\rm eff}(x)$. For $T\equiv 0$, this critical amplitude $A$ is plotted in the inset of Fig.\,\ref{F6}(a), as a function of the size of the condensate $hN$. Within the shaded area, the condensate is pinned by the effective external potential $V_{\rm eff}$; depinning occurs either to the left or to the right, depending on the drive orientation. Note that for $hN \to 0$, the right and left critical amplitudes coincide with the single particle depinning thresholds, $A_{L}$ and $A_R$, introduced in Sec. \ref{DC}. Similarly to the case for pointlike particles \cite{sav04}, selecting an appropriate combinations of $h$ and $N$, one can achieve the complete locking of the condensed mode in one direction, but not in the other \cite{PA10}, which yields the upper bound $|A|/2$ for the modulus of $\langle J \rangle$.
Finally, using Eq.\,(\ref{langevin2}), we compare in Fig.\ref{F6} the efficiency of the low frequency transport of strongly attracting ($\alpha \to \infty$) and non-interacting ($\alpha =0$) particles. We fix the size of the particles $h$ and change the average density $\bar \rho$ by changing the number of particles $N$ in the system. The unidirectional currents for the condensate oscillate with $N$ and hit the respective upper (lower) bound, $J^{\pm}=\pm |A|$, for $hN$ equal to a multiple of $L$. In the absence of particle attraction, $|J^{\pm}|$ increases monotonically with $N$ and attain the same upper bound only for $hN=S$. The corresponding rectification currents are shown in Fig.\ref{F6}(b). For certain combinations of $h$ and $N$, the magnitude of the current of the condensate is several orders of magnitude larger than for non-attracting particles.
\section{Concluding remarks}
\label{conc}
In this paper we have developed a DDFT for studying the dynamics of a file of attracting colloidal particles confined within a channel that exerts a periodic ratchet potential on the colloids. We find that the attraction between the colloids leads to rather rich behavior in the DDFT model when the particles are driven, including transitions from stationary to time-periodic density profiles as the strength of the attraction between the particles is increased. We also find that for strong enough attraction, there can be coexistence of stable stationary density profiles with different spatial periods and time-periodic density profiles, each with different values for the particle current $J$.
These dynamical transitions in our model stem from the fact that the approximate free energy functional \eqref{dft_energy} on which our DDFT is based, predicts that the system exhibits gas-liquid phase separation for sufficiently large values of the ratio $\alpha/Th\lambda$. This prediction comes as a consequence of the mean-field approximation made in constructing the free energy. In reality, for a system containing a finite number of particles, there is no true phase transition. Furthermore, since the system is one-dimensional, there is no phase transition even in the infinite sized system (i.e\ in the thermodynamic limit when $N,S\to\infty$, with average density $\bar{\rho}=N/S$ remaining constant). In 1D systems such as that studied here, as the attraction strength $\alpha$ is increased, the particles increasingly tend to gather together, but no true phase transition can be defined. Thus, in reality, as can be inferred from our Brownian dynamics simulation results, there are no `sharp' transitions from the pinned to the depinned (time-periodic) state, as $\alpha$ is increased. Thus, we expect that fluctuations will round the predicted transitions. Nonetheless, as the comparison with the Brownian dynamics simulations show, the results from our DDFT do capture the main features of the system - i.e.\ that for lower values of the attraction strength $\alpha$, the particles are uniformly distributed and that at higher values of $\alpha$ the particles gather to form a cluster, and that if the system length $S$ is sufficiently long, this clustering leads to time-periodic currents $J$ when the system is driven.
In our discussions above we have pointed out that similarities exist between the DDFT equation \eqref{eq3} for the particle density employed here and thin film equations that are used to model the dynamics of films and drops of partially wetting liquids on heterogenous solid substrates with and without additional driving forces \cite{KaTh07}.
The similarities result from the fact that in both cases kinetic equations for conserved fields are used, and that the respective free energy functionals contain terms which result in similar physical effects. For instance, the role of the particle-particle interactions in the present work is taken by wettability effects in the context of droplet dynamics. The parallels between the two systems have allowed us to use the knowledge gained from studying one system to understand aspects of the other. In particular, in the present work we have drawn on the understanding of depinning mechanisms developed for thin films in Refs.~\cite{dep2,BKHT11}. Furthermore, Ref.~\cite{BKHT11} indicates that one might encounter rich nonlinear behaviour when considering the behaviour of attractive hard-core particles in wider corrugated channels, i.e., without the `restriction' of single file motion. Note, however, that there are clear limits to the similarities: The thin film models referred to above do not account for any effect that is equivalent to the freezing
instability discussed above. We believe that studying in detail the similarities and differences between DDFT and thin film models is worthwhile, as it will allow for much cross fertilisation of ideas and techniques between the two fields.
One of the most striking features of our system is that the current $J$ depends very sensitively on the size of the particles $h$ and on the total number of particles in the system $N$, particularly when the particles are strongly attracted to one another so that they are bound together to form a cluster that moves as a unit through the system, when there is an external drive on the system. In fact, the direction of travel can be completely reversed when the system is driven by an AC potential, simply by changing the number of particles in the file by one -- i.e.\ adding an extra particle to a file can cause it to reverse its direction of motion without changing the external drive. This means that one can use the present system to form a molecular shuttle that moves back and forth between two docking stations, loading and unloading single particles from a source to a sink docking station \cite{PA10}. As the process repeats, a steady flux is established along the channel. This mechanism can be highly efficient if the system parameters are carefully tuned.
The present model thus provides a useful system for developing a deep understanding of the behavior of driven macromolecular and colloidal systems occurring in nanoscience and biology. In particular, by using DDFT, which is based on a fully microscopic expression for the Helmholtz free energy functional \eqref{dft_energy}, we are able to build into our theory a reliable description of the correlations between the particles, and their influence on the dynamics of the system as a whole.
\section*{Acknowledgements}
This work was partly supported by the HPC-Europa2 Transnational Access Programme, proposal No. 278. AJA gratefully acknowledges support from RCUK. FM acknowledges partial support from the Seventh Framework Programme under grant agreement No. 256959, project NANOPOWER.
|
\section{#1}\setcounter{equation}{0}}
\renewcommand{\theequation}{\thesection.\arabic{equation}}
\def{\hbox{ 1\kern-.8mm l}}{{\hbox{ 1\kern-.8mm l}}}
\def{\hbox{ 0\kern-1.5mm 0}}{{\hbox{ 0\kern-1.5mm 0}}}
\def\,{\rm Tr}\, {\,{\rm Tr}\, }
\def\,{\rm Tr}_M^{q,y}\, {\,{\rm Tr}_M^{q,y}\, }
\def\tau{\tau}
\def\eta{\,{\rm e}}
\def\hat{n}{\hat{n}}
\def\tilde{g}{\tilde{g}}
\def{\mathcal J}{{\mathcal J}}
\def{\mathcal K}{{\mathcal K}}
\def{\mathcal N}{{\mathcal N}}
\def{\mathcal O}{{\mathcal O}}
\def{\mathcal U}{{\mathcal U}}
\def\tilde{\mathcal O}{\tilde{\mathcal O}}
\def{\mathbb Z}{{\mathbb Z}}
\def{\rm Tr}{{\rm Tr}}
\def\slashed{\slashed}
\def\left{\left}
\def\right{\right}
\def\iota{\iota}
\def\phi{\phi}
\def\Delta{\Delta}
\def\sigma{\sigma}
\def\gamma{\gamma}
\def\Gamma{\Gamma}
\def\mu{\mu}
\def\bar{\mu}{\bar{\mu}}
\def\bar{w}{\bar{w}}
\def\eta{\eta}
\def\alpha{\alpha}
\def\dot{\alpha}{\dot{\alpha}}
\def\bar{a}{\bar{a}}
\def\tilde{\eta}{\tilde{\eta}}
\def\bar{Q}{\bar{Q}}
\def\o[#1]{{\rm O}\left({#1}\right)}
\def\dotl[#1,#2]{\left\langle #1, #2 \right\rangle}
\def\dotlb[#1,#2]{[ #1, #2 ]}
\def\dotp[#1,#2]{(#1) \cdot (#2)}
\def\>{\rangle}
\def\<{\langle}
\def\tilde{l}{\tilde{l}}
\def\tilde{\mu}{\tilde{\mu}}
\title{The Heat Kernel on $AdS$}
\author{Rajesh Gopakumar$^{a}$\footnote{gopakumr AT hri DOT res DOT in}, Rajesh Kumar Gupta$^{a,b}$\footnote{R.K.Gupta AT uu DOT nl} and Shailesh Lal$^{a}$\footnote{shailesh AT hri DOT res DOT in} \\
$^a$Harish-Chandra Research Institute, \\
$\;$Chhatnag Road,\\
$\;$Jhusi, India 211019\\
$^b$Institute of Theoretical Physics,\\
$\;$Utrecht University, Leuvenlaan 4, 3584 CE,\\
$\;$Utrecht, The Netherlands\\}
\abstract{We explicitly evaluate the heat kernel for the Laplacian of arbitrary spin tensor fields on the thermal quotient of (Euclidean) $AdS_N$ for $N\geq 3$ using the group theoretic techniques employed for $AdS_3$ in arXiv:0911.5085. Our approach is general and can be used, in principle, for other quotients as well as other symmetric spaces.}
\preprint{HRI/ST/1104}
\begin{document}
\section{Introduction}
Quantum effects in $AdS$ spacetimes are important in the AdS/CFT correspondence in trying to go beyond the conventional tree level gravity description. The leading effects come at one loop and capture, for instance, important information about the spectrum of the theory. In the language of the boundary gauge theory these come from diagrams which are suppressed by ${1\over N^2}$ compared to the planar ones. For example, this genus one contribution to the free energy of the QFT captures all the quadratic fluctuations about the bulk $AdS$ background. From a worldsheet perspective, for the closed string theory in the bulk, this is the torus contribution to the free energy.
The string theory sigma model with an $AdS$ target space is as yet fairly intractable at the quantum level. Thus some important clues about the worldsheet structure might be gained from knowing the spectrum of the gauge theory and trying to reproduce the one loop answer coming from the quadratic fluctuations of the dual fields. Typically the quadratic terms involve Laplacians acting on arbitrary tensor fields. Evaluation of the path integral at quadratic level then requires one to compute the determinants of such general laplacians about an $AdS$ background (more precisely a thermal quotient to define a generating function). The best way to evaluate these is the heat kernel method. This, in addition, has the virtue that the proper time that enters here has an interpretation in terms of the modulus of the genus one worldsheet \cite{Polchinski:1985zf}. Essentially the heat kernel captures the first quantised description of the particles which go into constituting the string spe
ctrum.
Recall that given the normalised eigenfunctions $\psi^{(S)}_{n,a}\left(x\right)$ of the Laplacian $\Delta_{\left(S\right)}$ for a spin-$S$ field on a manifold ${\mathcal M}_{d+1}$, and the spectrum of eigenvalues $E_n^{(S)}$, we can define the heat kernel between two points $x$ and $y$ as
\begin{equation}
\label{genkernel}
K^{(S)}_{ab}\left(x,y;t\right) = \langle y,b | e^{t\Delta_{\left(S\right)}} | x,a \rangle = \sum_n \psi^{(S)}_{n,a}\left(x\right)\psi^{(S)}_{n,b}\left(y\right)^*e^{tE_n^{(S)}},
\end{equation}
where $a$ and $b$ are local Lorentz indices for the field. We can trace over the spin and spacetime labels to define the traced heat kernel as
\begin{equation}
\label{intkernel}
K^{(S)}\left(t\right)\equiv {\rm Tr} e^{t\Delta_{\left(S\right)}} =\int_{\mathcal M}\sqrt{g} d^{d+1}x\sum_a K^{(S)}_{aa}\left(x,x;t\right).
\end{equation}
The one-loop partition function is related to the trace of the heat kernel through
\begin{equation}
ln\,Z^{(S)}=\ln\,det\left(-\Delta_{\left(S\right)}\right)= {\rm Tr} \ln\left(-\Delta_{\left(S\right)}\right) =-\int_{0}^\infty\frac{dt}{t}{\rm Tr} e^{t\Delta_{\left(S\right)}}.
\end{equation}
For a general manifold, the computation of the heat kernel is a formidable task, even for the scalar Laplacian. Typically one has asymptotic results. However, for symmetric spaces such as spheres and hyperbolic spaces (Euclidean $AdS$) there are many simplifications. This is because these spaces can be realised as cosets $G/H$ and one can therefore use the powerful methods of harmonic analysis on group manifolds. In \cite{David:2009xg} these techniques were used to explicitly
compute the heat kernel on thermal $AdS_3$ for fields of arbitrary spin. Though \cite{David:2009xg} exploited the group theory, it also used at several places the particular fact that $S^3$ (from which one continued to $AdS_3$) itself is the group manifold $SU(2)$. In fact, many properties of $SU(2)$ were used in intermediate steps and it was not completely obvious how these generalise to higher dimensional spheres or $AdS$ spacetimes.
In this paper, we will generalise the methods of \cite{David:2009xg} to compute the heat kernel for the Laplacian for arbitrary spin tensor fields on
thermal $AdS$ spacetimes. This therefore includes the cases of $AdS_4$, $AdS_5$ and $AdS_7$ which play a central role in the AdS/CFT correspondence.
Since we are primarily interested in evaluating the one-loop partition function for a spin-$S$ particle, we shall concentrate on the traced heat kernel though we will see that the techniques are sufficiently general. One may, for instance, also adapt these techniques to evaluate other objects of interest such as the (bulk to bulk) propagator.
We shall
mostly focus on ${\cal M}=AdS_{2n+1}/\Gamma$
where $\Gamma$ is the thermal quotient.
For practical reasons we will obtain the answer for thermal $AdS$ by analytic continuation of the answer for an appropriate ``thermal quotient" of
the $N$-sphere. As explained in \cite{David:2009xg} there are good reasons to believe that this analytic continuation of the harmonic analysis works for odd dimensional spheres and hyperboloids (Euclidean AdS). For even dimensional hyperboloids there are some additional discrete representations other than the continuous ones one obtains by a straightforward analytic continuation. However, this additional set of representations does not contribute for Laplacians over a wide class of tensor fields, which in particular include the symmetric transverse traceless (STT) tensors. We provide necessary details in Appendix \ref{evendim}.
We expect that the results here will have many applications. In the three dimensional case the corresponding results have been used to clarify the quantum nature of topologically massive gravity in $AdS_3$ \cite{Gaberdiel:2010xv}. They were also used to show that the higher spin gauge theories (with spins $s=2, 3 \ldots N$) realise ${\cal W}_N$ symmetry at the quantum level \cite{Gaberdiel:2010ar} - generalising the classical Brown-Henneaux like result for such theories \cite{Campoleoni:2010zq,Henneaux:2010xg}. This was an important input in formulating a duality between these higher spin theories (with additional scalar fields) and ${\cal W}_N$ minimal models \cite{Gaberdiel:2010pz}.
In the higher dimensional cases of interest here one can apply our results to study the Vasiliev higher spin theories. Such theories are conceivably related to a subsector of free Yang-Mills theories and perhaps the higher spin theories are higgsed in an interesting way in going away from the free theory.
Evaluating the one loop fluctuations in the bulk can help us in checking these conjectures with more precision.
The plan of the paper is as follows. In the next section we briefly review the harmonic analysis on homogeneous spaces that is the mainstay of these computations and illustrate it with the
case of $S^{2n+1}$. In Sec. 3, we describe how to generally consider quotients of these symmetric spaces. Sec. 4 describes the analytic continuation of $S^{2n+1}$ to the Euclidean hyperboloids. We evaluate the coincident heat kernels for the general class of symmetric tensor representations and check against existing results in the literature. In Sec. 5 we obtain the answer for the traced heat kernel on thermal $AdS_{2n+1}$. Sec. 6 contains an application to the scalar one loop partition function. Finally Sec. 7 has some concluding remarks.
\section{The Heat Kernel on Homogeneous Spaces}
\label{HarmonicAnalysis}
The heat kernel of the spin-$S$ Laplacian\footnote{By spin-$S$ we refer to the representation under which the field transforms under tangent space rotations. In the case of $S^{2n+1}$ or (euclidean) $AdS_{2n+1}$, this will be a representation of $SO(2n+1)$. The Laplacian is that of a tensor field transforming in this representation. In the case of spheres and hyberboloids we will consider the Laplacian with the Christoffel connection.}
may be evaluated over the spacetime manifold $\mathcal M$ by solving the appropriate heat equation. Alternatively, one may attempt a direct evaluation by constructing the eigenvalues and eigenfunctions of the spin-$S$ Laplacian and carrying out the sum over $n$ that appears in (\ref{genkernel}). Both these methods quickly become forbidding when applied to an arbitrary spin-$S$ field. However, if $\mathcal M$ is a homogeneous space $G/H$, then the use of group-theoretic techniques greatly simplifies the evaluation. The main simplifications arise from the following facts which we will review below and then heavily utilise:
\begin{enumerate}
\item
The eigenvalues $E_n^{(S)}$ of the Laplacian $\Delta_{(S)}$ are determined in terms of the quadratic Casimirs of the symmetry group $G$ and the isotropy subgroup $H$. There is thus a large degeneracy of eigenvalues.
\item
The eigenfunctions $\psi_{n,a}^{(S)}(x)$ are matrix elements of unitary representation matrices of $G$.
\item
This enables one to carry out the sum over degenerate eigenstates using the group multiplication properties of the matrix elements. Thus a large part of the sum in (\ref{genkernel}) can be explicitly carried out.
\end{enumerate}
We begin with a brief recollection of some basic facts about harmonic analysis on coset spaces. This will collate the necessary tools with which we evaluate (\ref{genkernel}) and (\ref{intkernel}), and will set up our notation. The interested reader is referred to \cite{Salam:1981xd,Camporesi:1990wm,Camporesi:1995fb} for introduction and details and to \cite{David:2009xg} for explicit examples of these constructions. Given compact Lie groups $G$ and $H$, where $H$ is a subgroup of $G$, the coset space $G/H$ is constructed through the right action of $H$ on elements of $G$
\begin{equation}
G/H=\lbrace gH \rbrace.
\end{equation}
(We will also need to consider left cosets, $\Gamma\backslash G$, where $\Gamma$ will act on elements of $G$ from the left.)
We recall that $G$ is the principal bundle over $G/H$ with fibre isomorphic to $H$. Let $\pi$ be the projection map from $G$ to $G/H$, \textit{i.e.}
\begin{equation}
\pi\left(g\right)=gH\quad\forall g\in G.
\end{equation}
Then a section $\sigma\left(x\right)$ in the principal bundle is a map
\begin{equation}
\sigma: G/H \mapsto G,\quad \hbox{such that } \pi\circ\sigma=e,
\end{equation}
where $e$ is the identity element in $G$, and $x$ are coordinates in $G/H$.
A class of sections which will be useful later is of the form
\begin{equation}
\label{thermalex}
\sigma\left(gH\right)=g_{\circ},
\end{equation}
where $g_{\circ}$ is an element of the coset $gH$, which is chosen by some well-defined prescription. The so called `thermal section' that we choose in Section \ref{thermalquotient} is precisely of this form.
Let us label representations of $G$ by $R$ and representations of $H$ by $S$. We will sometimes refer to representations $S$ of $H$ as spin-$S$ representations.\footnote{In the case of sphere and hyperboloids, $H$ is isomorphic to the group of tangent space rotations for the manifold $G/H$. See footnote 1.} The vector space which carries the representation $R$ is called $\mathcal V_R$, and has dimension $d_R$, while the corresponding vector space for the representation $S$ is $\mathcal V_S$ of dimension $d_S$.
Eigenfunctions of the spin-$S$ Laplacian are then given by the matrix elements
\begin{equation}
\label{eigenf}
{\psi^{\left(S\right)I}_a}\left(x\right)={{{\mathcal U}^{\left(R\right)}}\left(\sigma\left(x\right)^{-1}\right)_a}^I,
\end{equation}
where $S$ is the unitary irreducible representation of $H$ under which our field transforms, and $R$ is any representation of $G$ that contains $S$ when restricted to $H$. $a$ is an index in the subspace $\mathcal V_S$ of $\mathcal V_R$, while $I$ is an index in the full vector space $\mathcal V_R$. Generally, a given representation $S$ can appear more than once in $R$. However, we shall be interested in the coset spaces $SO\left(N+1\right)/SO\left(N\right)$ and $SO\left(N,1\right)/SO\left(N\right)$, for which a representation $S$ appears at most once \cite{Barut:1986dd,Camporesi}. We have therefore dropped a degeneracy factor associated with the index $a$, which appears in the more complete formulae given in \cite{Camporesi}.
The corresponding eigenvalues are given by
\begin{equation}
\label{eigenval}
-E_{R,I}^{(S)}= C_2\left(R\right)- C_2\left(S\right).
\end{equation}
The index $n$ for the eigenvalues of the spin-$S$ Laplacian that appeared in (\ref{genkernel}) is therefore a pair of labels, viz. $\left(R,I\right)$,\footnote{Note that $a$ labels the components of the eigenfunction and is not a part of the index $n$.} where the eigenfunctions that have the same label $R$ but a different $I$ are necessarily degenerate.
We will therefore drop the subscript $I$ for $E^{(S)}$.
The expression (\ref{genkernel}) for the heat kernel then reduces to
\begin{equation}
\label{kernelintermsofeigenfunctions}
K^{(S)}_{ab}\left(x,y;t\right) = \sum_{R,I}a_R^{(S)}\psi^{(S)}_{(R,I),a}\left(x\right)\psi^{(S)}_{(R,I),b}\left(y\right)^*e^{tE_R^{(S)}},
\end{equation}
where $ a_R^{(S)}=\frac{d_R}{d_S}\frac{1}{V_{G/H}}$ is a normalisation constant (see Appendix \ref{normalisationkernel}).
This can be further simplified by putting in the expression (\ref{eigenf}) for the eigenfunctions.
\begin{eqnarray}
\label{untracedkernel}
K^{(S)}_{ab}\left(x,y;t\right) &=& \sum_{R}\sum_{I=1}^{d_R} a_R^{(S)}{{{\mathcal U}^{\left(R\right)}}\left(\sigma\left(x\right)^{-1}\right)_a}^I \left[{{{\mathcal U}^{\left(R\right)}}\left(\sigma\left(y\right)^{-1}\right)_b}^I\right]^*e^{tE_R^{(S)}} \nonumber \\
&=& \sum_R a_R^{(S)} {{{\mathcal U}^{\left(R\right)}}\left(\sigma\left(x\right)^{-1}\sigma\left(y\right)\right)_a}^b e^{tE_R^{(S)}},
\end{eqnarray}
where we have used the fact that the $\mathcal U^{\left(R\right)}$ furnish a unitary representation of $G$. As an aside, we note that this matrix representation of the group composition law is the generalisation of the addition theorem for spherical harmonics on $S^2$ to arbitrary homogeneous vector bundles on coset spaces.
To establish notation for later use, we define the heat kernel with traced spin indices
\begin{equation}
\label{kernel}
K^{(S)}\left(x,y;t\right)\equiv \sum_{a=1}^{d_S} K^{(S)}_{aa}\left(x,y;t\right)
= \sum_R a_R^{(S)} {\rm Tr}_S\left({{\mathcal U}^{\left(R\right)}}\left(\sigma\left(x\right)^{-1}\sigma\left(y\right)\right)\right) e^{tE_R^{(S)}},
\end{equation}
where the symbol
\begin{equation}
{\rm Tr}_S\left({\mathcal U}\right)\equiv\sum_{a=1}^{d_S}\<a,S\vert{\mathcal U}\vert a,S\>,
\end{equation}
and can be thought of as a trace over the subspace $\mathcal V_S$ of $\mathcal V_R$. Note that this restricted trace is invariant under a unitary change of basis of $\mathcal V_S$ and
{\it not} invariant under the most general unitary change of basis in $\mathcal V_R$.
\subsection{The Heat Kernel on $S^{2n+1}$}
As a prelude to evaluating the traced heat kernel on the ``thermal quotient" of the odd-dimensional sphere, let us evaluate (\ref{untracedkernel}) for the case without any quotient.
That is, we focus first on $S^{2n+1}\simeq SO(2n+2)/SO(2n+1)$. We will describe the eigenfunctions (\ref{eigenf}) and define the sum over $R$ explicitly. This will be useful when we analytically continue our results to the corresponding hyberbolic space. We begin by recalling some facts from the representation theory of special orthogonal groups.
Unitary irreducible representations of $SO(2n+2)$ are characterised by a highest weight, which can be expressed in the orthogonal basis as the array
\begin{equation}
\label{so2nrep}
R=\left(m_1,m_2,\ldots ,m_n,m_{n+1}\right),\quad m_1\geq m_2\geq \ldots \geq m_n \geq \vert m_{n+1}\vert\geq 0
\end{equation}
where the $m_1\ldots m_{n+1}$ are all (half-)integers.
Similarly, unitary irreducible representations of $SO\left(2n+1\right)$ are characterised by the array
\begin{equation}
S=\left(s_1,s_2,\ldots ,s_n\right),\quad s_1\geq s_2\geq \ldots \geq s_n \geq 0,
\end{equation} where the $s_1\ldots s_{n}$ are all (half-)integers.
Then the quadratic Casimirs for the unitary irreducible representations for an orthogonal group of
rank $n+1$ can be expressed as (see e.g.\cite{Barut:1986dd,Fulton}).
\begin{equation}
\label{Casimir}
C_2\left(m_1,\ldots,m_{n+1} \right)=m^2+2r\cdot m .
\end{equation}
Here the dot product is the usual euclidean one, and the Weyl vector $r$ is given by
\begin{equation}
r_i= \left\{\begin{array}{l} n-i+1\quad \rm{if}~ G=SO\left(2n+2 \right), \\ \left(n+\frac{1}{2}\right)-i\quad \rm{if}~ G=SO\left(2n+1\right),\\ \end{array}\right\}
\end{equation}
where $i$ runs from $1$ to $n+1$.
Let us now consider the expression (\ref{untracedkernel}) for the spin-$S$ Laplacian on $S^{2n+1}$. The eigenvalues $E_R^{\left(S\right)}$ are given by (\ref{eigenval}) and (\ref{Casimir}) and may be written down compactly as
\begin{equation}
\label{eigenvals2n1}
-E_R^{\left(S\right)}=m^2+2r_{SO\left(2n+2\right)}\cdot m-s^2-2r_{SO\left(2n+1\right)}\cdot s.
\end{equation}
The corresponding eigenfunctions are given by (\ref{eigenf}) where we have to specify which are the representations $R$ of $SO\left(2n+2\right)$ that contain a given representation $S$ of $SO\left(2n+1\right)$. This is determined by the branching rules, which for our case state that a representation $R$ given by (\ref{so2nrep}) contains the representation $S$ if
\begin{equation}
\label{snbranching}
\quad m_1\geq s_1 \geq m_2\geq s_2\geq \ldots \geq m_n\geq s_n\geq\vert m_{n+1}\vert,\quad \left(m_i-s_i\right)\in{\mathbb Z}.
\end{equation}
Using these branching rules, one can show that the expression that appears on the right of (\ref{eigenvals2n1}) is indeed positive definite, so that the eigenvalue itself is negative definite as per our conventions.
These rules further simplify if we restrict ourselves to symmetric transverse traceless (STT) representations of $H$. These tensors of rank $s$
correspond to the highest weight $\left(s,0,\ldots,0\right)$. In this case, some of these inequalities get saturated, and one obtains the branching rule
\begin{equation}
\label{snbranchingSTT}
\quad m_1\geq s= m_2\geq 0,
\end{equation}
with all other $m_i, s_i$ zero, \footnote{For the special case of $n=1$, i.e. $S^3$, the branching rule is $m_1\geq s= \vert m_2\vert\geq 0$.} and the equality follows from requiring that $R$ contain $S$ in the maximal possible way. Essentially, this is equivalent to the transversality condition.
The sum over $R$ that appears in (\ref{untracedkernel}) is now a sum over the admissible values of $m_1$ in the above inequality. Thus if we restrict ourselves to evaluating the heat kernel for STT tensors, then we will be left with a sum over $m_1$ only.
The expression (\ref{eigenvals2n1}) for the eigenvalue also simplifies in this case. With the benefit of hindsight, we will write this in a form that is suitable for analytic continuation to $AdS$
\begin{equation}
\label{eigenvals2n1stt}
-E_R^{\left(S\right)}=\left(m_1+n\right)^2- s -n^2.
\end{equation}
Using these tools, one can write down a formal expression for the heat kernel for a spin-$S$ particle between an arbitrary pair of points $x$ and $y$ on $S^{2n+1}$ using (\ref{untracedkernel}). This is given by
\begin{equation}
\label{kernelsphere}
K^S_{ab}\left(x,y;t\right) = \sum_{m_i} \frac{n!}{2\pi^{n+1}}\frac{d_R}{d_S} {{{\mathcal U}^{\left(R\right)}}\left(\sigma\left(x\right)^{-1}\sigma\left(y\right)\right)_a}^b e^{tE_R^{(S)}},
\end{equation}
where we have simply expanded out the sum over $R$ into a sum over the permissible values of $m_i$ determined by (\ref{snbranching}), and inserted the expression for the volume of the $\left(2n+1\right)$-sphere. Expressions for the dimensions $d_R$ and $d_S$ are well known (see for example\cite{Barut:1986dd,Fulton}). We list them here for the reader's convenience.
\begin{equation}
\label{dimensions}
d_R=\prod_{i<j=1}^{n+1}\frac{l_i^2-l_j^2}{\mu_i^2-\mu_j^2},\quad d_S= \prod_{i<j=1}^{n}\frac{\tilde{l}_i^2-\tilde{l}_j^2}{\tilde{\mu}_i^2-\tilde{\mu}_j^2}\prod_{i=1}^n \frac{\tilde{l}_i}{\tilde{\mu}_i},
\end{equation}
where $l_i=m_i+\left(n+1\right)-i$, $\mu_i=\left(n+1\right)-i$, $\tilde{l}_i=s_i+n-i+\frac{1}{2}$, and $\tilde{\mu}_i=n-i+\frac{1}{2}$. As explained above, this expression further simplifies for the STT tensors, and we obtain
\begin{equation}
\label{kernelsttsphere}
K^{S}_{ab}\left(x,y;t\right) = \sum_{m_1} \frac{n!}{2\pi^{n+1}} \frac{d_{\left(m_1,s\right)}}{d_s} {{{\mathcal U}^{\left(m_1,s\right)}}\left(\sigma\left(x\right)^{-1}\sigma\left(y\right)\right)_a}^b e^{tE_R^{(S)}},
\end{equation}
where the labels $\left(m_1,s\right)$ and $s$ that appear on the RHS are shorthand for $R=\left(m_1,s,0\ldots,0\right)$ and $S=\left(s,0\ldots,0\right)$ respectively. This expression should be compared with the equation $\left(3.9\right)$ obtained in \cite{David:2009xg} for the case of $3$ dimensions. The traced heat kernel for the STT tensors is then given by
\begin{equation}
\label{kerneltracesttsphere}
K^{S}\left(x,y;t\right) = \sum_{m_1} \frac{n!}{2\pi^{n+1}} \frac{d_{\left(m_1,s\right)}}{d_s} {\rm Tr}_S\left({\mathcal U}^{\left(m_1,s\right)}\left(\sigma\left(x\right)^{-1}\sigma\left(y\right)\right)\right) e^{tE_R^{(S)}}.
\end{equation}
As mentioned earlier, we can in principle use these formulae to construct explicit expressions for the heat kernel between two points $\grave{a}$ \textit{la} \cite{David:2009xg}, and thus for the bulk to bulk propagator. This would, however also require using explicit matrix elements of $SO\left(2n+2\right)$ representations, and we shall not pursue this direction further here.
\section{The Heat Kernel on Quotients of Symmetric Spaces}
We consider the heat kernel on the quotient spaces $\Gamma\backslash G/H$ where $\Gamma$ is a discrete group which can be embedded in $G$. Though it is not essential to our analysis, we will assume that $\Gamma$ is of finite order and is generated by a single element. This is indeed true for the quotients (\ref{quotient}) we consider here. In particular, for the ``thermal quotient" on the $N$-sphere, $\Gamma$ is isomorphic to ${\mathbb Z}_N$.
To evaluate the heat kernel (\ref{genkernel}) on this space, a choice of section that is \textit{compatible} with the quotienting by $\Gamma$ is useful. By this we mean that if
$\gamma\in\Gamma$ acts on points $x=gH \in G/H$ by $\gamma:gH \mapsto\gamma\cdot gH$, then a section $\sigma\left(x\right)$ is said to be compatible with the quotienting $\Gamma$ iff
\begin{equation}
\label{compatible}
\sigma\left(\gamma\left(x\right)\right) = \gamma\cdot\sigma\left( x\right).
\end{equation}
The utility of this choice of section will become clear when we explicitly evaluate the traced heat kernel (\ref{intkernel}) for such geometries.
\subsection{The Thermal Quotient of $S^{5}$}
\label{thermalquotient}
As an example, we consider the thermal quotient of $S^5$. This will serve as a useful prototype to keep in mind. We shall also see that we can extrapolate the analysis to the general odd-dimensional sphere. To begin with, let us express the thermal quotient in terms of ``triple-polar" coordinates on $S^5$, which are complex numbers $\left(z_1,z_2,z_3\right)$ such that
\begin{equation}
|z_1|^2+|z_2|^2+|z_3|^2=1.
\end{equation}
We consider the quotient
\begin{equation}
\label{quotient}
\gamma: \{\phi_i \} \mapsto \{ \phi_i+\alpha_i \}.
\end{equation}
Here $\phi_1,\phi_2,\phi_3$ are the phases of the $z$'s and $n_i\alpha_i=2\pi$ for some $n_i\in\mathbb{Z}$ and not all $n_i$s are simultaneously zero.\footnote{Note that this is a more general identification than the thermal quotient we will need, where one can take $\alpha_i=0\, (\forall i\neq 1)$} However, to embed $\Gamma$ in $SO\left(6\right)$, it is more useful to decompose these complex numbers into 6 real coordinates that embed the $S^5$ into ${\mathbb R}^6$,
\begin{eqnarray}
\label{s5pt}
x_1=\cos\theta \, \cos\phi_1 &\quad& x_2=\cos\theta \, \sin\phi_1, \nonumber \\
x_3=\sin\theta \, \cos\psi \, \cos\phi_2 &\quad& x_4=\sin\theta \, \cos\psi \, \sin\phi_2, \nonumber \\
x_5=\sin\theta \, \sin\psi \, \cos\phi_3 &\quad& x_6=\sin\theta \, \sin\psi \, \sin\phi_3.
\end{eqnarray}
Now, we construct a coset representative in $SO\left(6\right)$ for this point $x$ with coordinates as in $(\ref{s5pt})$. To do so, we start with the point $\left(1,0,0,0,0,0\right)$ in $R^6$, and construct a matrix $g\left(x\right)$ that rotates this point, the north pole, to the generic point $x$. By construction, $g\left(x\right) \in SO\left(6\right)$, and there is a one-to-one correspondence between the points $x$ on $S^5$ and matrices $g\left(x\right)$, upto a right multiplication by an element of the $SO\left(5\right)$ which leaves the north pole invariant. Such a representative matrix $g\left(x\right)$ can be taken to be
\begin{equation}
\label{cosetrep}
g\left(x\right)=e^{i\phi_1Q_{12}}e^{i\phi_2Q_{34}}e^{i\phi_3Q_{56}}e^{i\psi Q_{35}}e^{i\theta Q_{13}},
\end{equation}
where $Q$'s are the generators of $SO\left(6\right)$. This is clearly an instance of a section in $G$ over $G/H$.
The action of the thermal quotient (\ref{quotient}) on the coset representative $g\left(x\right)$ is
\begin{equation}
\label{gamma}
\gamma:g\left(x\right) \mapsto g\left(\gamma\left(x\right)\right)= e^{i\alpha_1Q_{12}}e^{i\alpha_2Q_{34}}e^{i\alpha_3Q_{56}}\cdot x=\gamma\cdot g\left(x\right),
\end{equation}
where the composition `$\cdot$' is the usual matrix multiplication.\footnote{This gives the embedding of $\Gamma$ in $SO\left(6\right)$.} This section has the property (\ref{compatible}) that we demand from a thermal section. Hence, we choose the thermal section to be
\begin{equation}
\label{thermal}
\sigma_{th}\left(x\right)=g\left(x\right).
\end{equation}
This analysis can be repeated for any odd-dimensional sphere to find the same expression for the thermal section. Essentially the only difference is that for a $(2n+1)$-dimensional sphere, we need to consider $(n+1)$ complex numbers $z_i$ and proceed exactly as above. We find that the thermal section can always be chosen to be (\ref{thermal}).
\subsection{The Method of Images}
Since the heat kernel obeys a linear differential equation- the heat equation- we can use the method of images to construct the heat kernel on $\Gamma\backslash G/H$ from that on $G/H$ (see, for example \cite{Giombi:2008vd}). The relation between the two heat kernels is
\begin{equation}
\label{kernelonquotients}
K^{\left(S\right)}_{\Gamma}\left(x,y;t\right)=\sum_{\gamma\in\Gamma}K^{\left(S\right)}\left(x,\gamma\left(y\right);t\right),
\end{equation}
where $K^{\left(S\right)}_{\Gamma}$ is the heat kernel between two points $x$ and $y$ on $\Gamma\backslash G/H$, $K^{\left(S\right)}$ is the heat kernel on $G/H$ and the spin indices have been suppressed.
We shall use this relation to determine the traced heat kernel on the thermal quotient of $S^{2n+1}$.
\subsection{The Traced Heat Kernel on thermal $S^{2n+1}$}
We use the formalism developed above to evaluate the traced heat kernel for the thermal quotient of the odd-dimensional sphere. Using the method of images, the quantity of interest is
\begin{equation}
K^{\left(S\right)}_{\Gamma}\left(t\right)=\sum_{k\in\mathbb Z_N} \int_{\Gamma\backslash G/H}d\mu\left(x\right)\sum_a K^{\left(S\right)}_{aa}\left(x,\gamma^k\left(x\right);t\right),
\end{equation}
where $d\mu(x)$ is the measure on $\Gamma\backslash G/H$ obtained from the Haar measure on $G$, and $x$ labels points in $\Gamma\backslash G/H$.
We have also used the fact that $\Gamma\simeq\mathbb Z_N$, and since the sum over $m$ is a finite sum, the integral has also been taken through the sum.
Further, since
\begin{equation}
\sum_a K^{\left(S\right)}_{aa}\left(\gamma x,\gamma^k\left(\gamma x\right);t\right)= \sum_a K^{\left(S\right)}_{aa}\left(x,\gamma^k\left(x\right);t\right),
\end{equation}
the integral over $\Gamma\backslash G/H$ can be traded in for the integral over $G/H$. We therefore multiply by an overall volume factor, and evaluate
\begin{equation}
\int_{G/H} d\mu\left(x\right) \sum_a K^{\left(S\right)}_{aa}\left(x,\gamma^k\left(x\right);t\right),
\end{equation}
where $d\mu(x)$ is the left invariant measure on $G/H$ obtained from the Haar measure on $G$, and $x$ now labels points in the full coset space $G/H$. Putting the expression (\ref{kernel}) into this, and choosing the section (\ref{thermal}) we obtain
\begin{equation}
\int_{G/H} d\mu\left(x\right) \sum_a K^{\left(S\right)}_{aa}\left(x,\gamma^k\left(x\right);t\right)=\int_{G/H} d\mu\left(x\right)\sum_R a_R^{(S)} {\rm Tr}_S\left(g_x^{-1}\gamma^k g_x\right)^{\left(R\right)} e^{tE_R^{(S)}},
\end{equation}
where $\left(g_x^{-1}\gamma^k g_x\right)^{\left(R\right)}$ is an abbreviation for ${{\mathcal U}^{\left(R\right)}}\left(g\left(x\right)^{-1}\gamma^k g\left(x\right)\right)$. As this expression stands, the trace is only over some subspace $\mathcal V_S \subset \mathcal V_R$ so the cyclic property of the trace cannot be used to annihilate the $g_x^{-1}$ with the $g_x$. To proceed further, we move the integral into the summation to obtain
\begin{equation}
\int_{G/H} d\mu\left(x\right) \sum_a K^{\left(S\right)}_{aa}\left(x,\gamma^k\left(x\right);t\right)=\sum_R a_R^{(S)} \int_{G/H} d\mu\left(x\right){\rm Tr}_S\left(g_x^{-1}\gamma^k g_x\right)^{\left(R\right)} e^{tE_R^{(S)}}.
\end{equation}
Since $G$ and $H$ are compact, we may use the property that \cite{Helgason}
\begin{equation}
\label{integraldecomposition}
\int_G dg f\left(g\right)=\int_{G/H}d\mu(\tilde{x})\left[\int_H dh f\left(\tilde{x}h\right)\right],
\end{equation}
where $dg$ is the Haar measure on $G$, $d\mu$ and $dh$ are the invariant measures on $G/H$ and $H$ respectively. $\tilde{x}$ is an arbitrary choice of coset representatives that we make to label points in $G/H$. In what follows we shall choose the coset representative to be $g_x$. Let us consider the function
\begin{equation}
f\left(g\right)= {\rm Tr}_S\left(g^{-1}\gamma^k g\right)^{\left(R\right)}.
\end{equation}
This function has the property that $f\left(g_x h\right)=f\left(g_x\right)$.
Putting this in (\ref{integraldecomposition}), we see that the integral over $H$ becomes trivial, and we get
\begin{equation}
\label{integralrelation}
\int_{G/H} d\mu\left(x\right) {\rm Tr}_S\left(g_x^{-1}\gamma^k g_x\right)^{\left(R\right)}=\frac{1}{V_H}\int_G dg {\rm Tr}_S\left(g^{-1}\gamma^k g\right)^{\left(R\right)}.
\end{equation}
Now, as in \cite{David:2009xg}, we note that the integral $I_G=\int_G dg \left(g^{-1}\gamma^k g\right)$ commutes with all group elements $\tilde{g}\in G$ because
\begin{equation}
I_G\cdot \tilde{g}=\int_G dg \left(g^{-1}\gamma^k g\right)\cdot \tilde{g}=\tilde{g}\cdot \int_G d\left(g\tilde{g}\right) \left(\left(g\tilde{g}\right)^{-1}\gamma^k \left(g\tilde{g}\right)\right)=\tilde{g}\cdot I_G,
\end{equation}
where we have used the right invariance of the measure, viz. $dg=d\left(g\tilde{g}\right)$. We therefore have, from Schur's lemma, that
\begin{equation}
I_G=\int_G dg \left(g^{-1}\gamma^k g\right)\varpropto \mathbb{I},
\end{equation}
from which we obtain that
\begin{equation}
\label{tracerel}
\int_G dg {\rm Tr}_S\left(g^{-1}\gamma^k g\right)^{\left(R\right)}=\frac{d_S}{d_R}\int_G dg {\rm Tr}_{R}\left(g^{-1}\gamma^k g\right)^{\left(R\right)}
=\frac{d_S}{d_R}V_G {\rm Tr}_R\left(\gamma^k\right).
\end{equation}
The quotient $\gamma$ is just the exponential of the Cartan generators of $SO\left(2n+2\right)$ (see, for example (\ref{gamma}) for the $S^5$). This trace, therefore is just the $SO\left(2n+2\right)$ character $\chi_R$ in the representation $R$.
Putting this result in (\ref{integralrelation}), we find that
\begin{equation}
\int_{G/H} d\mu\left(x\right) {\rm Tr}_S\left(g_x^{-1}\gamma^k g_x\right)^{\left(R\right)}=V_{G/H}\frac{d_S}{d_R}\chi_{R}\left(\gamma^k\right),
\end{equation}
where we have normalised volumes so that $V_{G}=V_{G/H}V_{H}.$ Therefore,
\begin{equation}
\sum_{k\in {\mathbb Z}_N}\int_{G/H} d\mu\left(x\right)K^{\left(S\right)}\left(x,\gamma^k\left(x\right);t\right)
=\sum_{k\in {\mathbb Z}_N}\sum_R \chi_{R} \left(\gamma^k\right) e^{tE_R^{(S)}},
\end{equation}
where we have inserted the value of the normalisation constant $a_R^{\left(S\right)}$ from (\ref{normalisation}). Now for the thermal quotient, we have $\gamma$ such that
\begin{equation}
\label{thermalsphere}
\alpha_1\neq 0,\, \alpha_i=0,\quad\forall\, i=2,\ldots,n+1.
\end{equation}
The volume factor for this quotient is just $\frac{\alpha_1}{2\pi}$. This gives us the traced heat kernel on the thermal $S^{2n+1}$
\begin{equation}
\label{s5intkernel}
K^{\left(S\right)}_{\Gamma}\left(t\right) =\frac{\alpha_1}{2\pi}\sum_{k\in {\mathbb Z}_N}\sum_R \chi_{R} \left(\gamma^k\right) e^{tE_R^{(S)}}.
\end{equation}
We find that the answer assembles naturally into a sum of characters, in various representations of $G$, of elements of the quotient group $\Gamma$. Here the representations $R$ of $SO(2n+2)$ are those which contain $S$ when restricted to $SO(2n+1)$. The reader should compare this expression to the equation $\left(4.20\right)$ obtained in \cite{David:2009xg}.
\section{The Heat Kernel on $AdS_{2n+1}$}
\label{kernelAdS}
We have so far obtained an expression for the heat kernel on a compact symmetric space (\ref{untracedkernel}) and have extended our analysis to its left quotients. In particular, we have shown how the traced heat kernel on (a class of) quotients of $S^{2n+1}$ assembles into a sum over characters of the orbifold group $\Gamma$. We now extend our analysis to the hyperbolic space ${\mathbb{H}}_{2n+1}$. Following the analysis of \cite{David:2009xg}, we will use the fact that the $N$-dimensional sphere admits an analytic continuation to the corresponding euclidean $AdS$ geometry. We now give an account of how one can exploit this fact to determine the heat kernel on $AdS_{2n+1}$ and its quotients.
\subsection{Preliminaries}
Euclidean $AdS$ is the $N$ dimensional hyperbolic space ${\mathbb H}_N^+$ which is the coset space
\begin{equation}
{\mathbb H}_N\simeq SO\left(N,1\right)/SO\left(N\right),
\end{equation}
where the quotienting is done, as in the sphere, by the right action. As was done for the three-dimensional case, we will view $SO\left(N,1\right)$ as an analytic continuation of $SO\left(N+1\right)$.
For explicit expressions, we shall employ the generalisation of the triple-polar coordinates that we introduced in (\ref{s5pt}) to the general $S^{2n+1}$. In these coordinates, the $S^{2n+1}$ metric is
\begin{equation}
\label{s5metric}
d\theta^2+cos^2\theta d\phi_1^2 + sin^2\theta\, d\Omega^2_{2n-1}.
\end{equation}
We perform the analytic continuation
\begin{equation}
\label{continuation}
\theta \mapsto -i\rho, \quad \phi_1 \mapsto it,
\end{equation} where $\rho$ and $t$ take values in $\mathbb{R}$, to obtain
\begin{equation}
ds^2=-\left(d\rho^2+cosh^2\rho dt^2 + sinh^2\rho\, d\Omega^2_{2n-1}\right).
\end{equation}
The reader will recognize this as the metric on global $AdS_{2n+1}$, upto a sign.
Now, to construct eigenfunctions on ${\mathbb{H}}_N$, we need to write down a section in $SO(N,1)$. The Lie algebra of $SO(N,1)$ is an analytic continuation of $SO(N+1)$ where we choose a particular axis- say `1'- as the time direction and perform the analytic continuation $Q_{1j}\rightarrow iQ_{1j}$ to obtain the $so(N,1)$ algebra from the $so(N+1)$ algebra. This is equivalent to the analytic continuation of the coordinates described above. Therefore, the section in $SO(N,1)$ can be obtained from that in $SO\left(N+1\right)$ by analytically continuing the coordinates via (\ref{continuation}).
\subsection{Harmonic Analysis on ${\mathbb{H}}_{2n+1}$}
\label{HarmonicAnalysisAdS}
We have recollected basic results from harmonic analysis on coset spaces in Section \ref{HarmonicAnalysis}, which we have exploited for compact groups $G$ and $H$. In fact, all the basic ingredients that we have employed in our analysis can be carried over to the case of non-compact groups as well. The eigenvalues of the spin-$S$ Laplacian are still given by (\ref{eigenval}), and the eigenfunctions are still (\ref{eigenf}), \textit{i.e.} they are determined by matrix elements of unitary representations of $G$. These unitary representations are now infinite dimensional, given that $G$ is non-compact. However, for $SO(N,1)$, these representations have been classified \cite{Hirai,Camporesi}, and we shall use these results to determine the traced heat kernel on $AdS_{2n+1}$.
The only unitary representations of $SO(N,1)$ that are relevant to us are those that contain unitary representations of $SO(N)$. For odd-dimensional hyperboloids, where $N=2n+1$, these are just the so-called principal series representations of $SO\left(2n+1,1\right)$ which are labelled by the array
\begin{equation}
\label{principalseries}
R=\left(i\lambda,m_2,m_3,\cdots,m_{n+1}\right), \quad \lambda \in {\mathbb R},\,m_2\geq m_3\geq\cdots\geq m_n\geq\vert m_{n+1}\vert,
\end{equation} where the $m_2,\cdots ,m_n$ and $\vert m_{n+1}\vert$ are non-negative (half-)integers. We shall usually denote the array $\left(m_2,m_3,\ldots,m_{n+1}\right)$ by $\vec{m}$. We also note that the principal series representations $\left(i\lambda,\vec{m}\right)$ that contain a representation $S$ of $SO\left(2n+1\right)$ are determined by the branching rules \cite{Ottoson,Camporesi}
\begin{equation}
\label{adsbranching}
s_1 \geq m_2\geq s_2\geq \ldots \geq m_n\geq s_n\geq\vert m_{n+1}\vert,
\end{equation}
which, for the STT tensors just reduce to $m_2=s$ with all other $m_i$s and $s_i$s set to zero (\textit{cf.} \ref{snbranchingSTT}), except for the case of $n=1$, where the branching rule is $\vert m_2\vert=s$.
Comparing (\ref{principalseries}) to (\ref{so2nrep}) suggests that the appropriate analytic continuation is
\begin{equation}
m_1\mapsto i\lambda-n, \quad \lambda\in\mathbb{R}_+,
\end{equation}
which is indeed the analytic continuation used by \cite{Camporesi:1994ga}.
Let us consider how the eigenvalues $E_R^{\left(S\right)}$ transform under this analytic continuation. It turns out that the eigenvalues (\ref{eigenvals2n1}) get continued to
\begin{equation}
E_{R,AdS_{2n+1}}^{\left(S\right)}=-\left(\lambda^2+\zeta\right),\quad \zeta\equiv C_2\left(S\right)-C_2\left(\vec{m}\right)+n^2,
\end{equation}
and that the eigenvalue for the STT tensors (\ref{eigenvals2n1stt}) gets continued to
\begin{equation}
E_{R,AdS_{2n+1}}^{\left(S\right)}=-\left(\lambda^2+ s+n^2\right).
\end{equation}
The eigenvalues on $AdS$ have an extra minus sign apart from what is obtained by the analytic continuation because the metric $S^{2n+1}$ under the analytic continuation goes to minus of the metric on $AdS_{2n+1}$.
This analytic continuation preserves the corresponding energy eigenvalue as a negative definite real number, which it must, because the Laplacian on Euclidean $AdS$ is an elliptic operator, and its eigenvalues must be of definite sign.
\subsection{The coincident Heat Kernel on $AdS_{2n+1}$}
In computing the heat kernel over $AdS_{2n+1}$ by analytically continuing from $S^{2n+1}$, the sum over $m_1$ that entered in (\ref{kernelsphere}), (\ref{kernelsttsphere}) and (\ref{kerneltracesttsphere}) is now continued to an integral over $\lambda$. In general this integral over $\lambda$ is hard to perform, but it simplifies significantly in the coincident limit and is evaluated below for this case. The traced heat kernel for STT tensors has previously been obtained directly in this limit by \cite{Camporesi:1994ga} and this calculation therefore serves as a check of the prescription (\ref{continuation}) of analytic continuation that we have employed. We will also see that the normalisation constant $a_{R}^S$ that appeared for the $S^{2n+1}$ gets continued to $\mu_{R}^S$, which is essentially the measure for this integral. This is a brief summary of the calculation, the reader will find more details in Appendix \ref{plancherelcalc}.
On using (\ref{dimensions}) for the special case of $R=\left(m,s,0,\ldots,0\right)$, one can show that $a_R^S$ gets continued via (\ref{continuation}) to $\mu_R^S$, where
\begin{equation}
\mu_R^S=\frac{1}{d_s}\frac{\left[\lambda^2+\left(s+\frac{N-3}{2}\right)^2\right]\prod_{j=0}^{\frac{N-5}{2}}\left(\lambda^2+j^2\right)} {2^{N-1}\pi^{\frac{N}{2}}\Gamma\left(\frac{N}{2}\right)} \frac{\left(2s+N-3\right)\left(s+N-4\right)!}{s!\left(N-3\right)!},
\end{equation}
where $N=2n+1$. A little algebra reveals this as the combination
\begin{equation}
\mu_R^S=\frac{C_N g\left(s\right)}{d_S}\frac{\mu\left(\lambda\right)}{\Omega_{N-1}},
\end{equation}
in the notation of \cite{Camporesi:1994ga}, (see their expressions $2.7$ to $2.13$ and $2.108$ or our Appendix \ref{plancherelcalc}). We have omitted the overall sign $\left(-1\right)^{\frac{N-1}{2}}$ in writing the above. The quantity $\mu\left(\lambda\right)$ is known as the Plancherel measure.
We now consider the expression (\ref{kerneltracesttsphere}) on $S^{2n+1}$, which in the coincident limit, reduces to
\begin{equation}
K^{S}\left(x,x;t\right) = \sum_{m_1} a_R^S d_S e^{tE_R^{(S)}}.
\end{equation}
where $a_R^S d_S=\frac{n!}{2\pi^{n+1}}d_{\left(m_1,s\right)}$. We analytically continue this expression via our prescription (\ref{continuation}) to the coincident heat kernel on $AdS_{2n+1}$.
\begin{equation}
K^{S}\left(x,x;t\right) = \int d\lambda\, \mu_R^S\left(\lambda\right) d_S e^{tE_R^{(S)}}= \frac{C_N}{\Omega_{N-1}} g\left(s\right) \int d\lambda\,\mu\left(\lambda\right) e^{-t\left(\lambda^2+ s+n^2\right),}
\end{equation}
which is precisely the expression obtained by \cite{Camporesi:1994ga}.
\section{The Heat Kernel on Thermal $AdS_{2n+1}$}
\label{kernelthermalads}
\subsection{The Thermal Quotient of $AdS$}
We are now in a position to calculate the traced heat kernel of an arbitrary tensor particle on thermal $AdS_{2n+1}$. This space is the hyperbolic space ${\mathbb H}_{2n+1}$ with a specific $\mathbb Z$ identification (in the generalised polar coordinates)
\begin{equation}
t\sim t+\beta,\quad \beta=i\alpha_1
\end{equation}
which is just the analytic continuation by (\ref{continuation}) of the identification (\ref{thermalsphere}) on the sphere. Since $t$ is the global time coordinate, $\beta$ is to be interpreted as the inverse temperature.
\subsection{The Heat Kernel}
In section \ref{kernelAdS} we have discussed how the heat kernel on ${\mathbb H}_{2n+1}$ can be calculated by analytically continuing the harmonic analysis on $S^{2n+1}$ to ${\mathbb H}_{2n+1}$. As discussed in Sec. 6.2 of \cite{David:2009xg}, we expect to be able to continue the expressions for the heat kernel on the thermal sphere to thermal $AdS$, with the difference that now $\Gamma\simeq\mathbb Z$, rather than $\mathbb Z_N$. Also, as noted in \cite{David:2009xg}, essentially the only difference that arises for the traced heat kernel is that the character of $SO(2n+2)$ that appears in (\ref{s5intkernel}) is now replaced by the Harish-Chandra (or global) character for the non-compact group $SO(2n+1,1)$.
With these inputs, the traced heat kernel on thermal $AdS_{2n+1}$ is given by
\begin{equation}
\label{ads5intkernel}
K^{\left(S\right)}\left(\gamma,t\right)
=\frac{\beta}{2\pi}\sum_{k\in {\mathbb Z}}\sum_{\vec{m}}\int_0^\infty d\lambda\, \chi_{\lambda,\vec{m}} \left(\gamma^k\right) e^{tE_R^{(S)}},
\end{equation}
where $\chi_{\lambda,\vec{m}}$ is the Harish-Chandra character in the principal series of $SO\left(2n+1,1\right)$, which has been evaluated \cite{HiraiChar} to be
\begin{equation}
\label{fullHCchar}
\chi_{\lambda,\vec{m}}\left(\beta,\phi_1,\phi_2,\ldots,\phi_n\right)=
\frac{e^{-i\beta\lambda}\chi^{SO\left(2n\right)}_{\vec{m}}\left(\phi_1,\phi_2,\ldots,\phi_n\right)+e^{i\beta\lambda}\chi^{SO\left(2n\right)}_{\check{\vec{m}}}\left(\phi_1,\phi_2,\ldots,\phi_n\right)}{e^{-n\beta}\prod_{i=1}^n\vert e^\beta - e^{i\phi_i}\vert^2},
\end{equation}
for the group element
\begin{equation}
\gamma=e^{i\beta Q_{12}}e^{i\phi_1 Q_{23}}\ldots e^{i\phi_n Q_{2n+1,2n+2}},
\end{equation}
where ${\check{\vec{m}}}$ is the conjugated representation, with the highest weight $\left(m_2,\ldots,-m_{n+1}\right)$, and $\chi^{SO\left(2n\right)}_{\vec{m}}$ is the character in the representation $\vec{m}$ of $SO\left(2n\right)$.
The sum over $\vec{m}$ that appears in (\ref{ads5intkernel}) is the sum over permissible values of $m$ as determined by the branching rules (\ref{adsbranching}). We also recall that `1' is the time-like direction.
For the thermal quotient, we have $\beta\neq 0$ and $\phi_i=0\;\forall\, i$. The character of $SO\left(2n\right)$ that appears in the character formula (\ref{fullHCchar}) above is then just the dimension of the corresponding representation. Using the fact that the dimensions of this representation is equal to that of its conjugate, we have for the character
\begin{equation}
\label{thermalHCchar}
\chi_{\lambda,\vec{m}}\left(\beta,\phi_1,\phi_2,\ldots,\phi_n\right)=
\frac{cos\left(\beta\lambda\right)}{2^{2n-1}\sinh^{2n}{\beta\over 2}}d_{\vec{m}}.
\end{equation}
We therefore obtain, for the traced heat kernel $AdS_{2n+1}$ (\ref{ads5intkernel}),
\begin{equation}
K^{\left(S\right)}\left(\beta,t\right)
=\frac{\beta}{2^{2n}\pi}\sum_{k\in {\mathbb Z}}\sum_{\vec{m}}d_{\vec{m}}\int_0^\infty d\lambda\, \frac{cos\left(k\beta\lambda\right)}{\sinh^{2n}{k\beta\over 2}} e^{-t\left(\lambda^2+\zeta\right)}.
\end{equation}
The integral over $\lambda$ is a Gaussian integral, which we can evaluate to obtain
\begin{equation}
K^{\left(S\right)}\left(\beta,t\right)
=\frac{\beta}{2^{2n}\sqrt{\pi t}}\sum_{k\in {\mathbb Z}_+}\sum_{\vec{m}}d_{\vec{m}}\frac{1}{\sinh^{2n}{k\beta\over 2}}e^{-\frac{k^2 \beta ^2}{4t}-t\zeta},
\end{equation}
where we have dropped the term with $k=0$, which diverges. This divergence arises due to the infinite volume of $AdS$, over which the coincident heat kernel on the full $AdS_{2n+1}$ is integrated. It can be reabsorbed into a redefinition of parameters of the gravity theory under study and is independent of $\beta$ and is therefore not of interest to us.
This expression further simplifies for the case of the STT tensors. The branching rules determine that $\vec{m}=\left(s,0,\ldots,0\right)$ and therefore the sum over $\vec{m}$ gets frozen out and we obtain
\begin{equation}
K^{\left(S\right)}\left(\beta,t\right)
=\frac{\beta}{2^{2n}\sqrt{\pi t}}\sum_{k\in {\mathbb Z}_+}\frac{d_{\vec{m}}}{\sinh^{2n}{k\beta\over 2}}e^{-\frac{k^2 \beta ^2}{4t}-t\left(s+n^2\right)}.
\end{equation}
The reader may compare this expression to the equation $\left(3.9\right)$ obtained in \cite{David:2009xg} for the $AdS_3$ case (where one would have to specialise to $\tau_1=0$, $\tau_2=\beta$, and $d_{\vec{m}}\equiv 1$, and further include a factor of $2$ that appears because the branching rule for $AdS_3$ leads us to sum over $m_2=\pm s$ rather than $m_2=s$ for $s>0$).
\section{The one-loop Partition Function}
As a consequence of the above, we can calculate the one-loop determinant of a spin-$S$ particle on $AdS_5$. To do so, we need the result that
\begin{equation}
\int_0^\infty \frac{dt}{t^{\frac{3}{2}}}e^{-\frac{\alpha^2}{4t}-\beta^2t}=\frac{2\sqrt{\pi}}{\alpha}e^{-\alpha\beta}.
\end{equation}
Then the one-loop determinants can be deduced from the heat kernel by using
\begin{equation}
-log\,det\left(-\Delta_{\left(S\right)}+m_S^2\right)=\int_0^\infty \frac{dt}{t}K^{\left(S\right)}\left(\beta,t\right)e^{-m_S^2t},
\end{equation}
which we can simplify to obtain
\begin{equation}
-log\,det\left(-\Delta_{\left(S\right)}+m_S^2\right)=\sum_{k\in {\mathbb Z}_+}\sum_{\vec{m}}d_{\vec{m}}\frac{2}{e^{-nk\beta}\left( e^{k\beta} -1\right)^{2n}}\frac{1}{k}e^{-k\beta\sqrt{\zeta+m_S^2}}.
\end{equation}
This expression further simplifies for the case of STT tensors and one obtains
\begin{equation}
-log\,det\left(-\Delta_{\left(S\right)}+m_S^2\right)=\sum_{k\in {\mathbb Z}_+}d_{\vec{m}}\frac{2}{e^{-nk\beta}\left( e^{k\beta} -1\right)^{2n}}\frac{1}{k}e^{-k\beta\sqrt{s+n^2+m_S^2}}.
\end{equation}
\subsection{The scalar on $AdS_5$}
Let us evaluate the above expression for scalars in $AdS_5$, where $s=0$. In units where the $AdS$ radius is set to one, $\sqrt{m_S^2+4}=\Delta-2$, where $\Delta$ is the conformal dimension of the scalar. We therefore have
\begin{equation}
-log\,det\left(-\Delta_{\left(S\right)}+m_S^2\right)=\sum_{k\in {\mathbb Z}_+}\frac{2}{k\left(1-e^{-k\beta}\right)^{4}}e^{-k\beta\Delta}
\end{equation}
We can evaluate the sum to find that the one-loop determinant is given by
\begin{equation}
-log\,det\left(-\Delta_{\left(S\right)}+m_S^2\right)=-2\sum_{n=0}^\infty \frac{\left(n+1\right)\left(n+2\right)\left(n+3\right)}{6}log\left(1-e^{-\beta\left(\Delta+n\right)}\right).
\end{equation}
Now since $logZ_{\left(S\right)}=-\frac{1}{2}log\,det\left(-\Delta_{\left(S\right)}+m_S^2\right)$, we have, for the one-loop partition function of a scalar,
\begin{equation}
\label{scalarpartition}
logZ_{\left(S\right)}=-\sum_{n=0}^\infty \frac{\left(n+1\right)\left(n+2\right)\left(n+3\right)}{6}log\left(1-e^{-\beta\left(\Delta+n\right)}\right).
\end{equation}
This expression matches exactly with that which is obtained with the method of \cite{Denef:2009kn} (see also, earlier work by \cite{Hartman:2006dy,Diaz:2007an}).
\textit{Note:} After version 1 of this paper appeared on the arXiv, we learned of \cite{Gibbons:2006ij} where expressions for the one-loop partition function for general spin on thermal $AdS$ were obtained by means of a Hamiltonian computation\footnote{We thank Gary Gibbons for bringing this to our attention.}. Our results agree with the expressions obtained there.
\section{Conclusions}
We have computed the principal ingredients that go into the calculation of one loop effects on odd dimensional thermal AdS spacetimes. As mentioned in the introduction, there are many potential applications of these results. Specifically, in the context of investigating higher spin gauge fields in these spacetimes. It would also be useful to complete the analysis for the even dimensional AdS spacetimes as well. Finally, the ambitious goal, which was the initial motivation for this work, is to obtain some clues about the one loop partition function of the string theory on AdS. The heat kernel method is ideally suited for this purpose and we expect our explicit results will be helpful in this regard.
\section*{Acknowledgments}
We would like to thank M. Gaberdiel for helpful comments on the manuscript. We would also like to thank Shamik Banerjee, Andrea Campoleoni, Justin David, Dileep Jatkar, and Suvrat Raju for helpful discussions. The work of R.G. was supported by a Swarnajayanthi Fellowship of the DST, Govt. of India and more generally by the generosity of the Indian people. The work of R. Gupta is a part of research programme of FOM, which is financially supported by the Netherlands Organization for Scientific Research (NWO).
|
\section{Introduction}
A non-trivial generalization of Chernoff bound type inequalities for matrix-valued random variables was introduced by Ahlswede and Winter~\cite{chernoff:matrix_valued:AW}. In parallel, Vershynin and Rudelson introduced similar matrix-valued concentration inequalities using different machinery~\cite{rudelson:isotropic,lowrank:rankone:VR}. Following these two seminal papers, many variants have been proposed in the literature~\cite{recht:simple_completion}; see~\cite{chernoff:matrix_valued:Tropp} for more. Such inequalities, similarly to their real-valued ancestors, provide powerful tools to analyze probabilistic constructions and the performance of randomized algorithms. There is a rapidly growing line of research exploiting the power of these inequalities including new proofs of probabilistic constructions of expander graphs~\cite{expander:AlonRoichman:orig,expander:AlonRoichman:RusLan,chernoff:matrix_valued:Azuma_Naor}, matrix approximation by element-wise sparsification~\cite{matrix:sparsification:IPL2011}, graph approximation via edge sparsification~\cite{graph:sparsifiers:eff_resistance}, analysis of algorithms for matrix completion and decomposition of low rank matrices~\cite{recht:simple_completion,chernoff:matrix_valued:MZ11}, semi-definite relaxation and rounding of quadratic maximization problems~\cite{chernoff:matrix_valued:opt:journal}.
In many settings, it is desirable to convert the above probabilistic proofs into \emph{efficient} deterministic procedures. That is, to derandomize the proofs. Wigderson and Xiao presented an efficient derandomization of the matrix Chernoff bound by generalizing Raghavan's method of pessimistic estimators to the matrix-valued setting~\cite{chernoff:matrix_valued:derand:WX08}. In this paper, we generalize Spencer's hyperbolic cosine algorithm to the matrix-valued setting~\cite{hyperbolic_cosine:Spencer}. In an earlier, preliminary version of our paper~\cite{matrix:hypercosine_zouzias} the generalization of Spencer's hyperbolic cosine algorithm was also based on the method of pessimistic estimators. However, here we present a proof which is based on a simple averaging argument. Next, we carefully analyze two special cases of matrices; one in which the matrices have a group structure and the other in which they have rank-one. We apply our main result to the following problems: deterministically constructing Alon-Roichman expanding Cayley graphs, approximating graphs via edge sparsification and approximating matrices via element-wise sparsification.
The Alon-Roichman theorem asserts that Cayley graphs obtained by choosing a logarithmic number of group elements independently and uniformly at random are expanders~\cite{expander:AlonRoichman:orig}. The original proof of Alon and Roichman is based on Wigner's trace method, whereas recent proofs rely on matrix-valued deviation bounds~\cite{expander:AlonRoichman:RusLan}. Wigderson and Xiao's derandomization of the matrix Chernoff bound implies a deterministic $\mathcal{O}(n^4 \log n )$ time algorithm for constructing Alon-Roichman graphs. Independently, Arora and Kale generalized the multiplicative weights update (MWU) method to the matrix-valued setting and, among other interesting implications, they improved the running time to $\mathcal{O}(n^3\polylog{n})$~\cite{phdthesis:Kale:2008}. Here we further improve the running time to $\mathcal{O}(n^2 \log^3 n)$ by exploiting the group structure of the problem. In addition, our algorithm is combinatorial in the sense that it only requires counting the number of all closed (even) paths of size at most $\mathcal{O}(\log n)$ in Cayley graphs. All previous algorithms involve numerical matrix computations such as eigenvalue decompositions and matrix exponentiation.
The second problem that we study is the graph sparsification problem. This problem poses the question whether any dense graph can be approximated by a sparse graph under different notions of approximation. Given any undirected graph, the most well-studied notions of approximation by a sparse graph include approximating, \emph{all} pairwise distances up to an additive error~\cite{graph:spanners:PelegS89}, every cut to an arbitrarily small multiplicative error~\cite{graph:sparsifier:BenczurK96} and every eigenvalue of the difference of their Laplacian matrices to an arbitrarily small relative error~\cite{graph:sparsifier:ICM2010}; the resulting graphs are usually called \emph{graph spanners}, \emph{cut sparsifiers} and \emph{spectral sparsifiers}, respectively. Given that the notion of spectral sparsification is stronger than cut sparsification, so we focus on spectral sparsifiers. An efficient randomized algorithm to construct an $(1+\varepsilon)$-spectral sparsifier with $\mathcal{O}(n\log n /\varepsilon^2)$ edges was given in~\cite{graph:sparsifiers:eff_resistance}. Furthermore, an $(1+\varepsilon)$-spectral sparsifier with $\mathcal{O}(n/\varepsilon^2)$ edges can be computed in $\mathcal{O}(mn^3/\varepsilon^2)$ deterministic time~\cite{graph:sparsifiers:twice_ram}. The latter result is a direct corollary of the spectral sparsification of positive semi-definite (psd) matrices problem as defined in~\cite{phdthesis:Srivastava:2010}; see also~\cite{graph:sparsification:Naor} for more applications. Here we present a fast deterministic algorithm for spectral sparsification of psd matrices and, as a consequence, we obtain an improved deterministic spectral graph sparsification algorithm for the case of dense graphs.
The last problem that we analyze is the element-wise matrix sparsification problem. This problem was first introduced by Achlioptas and McSherry in~\cite{matrix:sparsification:optas}. They described sampling-based algorithms that select a small number of entries from an input matrix $A$, forming a sparse matrix $\widetilde{A}$, which is close to $A$ in the operator norm sense. The motivation to study this problem lies on the need to speed up several matrix computations including approximate eigenvector computations~\cite{matrix:sparsification:optas} and semi-definite programming solvers~\cite{fast_SDP:AHK05,Asp09}. Recently, there are many follow-up results on this problem~\cite{matrix:sparsification:arora,matrix:sparsification:IPL2011}. To the best of our knowledge, all known algorithms for this problem are randomized (see Table~$1$ of~\cite{matrix:sparsification:IPL2011}). In this paper we present the first deterministic algorithm and strong sparsification bounds for self-adjoint matrices that have an approximate diagonally dominant\footnote{A self-adjoint matrix $A$ of size $n$ is called \emph{diagonally dominant} if $|A_{ii}| \geq \sum_{j\neq i} |A_{ij}|$ for every $i\in{[n]}$.} property. Diagonally dominant matrices arise in many applications such as the solution of certain elliptic differential equations via the finite element method~\cite{SDD:Vavasis}, several optimization problems in computer vision~\cite{SDD:vision:Koutis} and computer graphics~\cite{SDD:graphics:Joshi}, to name a few.
\paragraph{Organization of the Paper.}
The paper is organized as follows. In~\S~\ref{sec:derand_Bernstein}, we present the matrix hyperbolic cosine algorithm (Algorithm~\ref{alg:derand:Bernstein}). We apply the matrix hyperbolic cosine algorithm to derive improved deterministic algorithms for the construction of Alon-Roichman expanding Cayley graphs in~\S~\ref{sec:AR_graphs}, spectral sparsification of psd matrices in~\S~\ref{sec:fast_isotrop_sparse} and element-wise matrix sparsification in \S~\ref{sec:sparsification:matrix}. Due to space constraints, almost all proofs have been deferred to the Appendix.
\subsection*{Our Results}
The main contribution of this paper is a generalization of Spencer's hyperbolic cosine algorithm to the matrix-valued setting~\cite{hyperbolic_cosine:Spencer},~\cite[Lecture~$4$]{book:probmeth:Spencer}, see Algorithm~\ref{alg:derand:Bernstein}. As mentioned in the introduction, our main result has connections with a recent derandomization of matrix concentration inequalities~\cite{chernoff:matrix_valued:derand:WX08}. We should highlight a few advantages of our result compared to~\cite{chernoff:matrix_valued:derand:WX08}. First, our construction does not rely on composing two separate estimators (or potential functions) to achieve operator norm bounds and does not require knowledge of the sampling probabilities of the matrix samples as in~\cite{chernoff:matrix_valued:derand:WX08}. In addition, the algorithm of~\cite{chernoff:matrix_valued:derand:WX08} requires computations of matrix expectations with matrix exponentials which are computationally expensive, see~\cite[Footnote~$6$, p. $63$]{chernoff:matrix_valued:derand:WX08}. In this paper, we demostrate that overcoming these limitations leads to faster and in some cases simpler algorithms.
Next, we demonstrate the usefulness of the main result by analyzing its computational efficiency under two special cases of matrices. We begin by presenting the following result
\begin{theorem}[Restatement of Theorem~\ref{thm:AR_graphs}]
There is a deterministic algorithm that, given the multiplication table of a group $G$ of size $n$, constructs an Alon-Roichman expanding Cayley graph of logarithmic degree in $\mathcal{O}(n^2\log^3 n)$ time. Moreover, the algorithm performs only group algebra operations that correspond to counting closed paths in Cayley graphs.
\end{theorem}
To the best of our knowledge, the above theorem improves the running time of all previously known deterministic constructions of Alon-Roichman Cayley graphs~\cite{arora:fast_SDP,chernoff:matrix_valued:derand:WX08,phdthesis:Kale:2008}. Moreover, notice that the running time of the above algorithm is optimal up-to poly-logarithmic factors since the size of the multiplication table of a finite group of size $n$ is $\mathcal{O}(n^2)$.
In addition, we study the computational efficiency of the matrix hyperbolic cosine algorithm on the case of matrix samples with rank-one. The motivation for studying this special setting is its connection with problems such as graph approximation via edge sparsification as was shown in~\cite{graph:sparsifiers:twice_ram,phdthesis:Srivastava:2010} and matrix approximation via element-wise sparsification as we will see later in this paper. The main result for this setting can be summarized in the following theorem (see \S~\ref{sec:fast_isotrop_sparse}), which improves the $\mathcal{O}(mn^3 /\varepsilon^2)$ running time of~\cite{phdthesis:Srivastava:2010} when, say, $m = \Omega(n^2)$ and $\varepsilon$ is a constant.
\begin{theorem}
Suppose $0 < \varepsilon < 1$ and $A = \sum_{i=1}^{m} v_i \otimes v_i $ are given, with column vectors $v_i\in\mathbb{R}^n $. Then there are non-negative real weights $\{s_i\}_{i\leq m}$, at most $ \lceil n/\varepsilon^2 \rceil$ of which are non-zero, such that
\begin{equation*}
(1-\varepsilon)^3 A \preceq \widetilde{A} \preceq (1+\varepsilon)^3 A,
\end{equation*}
where $\widetilde{A} = \sum_{i=1}^{m}s_i v_i \otimes v_i$. Moreover, there is a deterministic algorithm which computes the weights $s_i$ in\footnote{The $\widetilde{\mathcal{O}}(\cdot)$ notation hides $\log\log n$ and $\log\log (1/\varepsilon)$ factors throughout the paper.} $\widetilde{\mathcal{O}}(mn^2 \log^3 n /\varepsilon^2 + n^4 \log n /\varepsilon^4)$ time.
\end{theorem}
First, as we have already mentioned the graph sparsification problem can be reduced to spectral sparsification of positive semi-definite matrix. Hence as a corollary of the above theorem (proof omitted, see~\cite{phdthesis:Srivastava:2010} for details), we obtain a fast deterministic algorithm for sparsifying dense graphs, which improves the currently best known $\mathcal{O}(n^5 /\varepsilon^2 )$ running time for this problem.
\begin{corollary}
Given a weighted dense graph $H=(V,E)$ on $n$ vertices with positive weights and $0< \varepsilon <1$, there is a deterministic algorithm that returns an $(1+\varepsilon)$-spectral sparsifier with $\mathcal{O}(n/ \varepsilon^2)$ edges in $\widetilde{\mathcal{O}}(n^4 \log n /\varepsilon^2$ $ \max\{ \log^2 n, 1/\varepsilon^2 \})$ time.
\end{corollary}
Second, we give an elementary connection between element-wise matrix sparsification and spectral sparsification of psd matrices. A direct application of this connection implies strong sparsification bounds for self-adjoint matrices that are close to being \emph{diagonally dominant}. More precisely, we give two element-wise sparsification algorithms for self-adjoint and diagonally dominant-like matrices; in its randomized and the other in its derandomized version (see Table~$1$ of~\cite{matrix:sparsification:IPL2011} for comparison).
Here, for the sake of presentation, we state our results for diagonally dominant matrices, although the results hold under a more general setting (see \S~\ref{sec:sparsification:matrix} for details).
\begin{theorem}
Let $A$ be any self-adjoint and diagonally dominant matrix of size $n$ and $ 0 < \varepsilon <1$. Assume for normalization that $\norm{A}=1$.
\begin{enumerate}[(a)]
\item
There is a randomized linear time algorithm that outputs a matrix $\widetilde{A}\in \mathbb{R}^{n\times n}$ with at most $\mathcal{O}( n \log n /\varepsilon^2)$ non-zero entries such that, with probability at least $1-1/n$, $\norm{A - \widetilde{A}} \leq \varepsilon.$
\item
There is a deterministic $\widetilde{\mathcal{O}}( \varepsilon^{-2} \nnz{A} n^2 \log n \max\{ \log^2 n,1/\varepsilon^2 \} )$ time algorithm that outputs a matrix $\widetilde{A}\in \mathbb{R}^{n\times n}$ with at most $\mathcal{O}(n/\varepsilon^2)$ non-zero entries such that $\norm{A - \widetilde{A}} \leq \varepsilon.$
\end{enumerate}
\end{theorem}
\paragraph{Preliminaries.}
The next discussion reviews several definitions and facts from linear algebra; for more details, see~\cite{book:matrix:Bhatia}. By $[n]$ to be the set $\{1,2,\ldots , n \}$. We denote by $\mathcal{S}^{n\times n}$ the set of symmetric matrices of size $n$. Let $x\in\mathbb{R}^n$, we denote by $\diag{x}$ the diagonal matrix containing $x_1,x_2,\ldots ,x_n$. For a square matrix $M$, we also write $\diag{M}$ to denote the diagonal matrix that contains the diagonal entries of $M$. Let $A$ be an $m\times n$ matrix. $A^{(j)}$ will denote the $j$-th column of $A$ and $A_{(i)}$ the $i$-th row of $A$. We denote $\norm{A}=\max \{ \norm{Ax}~|~\norm{x} =1 \}$, $\infnorm{A} = \max_{i\in{[m]}} \sum_{j\in{[n]}} |A_{ij}|$ and by $\frobnorm{A}=\sqrt{\sum_{i,j}{A_{ij}^2}}$ the Frobenius norm of $A$. Also $\sr{A}:=\frobnorm{A}^2/\norm{A}^2$ is the \emph{stable rank} of $A$ and by $\nnz{A}$ the number of its non-zero entries. The trace of a square matrix $B$ is denoted as $\trace{B}$. We write $\mathbf{J}_n$ for the all-ones square matrices of size $n$. For two self-adjoint matrices $X,Y$, we say that $Y\succeq X$ if and only if $Y-X$ is a positive semi-definite (psd) matrix. Let $x\in\mathbb{R}^n$, then $x\otimes x$ is the $n\times n$ matrix such that $ (x \otimes x)_{i,j} =x_i x_j$. Given any matrix $A$, its \emph{dilation} is defined as $\dil{A} = \left[\begin{matrix}
\mathbf{0} & A \\
A^\top & \mathbf{0}
\end{matrix}\right].
$
It is easy to see that $\lambda_{\max}(\dil{A}) = \norm{A}$, see e.g.~\cite[Theorem~$4.2$]{book:perturbation:stewart}.
\paragraph{Functions of Matrices.}
Here we review some basic facts about the matrix exponential and the hyperbolic cosine function, for more details see~\cite{book:Higham:Matrix_fcn}. All proofs of this section have been deferred to the appendix. The matrix exponential of a self-adjoint matrix $A$ is defined as $\expm{A} = \mathbf{I} + \sum_{k=1}^{\infty} \frac{A^k}{k!}$. Let $A=Q\Lambda Q^\top$ be the eigendecomposition of $A$. It is easy to see that $\expm{A} = Q \expm{\Lambda} Q^\top$. For any real square matrices $A$ and $B$ of the same size that commute, i.e., $AB=BA$, we have that $ \expm{A+B} = \expm{A}\expm{B}$. In general, when $A$ and $B$ do not commute, the following estimate is known for self-adjoint matrices.
\begin{lemma}\cite{ineq:trace_exp:Golden,ineq:trace_exp:Thompson}\label{lem:ineq:golden_thompson}
For any self-adjoint matrices $A$ and $B$, $\trace{\expm{ A + B}} \leq \trace{\expm{A}\expm{B}}$.
\end{lemma}
\ignore{
\begin{lemma}
For any self-adjoint matrix $A\in{\mathbb{R}^{n\times n}}$ such that $\mathbf{0} \preceq A \preceq \mathbf{I}$, and any $\alpha,\beta \in\mathbb{R}$,
\[\expm{\alpha A + \beta(\mathbf{I}-A)} \preceq A \exp(\alpha) + (\mathbf{I}-A) \exp(\beta).\]
\end{lemma}
The Trotter product formula~\cite{expm:Trotter} states that
\begin{theorem}
Let $A,B$ be two self-adjoint matrices. Then
\[\expm{A+B} = \lim_{l\to \infty} \left(\expm{A/l} \expm{B/l}\right)^l\]
\end{theorem}
}We will also need the following fact about matrix exponential for rank one matrices.
\begin{lemma}\label{lem:expm:outerprod}
Let $x$ be a non-zero vector in $\mathbb{R}^n$. Then $ \expm{x \otimes x} = \mathbf{I}_n + \frac{\ensuremath{{\rm e}}^{\norm{x}^2} - 1}{\norm{x}^2} x\otimes x$.
Similarly, $\expm{-x \otimes x} = \mathbf{I}_n - \frac{1 - \ensuremath{{\rm e}}^{-\norm{x}^2}}{\norm{x}^2} x \otimes x$.
\end{lemma}
Let us define the \emph{matrix hyperbolic cosine} function of a self-adjoint matrix $A$ as $ \coshm{A} := (\expm{A} + \expm{-A}) /2$. Next, we state a few properties of the matrix hyperbolic cosine.
\begin{lemma}\label{lem:dil_vs_expm}
Let $A$ be a self-adjoint matrix. Then $\trace{\expm{\dil{A}}} = 2 \trace{ \coshm{A} }$.
\end{lemma}
\begin{lemma}\label{lem:coshm_with_proj}
Let $A$ be a self-adjoint matrix and $P$ be a projector matrix that commutes with $A$, i.e., $PA=AP$. Then $\coshm{PA} = P\coshm{A} + \mathbf{I} - P$.
\end{lemma}
\begin{lemma}\cite[Lemma~$2.2$]{Tsuda05}\label{lem:trace:incr_psd}
For any positive semi-definite self-adjoint matrix $A$ of size $n$ and any two self-adjoint matrices $B,C$ of size $n$, $B\preceq C$ implies $\trace{AB} \leq \trace{AC}$.
\end{lemma}
\vspace*{-4.0ex}
\section{Balancing Matrices: a matrix hyperbolic cosine algorithm}\label{sec:derand_Bernstein}
We briefly describe Spencer's balancing vectors game and then generalize it to the matrix-valued setting~\cite[Lecture~$4$]{book:probmeth:Spencer}. Let a two-player perfect information game between Alice and Bob. The game consists of $n$ rounds. On the $i$-th round, Alice sends a vector $v_i$ with $\infnorm{v_i}\leq 1$ to Bob, and Bob has to decide on a sign $s_i\in{\{\pm 1\}}$ knowing only his previous choices of signs and $\{v_{k}\}_{k < i}$.
At the end of the game, Bob pays Alice $\infnorm{\sum_{i=1}^{n} s_i v_i}$. We call the latter quantity, the \emph{value} of the game.
It has been shown in~\cite{Spencer:balanc_vct} that, in the above limited online variant, Spencer's six standard deviations bound~\cite{sixDeviation:Spencer} does not hold and the best value that we can hope for is $\Omega(\sqrt{n \ln n})$. Such a bound is easy to obtain by picking the signs $\{s_i\}$ uniformly at random. Indeed, a direct application of Azuma's inequality to each coordinate of the random vector $\sum_{i=1}^{n} s_i v_i$ together with a union bound over all the coordinates gives a bound of $\mathcal{O}(\sqrt{n\ln n})$.
Now, we generalize the balancing vectors game to the matrix-valued setting. That is, Alice now sends to Bob a sequence $\{M_i\}$ of self-adjoint matrices of size $n$ with\footnote{A curious reader may ask him/her-self why the operator norm is the right choice. It turns out the the operator norm is the correct matrix-norm analog of the $\ell_\infty$ vector-norm, viewed as the \emph{infinity} Schatten norm on the space of matrices.} $\norm{M_i}\leq 1$, and Bob has to pick a sequence of signs $\{s_i\}$ so that, at the end of the game, the quantity $\norm{\sum_{i=1}^{n} s_i M_i} $ is as small as possible. Notice that the balancing vectors game is a restriction of the balancing matrices game in which Alice is allowed to send only diagonal matrices with entries bounded in absolute value by one. Similarly to the balancing vectors game, using matrix-valued concentration inequalities, one can prove that Bob has a randomized strategy that achieves at most $\mathcal{O}(\sqrt{n\ln n})$ w.p. at least $1/2$. Indeed,
\begin{lemma}\label{lem:balanc_mtx}
Let $M_i \in \mathcal{S}^{n\times n}$, $\norm{M_i} \leq 1$, $1 \leq i \leq n$. Pick $s_i^*\in{\{\pm 1\} }$ uniformly at random for every $i\in{[n]}$. Then $\norm{\sum_{i=1}^{n} s_i^* M_i} = \mathcal{O}(\sqrt{ n \ln n})$ w.p. at least $1/2$.
\end{lemma}
Now, let's assume that Bob wants to achieve the above probabilistic guarantees using a \emph{deterministic} strategy. Is it possible? We answer this question in the affirmative by generalizing Spencer's hyperbolic cosine algorithm (and its proof) to the matrix-valued setting. We call the resulting algorithm \emph{matrix hyperbolic cosine} (Algorithm~\ref{alg:derand:Bernstein}). It is clear that this simple greedy algorithm implies a deterministic strategy for Bob that achieves the probabilistic guarantees of Lemma~\ref{lem:balanc_mtx} (set $f_j\sim s_j M_j$, $t=n$ and $\varepsilon = \mathcal{O}(\sqrt{ \ln n / n})$ and notice that $\gamma,\rho^2$ are at most one).
Algorithm~\ref{alg:derand:Bernstein} requires an extra assumption on its random matrices compared to Spencer's original algorithm. That is, we assume that our random matrices have uniformly bounded their ``matrix variance'', denoted by $\rho^2$. This requirement is motivated by the fact that in the applications that are studied in this paper such an assumption translates bounds that depend quadratically on the matrix dimensions to bounds that depend linearly on the dimensions.
We will need the following technical lemma for proving the main result of this section, which is a Bernstein type argument generalized to the matrix-valued setting~ \cite{chernoff:matrix_valued:Tropp}.
\begin{lemma}\label{lem:bounding_w}
Let $f:[m] \to \mathcal{S}^{n\times n}$ with $\norm{f(i)} \leq \gamma$ for all $i\in{[m]}$. Let $X$ be a random variable over $[m]$ such that $\ensuremath{\mathbb{E}}{f(X)}=\mathbf{0}$ and $\norm{\ensuremath{\mathbb{E}} f(X)^2 } \leq \rho^2$. Then, for any $\theta >0$, $\norm{\ensuremath{\mathbb{E}} [ \expm{ \dil{ \theta f(X)} }]}\ \leq\ \exp\left( \rho^2( \ensuremath{{\rm e}}^{\theta \gamma } -1 - \theta \gamma )/ \gamma^2\right).$ In particular, for any $0 < \varepsilon < 1$, setting $\theta = \varepsilon /\gamma$ implies that $\ensuremath{\mathbb{E}} [ \expm{ \dil{ \varepsilon f(X) / \gamma} }] \preceq \ensuremath{{\rm e}}^{\varepsilon^2 \rho^2 / \gamma^2} \mathbf{I}_{2n}$.
\end{lemma}
Now we are ready to prove the correctness of the matrix hyperbolic cosine algorithm.
\vspace*{-4.0ex}
\begin{algorithm}{}
\caption{Matrix Hyperbolic Cosine}\label{alg:derand:Bernstein}
\begin{algorithmic}[1]
\Procedure{Matrix-Hyperbolic}{$\{f_j\}$, $\varepsilon$, $t$}\Comment{$f_j:[m] \to \mathcal{S}^{n\times n}$ as in Theorem~\ref{thm:hypercosine:main}, $0 < \varepsilon < 1$.}
\State Set $\theta = \varepsilon /\gamma $
\For {$i=1$ to $t$}
\State Compute $x_i^*\in{[m]}$: $ x_i^* = \operatorname*{arg\; min}_{k\in{[m]}}\trace{\coshm{ \theta \sum_{j=1}^{i-1} f_j(x_j^*) + \theta f_i(k) }} $
\EndFor
\State \textbf{Output:} $t$ indices $x_1^*, x_2^*, \ldots ,x_t^*$ such that $\norm{ \frac1{t} \sum_{j=1}^{t} f_j(x_j^*) } \leq \frac{\gamma \ln( 2n)}{t\varepsilon } + \frac{\varepsilon\rho^2}{\gamma} $
\EndProcedure
\end{algorithmic}
\end{algorithm}
\vspace*{-4.0ex}
\begin{theorem}\label{thm:hypercosine:main}
Let $f_j:[m] \to \mathcal{S}^{n\times n}$ with $\norm{f_j(i)} \leq \gamma$ for all $i\in{[m]}$ and $j=1,2,\ldots$. Suppose that there exists independent random variables $X_1,X_2,\ldots $ over $[m]$ such that $\ensuremath{\mathbb{E}}{f_j(X_j)}=\mathbf{0}$ and $\norm{\ensuremath{\mathbb{E}} f_j(X_j)^2 } \leq \rho^2$. Algorithm~\ref{alg:derand:Bernstein} with input $\{f_j\},\varepsilon, t$ outputs a set of indices $\{x_j^*\}_{j\in{[t]}}$ over $[m]$ such that $ \norm{ \frac1{t}\sum_{j=1}^{t} f_j(x_j^*)} \leq \frac{ \gamma \ln (2n)}{t\varepsilon} + \frac{\varepsilon \rho^2}{\gamma}.$
\end{theorem}
We conclude with an open question related to Spencer's six standard deviation bound~\cite{sixDeviation:Spencer}. Does Spencer's six standard deviation bound holds under the matrix setting? More formally, given any sequence of $n$ self-adjoint matrices $\{M_i\}$ with $\norm{ M_i}\leq 1$, does there exist a set of signs $\{s_i\}$ so that $\norm{ \sum_{i=1}^{n} s_i M_i} = \mathcal{O}(\sqrt{n})$?
\ignore{
\paragraph{Connection with Arora-Kale's Matrix MWU Method.}
The balancing matrices game can be summarized as follows: At each round Alice presents to Bob a finite set of matrices with norm bounded by one and the promise that there is a convex combination of them that sums up to the all-zeros matrix (zero-mean condition). Bob selects one matrix from Alice's set based on his previous choices. Bob's goal is to keep the operator norm of the sum of his selected matrices as small as possible.
In Arora-Kale's setting~\cite{phdthesis:Kale:2008}, Alice is more powerful and the ``randomness'' is on Bob's side. At each round, Alice picks any psd self-adjoint matrix $M$ with norm at most one and present it to Bob. Bob maintains a probability distribution over unit vectors (experts), encoded in a density matrix $P$, that updates after each round. At the end of the round, Bob pays to Alice $\trace{P M}$. The claim is that it is possible to upper bound Bob's payoff if he had selected a fixed unit vector a-priori with Bob's average payoff. In other words, the latter claim states that it is possible to bound the maximum eigenvalue of the sum of Alice's selected matrices.
These two different matrix games seems to be applicable only on different scenarios, although their proofs draw many similarities. Combining the ideas of Arora-Kale's result and few observations from the current work it is possible to extend Arora-Kale's framework to the set of non-square matrices. However, we do not have any interesting implications of such a result.
}
\vspace*{-3.0ex}
\section{Alon-Roichman Expanding Cayley Graphs}\label{sec:AR_graphs}
We start by describing expander graphs. Given a connected undirected $d$-regular graph $H=(V,E)$ on $n$ vertices, let $A$ be its adjacency matrix, i.e., $A_{ij}=w_{ij}$ where $w_{ij}$ is the number of edges between vertices $i$ and $j$. Moreover, let $\widehat{A}=\frac1{d}A$ be its normalized adjacency matrix. We allow self-loops and multiple edges. Let $\lambda_1(\widehat{A}),\ldots ,\lambda_n(\widehat{A})$ be its eigenvalues in decreasing order. We have that $\lambda_1(\widehat{A})=1$ with corresponding eigenvector $\mathbf{1}/\sqrt{n}$, where $\mathbf{1}$ is the all-one vector. The graph $H$ is called a spectral expander if $\lambda(\widehat{A}):=\max_{2\leq j}\{ |\lambda_j(\widehat{A})|\}\leq \varepsilon$ for some positive constant $\varepsilon<1$.
Denote by $m_k=m_k(H):= \trace{A^k}$. By definition, $m_k$ is equal to the number of self-returning walks of length $k$ of the graph $H$. A graph-spectrum-based invariant, recently proposed by Estrada is defined as $EE(A) := \trace{\expm{A}}$~\cite{estrada}, which also equals to $\sum_{k=0}^{\infty} m_k/k!$. For $\theta>0$, we define the \emph{even $\theta$-Estrada index} by $EE_{\text{even}}(A,\theta) := \sum_{k=0}^{\infty} m_{2k}(\theta A)/(2k)!$.
Now let $G$ be any finite group of order $n$ with identity element $\mathtt{id}$. Let $S$ be a multi-set of elements of $G$, we denote by $S\sqcup S^{-1}$ the symmetric closure of $S$, namely the number of occurrences of $s$ and $s^{-1}$ in $S\sqcup S^{-1}$ equals the number of occurrences of $s\in S$. Let $R$ be the right regular representation\footnote{In other words, represent each group algebra element with a permutation matrix of size $n$ that preserves the group structure. This is always possible due to Cayley's theorem.}, i.e., $(R(g_1)\phi)(g_2) = \phi(g_1 g_2)$ for every $\phi : G \to \mathbb{R}$ and $g_1,g_2\in G$. The Cayley graph $\Cay{G}{S}$ on a group $G$ with respect to the mutli-set $S\subset G$ is the graph whose vertex set is $G$, and where $g_1$ and $g_2$ are connected by an edge
if there exists $s\in S$ such that $g_2 = g_1 s$ (allowing multiple edges for multiple elements in $S$).
In this section we prove the correctness of the following greedy algorithm for constructing expanding Cayley graphs.
\begin{theorem}\label{thm:AR_graphs}
Algorithm~\ref{alg:estradaAR}, given the multiplication table of a finite group $G$ of size $n$ and $0<\varepsilon<1$, outputs a (symmetric) multi-set $S\subset G$ of size $\mathcal{O}(\log n /\varepsilon^2)$ such that $\lambda (\Cay{G}{S}) \leq \varepsilon$ in $\mathcal{O}(n^2\log^3 n /\varepsilon^3)$ time. Moreover, the algorithm performs only group algebra operations that correspond to counting closed paths in Cayley graphs.
\end{theorem}
\vspace*{-3.5ex}
\begin{algorithm}{}
\caption{Expander Cayley Graph via even Estrada Index Minimization}\label{alg:estradaAR}
\begin{algorithmic}[1]
\Procedure{GreedyEstradaMin}{$G$, $\varepsilon$}\Comment{Multiplication table of $G$, $0<\varepsilon <1$}
\State Set $S^{(0)}=\emptyset$ and $t=\mathcal{O}(\log n /\varepsilon^2)$
\For {$i=1,\ldots t$ }
\State Let $g_{*}\in G$ that (approximately) min. the even $\varepsilon/ 2$-Estrada index of $\Cay{G}{S^{(i-1)}\cup g \cup g^{-1}}$ over all $g\in G $ \Comment{Use Lemma~\ref{lem:fastEstrada:Cayley}}
\State Set $S^{(i)} = S^{(i-1)} \cup g_{*} \cup g_{*}^{-1}$
\EndFor
\State \textbf{Output:} A multi-set $S:=S^{(t)}$ of size $2t$ such that $\lambda(\Cay{G}{S}) \leq \varepsilon$
\EndProcedure
\end{algorithmic}
\end{algorithm}
\vspace*{-4.0ex}
Let $\widehat{A}$ be the normalized adjacency matrix of $\Cay{G}{S\sqcup S^{-1}}$ for some $S\subset G$. It is not hard to see that $ \widehat{A} = \frac1{2|S|} \sum_{s\in S}{ (R(s) + R(s^{-1}))}$.
We want to bound $\lambda (A)$. Notice that $\lambda(A)=\norm{(\mathbf{I} - \mathbf{J}/n)A}$. Since we want to analyze the second-largest eigenvalue (in absolute value), we consider $(\mathbf{I} - \mathbf{J}/n)A = \frac1{|S|} \sum_{s\in S}{ (R(s) + R(s^{-1})) /2} - \mathbf{J}/n.$
Based on the above calculation, we define our matrix-valued function as
\begin{equation}\label{eq:AR:samplesnew}
f(g) := (R(g) + R(g^{-1})) / 2 - \mathbf{J} /n
\end{equation}
for every $g\in G$. The following lemma connects the potential function that is used in Theorem~\ref{thm:hypercosine:main} and the even Estrada index.
\begin{lemma}\label{lem:cosh_Estrada}
Let $S\subset G $ and $A$ be the adjacency matrix of $\Cay{G}{S\sqcup S^{-1}}$. For any $\theta>0$, $\trace{ \coshm{ \theta \sum_{s\in S} f(s) } } = EE_{even} (A,\theta/2) + 1 - \cosh(\theta |S|).$
\end{lemma}
The following lemma indicates that it is possible to efficiently compute the (even) Estrada index for Cayley graphs with small generating set.
\begin{lemma}\label{lem:fastEstrada:Cayley}
Let $S\subset G $, $\theta,\delta >0$, and $A$ be the adjacency matrix of $\Cay{G}{S}$. There is an algorithm that, given $S$, computes an additive $\delta$ approximation to $EE(\theta A)$ or $EE_{\text{even}}(A,\theta)$ in $\mathcal{O}(n|S| \max\{ \log (n/\delta) , 2\ensuremath{{\rm e}}^2 |S| \theta \})$ time.
\end{lemma}
\begin{proof}(of Theorem~\ref{thm:AR_graphs})
By Lemma~\ref{lem:cosh_Estrada}, minimizing the even $\varepsilon/2$-Estrada index in the $i$-th iteration is equivalent to minimizing $\trace{ \coshm{ \theta \sum_{s\in S^{(i-1)}} f(s) +\theta f(g) } }$ over all $g\in G$ with $\theta = \varepsilon$. Notice that $f(g)\in \mathcal{S}^{n\times n}$ for $g\in G$, $\ensuremath{\mathbb{E}}_{g\in_R{G}}{f(g)} = \mathbf{0}_n$ since $\sum_{g\in G}R(g) = \mathbf{J}$. It is easy to see that $\norm{f(g)} \leq 2$ and moreover a calculation implies that $\norm{\ensuremath{\mathbb{E}}_{g\in_R{G}}{f(g)^2}} \leq 2$ as well. Theorem~\ref{thm:hypercosine:main} implies that we get a multi-set $S$ of size $t$ such that $\lambda (\Cay{G}{ S\sqcup S^{-1}})=\norm{\frac1{|S|} \sum_{s\in S} f(s) } \leq \varepsilon$. The moreover part follows from Lemma~\ref{lem:fastEstrada:Cayley} with $\delta = \frac{\ensuremath{{\rm e}}^{\varepsilon^2}}{n^c}$ for a sufficient large constant $c>0$. Indeed, in total we incur (following the proof of Theorem~\ref{thm:hypercosine:main}) at most an additive $\ln( \delta n \ensuremath{{\rm e}}^{\varepsilon^2 t}) / \varepsilon$ error which is bounded by $\varepsilon$.
\end{proof}
\section{Fast Isotropic Sparsification and Spectral Sparsification}\label{sec:fast_isotrop_sparse}
Let $A$ be an $m\times n$ matrix with $m\gg n$ whose columns are in isotropic position, i.e., $A^\top A = \mathbf{I}_n$. For $0 <\varepsilon < 1$, consider the problem of finding a small subset of (rescaled) rows of $A$ forming a matrix $\widetilde{A}$ such that $\norm{\widetilde{A}^\top \widetilde{A} - \mathbf{I} } \leq \varepsilon$. The matrix Bernstein inequality (see~\cite{chernoff:matrix_valued:Tropp}) tells us that there exists such a set with size $\mathcal{O}(n\log n /\varepsilon^2)$. Indeed, set $f(i)=A_{(i)} \otimes A_{(i)} / p_i - \mathbf{I}_n$ where $p_i = \norm{A_{(i)}}^2 / \frobnorm{A}^2$. A calculation shows that $\gamma$ and $\rho^2$ are $\mathcal{O}(n)$. Moreover, Algorithm~\ref{alg:derand:Bernstein} implies an $\mathcal{O}(mn^4 \log n /\varepsilon^2)$ time algorithm for finding such a set. The running time of Algorithm~\ref{alg:derand:Bernstein} for rank-one matrix samples can be improved to $\mathcal{O}(mn^3 \polylog{n} /\varepsilon^2)$ by exploiting their rank-one structure. More precisely, using fast algorithms for computing all the eigenvalues of matrices after rank-one updates~\cite{Gu:update}. Next we show that we can further improve the running time by a more careful analysis.
We show how to improve the running time of Algorithm~\ref{alg:derand:Bernstein} to $\mathcal{O}(\frac{mn^2}{\varepsilon^2} \polylog{n, \frac1{\varepsilon}})$ utilizing results from numerical linear algebra including the Fast Multipole Method~\cite{FMM:CGR} (FMM) and ideas from~\cite{Gu:update}. The main idea behind the improvement is that the trace is invariant under any change of basis. At each iteration, we perform a change of basis so that the matrix corresponding to the previous choices of the algorithm is diagonal. Now, Step $4$ of Algorithm~\ref{alg:derand:Bernstein} corresponds to computing all the eigenvalues of $m$ different eigensystems with special structure, i.e., diagonal plus a rank-one matrix. Such eigensystem can be solved in $\mathcal{O}(n \polylog{n})$ time using the FMM as was observed in~\cite{Gu:update}. However, the problem now, is that at each iteration we have to represent all the vectors $A_{(i)}$ in the new basis, which may cost $\mathcal{O}(mn^2)$. The key observation is that the change of basis matrix at each iteration is a Cauchy matrix (see Appendix). It is known that matrix-vector multiplication with Cauchy matrices can be performed efficiently and numerically stable using FMM. Therefore, at each iteration, we can perform the change of basis in $\mathcal{O}(mn\polylog{n})$ and $m$ eigenvalue computations in $\mathcal{O}(mn\polylog{n})$ time. The next theorem states that the resulting algorithm runs in $\mathcal{O}(mn^2 \polylog{n})$ time (see Appendix for proof).
\begin{theorem}\label{thm:derand:isotrop:fast}
Let $A$ be an $m\times n$ matrix with $A^\top A = \mathbf{I}_n$, $m\geq n$ and $ 0 < \varepsilon <1$. Algorithm~\ref{alg:fast:isotrop} returns at most $t=\mathcal{O}(n \ln n/\varepsilon^2)$ indices $x_1^*,x_2^*,\ldots x_t^*$ over $[m]$ with corresponding scalars $s_1,s_2,\ldots ,s_t$ using $\widetilde{\mathcal{O}}(mn^2 \log^3 n /\varepsilon^2 )$ operations such that
\begin{equation}\label{eq:main_thm:fast:main_eqn}
\norm{ \sum_{i=1}^{t} s_i A_{(x_i^*)} \otimes A_{(x_i^*)} - \mathbf{I}_n} \leq \varepsilon.
\end{equation}
\end{theorem}
\vspace*{-3.5ex}
\begin{algorithm}{}
\caption{Fast Isotropic Sparsification}\label{alg:fast:isotrop}
\begin{algorithmic}[1]
\Procedure{Isotrop}{$A$, $\varepsilon$} \Comment{$A\in{\mathbb{R}^{m\times n}}$, $\sum_{k=1}^{m}A_{(k)}\otimes A_{(k)} = \mathbf{I}_n$ and $0 < \varepsilon <1$}
\State Set $\theta = \varepsilon / n $, $t=\mathcal{O}( n \ln n/\varepsilon^2)$, and $A_{(k)} \leftarrow A_{(k)}/\sqrt{p_k}$ for every $k\in{[m]}$, where $p_k=\norm{A_{(k)}}^2/n$
\State Set $\Lambda_{\{0\}} = \mathbf{0}_n$ and $Z = \sqrt{\theta}\ A$
\For {$i=1$ to $t$}
\State $x_i^* = \operatorname*{arg\; min}_{k\in{[m]}}{\trace{\expm{ \Lambda_{\{i - 1\}} + Z_{(k)} \otimes Z_{(k)} }\ensuremath{{\rm e}}^{-\theta i} + \expm{- \Lambda_{\{i - 1\}} - Z_{(k)} \otimes Z_{(k)} }\ensuremath{{\rm e}}^{\theta i} } } $ \Comment{Apply $m$ times Lemma~\ref{lem:comp_eigs}}
\State $[\Lambda_{\{i\}}, U_{\{i\}}] = \textbf{eigs} ( \Lambda_{\{i - 1\}} + Z_{(x_i^*)} \otimes Z_{(x_i^*)} )$ \Comment{\textbf{eigs} computes eigensystem}
\State $Z = Z U_{\{i\}} $ \Comment{Apply fast matrix-vector multiplication }
\EndFor
\State \textbf{Output:} $t$ indices $x_1^*, x_2^*, \ldots ,x_t^*,\ x_i^* \in{[m]}$ s.t. $\norm{ \sum_{k=1}^{t} \frac{A_{(x_k^*)} \otimes A_{(x_k^*)} }{ tp_{x_k^*}} - \mathbf{I}_n } \leq \varepsilon $
\EndProcedure
\end{algorithmic}
\end{algorithm}
\vspace*{-4.0ex}
Next, we show that Algorithm~\ref{alg:fast:isotrop} can be used as a bootstrapping procedure to improve the time complexity of~\cite[Theorem~3.1]{phdthesis:Srivastava:2010}, see also~\cite[Theorem~$3.1$]{graph:sparsifiers:twice_ram}. Such an improvement implies faster algorithms for constructing graph sparsifiers and, as we will see in \S~\ref{sec:sparsification:matrix}, element-wise sparsification of matrices.
\begin{theorem}\label{thm:sparsification:here}
Suppose $0 < \varepsilon < 1$ and $A = \sum_{i=1}^{m} v_i \otimes v_i $ are given, with column vectors $v_i\in\mathbb{R}^n $ and $m\geq n$. Then there are non-negative weights $\{s_i\}_{i\leq m}$, at most $ \lceil n/\varepsilon^2 \rceil$ of which are non-zero, such that
\begin{equation}
(1-\varepsilon)^3 A \preceq \widetilde{A} \preceq (1+\varepsilon)^3 A,
\end{equation}
where $\widetilde{A} = \sum_{i=1}^{m}s_i v_i \otimes v_i$. Moreover, there is an algorithm that computes the weights $\{s_i\}_{i\leq m}$ in deterministic $\widetilde{\mathcal{O}}(mn^2 \log^3 n /\varepsilon^2 + n^4 \log n /\varepsilon^4)$ time.
\end{theorem}
\section{Element-wise Matrix Sparsification}\label{sec:sparsification:matrix}
A deterministic algorithm for the element-wise matrix sparsification problem can be obtained by derandomizing a recent result whose analysis is based on the matrix Bernstein inequality~\cite{matrix:sparsification:IPL2011}.
\begin{theorem}\label{thm:matrix_sparse:slow}
Let $A$ be an $n\times n$ matrix and $ 0 < \varepsilon <1$. There is a deterministic polynomial time algorithm that, given $A$ and $ \varepsilon$, outputs a matrix $\widetilde{A}\in \mathbb{R}^{n\times n}$ with at most $ 28 n \ln (\sqrt{2n} ) \sr{A} /\varepsilon^2$ non-zero entries such that $\norm{A - \widetilde{A}} \leq \varepsilon \norm{A}.$
\end{theorem}
Next, we give two improved element-wise sparsification algorithms for self-adjoint and diagonally dominant-like matrices; one of them is randomized and the other is its derandomized version. Both algorithms share a crucial difference with all previously known algorithms for this problem; during their execution they may densify the diagonal entries of the input matrices. On the one hand, there are at most $n$ diagonal entries, so this does not affect asymptotically their sparsity guarantees. On the other hand, as we will see later this twist turns out to give strong sparsification bounds.
Recall that the results of~\cite{graph:sparsifiers:eff_resistance,graph:sparsifiers:twice_ram} imply an element-wise sparsification algorithm that works only for Laplacian matrices. It is easy to verify that Laplacian matrices are also diagonally dominant. Here we extend these results to a wider class of matrices (with a weaker notion of approximation). The diagonally dominant assumption is too restrictive and we will show that our sparsification algorithms work for a wider class of matrices. To accommodate this, we say that a matrix $A$ is $\theta$-symmetric diagonally dominant (abbreviate by $\theta$-SDD) if $A$ is self-adjoint and the inequality $\infnorm{A} \leq \sqrt{\theta} \norm{A}$ holds.
By definition, any diagonally dominant matrix is also a $4$-SDD matrix. On the other extreme, every self-adjoint matrix of size $n$ is $n$-SDD since the inequality $\infnorm{A}\leq \sqrt{n} \norm{A}$ is always valid. The following elementary lemma gives a connection between element-wise matrix sparsification and spectral sparsification as defined in~\cite{phdthesis:Srivastava:2010}.
\begin{lemma}\label{lem:sparsif:decomp}
Let $A$ be a self-adjoint matrix of size $n$ and $R=\diag{R_1,R_2,\ldots ,R_n}$ where $R_i = \sum_{j\neq i} |A_{ij}|$. Then there is a matrix $C$ of size $n\times m$ with $m \leq \binom{n}{2}$ such that
\begin{eqnarray}\label{eqn:sparsify_lemma}
A = CC^\top +\diag{A} - R.
\end{eqnarray}
Moreover, each column of $C$ is indexed by the ordered pairs $(i,j)$, $i<j$ and equals to $C^{(i,j)} = \sqrt{|A_{ij}|} e_i + \sign{A_{ij}} \sqrt{|A_{ij}|} e_j$ for every $i<j$, $i,j\in[n]$.
\end{lemma}
\begin{remark}
In the special case where $A$ is the Laplacian matrix of some graph, the above decomposition is precisely the vertex-edge decomposition of the Laplacian matrix, since in this case $\diag{A} =R$.
\end{remark}
Using the above lemma, we give a randomized and a deterministic algorithm for sparsifying $\theta$-SDD matrices. First we present the randomized algorithm.
\begin{theorem}\label{thm:matrix_sparsif:rand}
Let $A$ be a $\theta$-SDD matrix of size $n$ and $ 0 < \varepsilon <1$. There is a randomized linear time algorithm that, given $A$, $\norm{A}$ and $\varepsilon$, outputs a matrix $\widetilde{A}\in \mathbb{R}^{n\times n}$ with at most $\mathcal{O}( n\theta \log n /\varepsilon^2)$ non-zero entries such that w.p. at least $1-1/n$, $\norm{A - \widetilde{A}} \leq \varepsilon \norm{A}.$
\end{theorem}
Next we state the derandomized algorithm of the above result.
\begin{theorem}\label{thm:matrix_sparsif:det}
Let $A$ be a $\theta$-SDD matrix of size $n$ and $ 0 < \varepsilon <1/2$. There is an algorithm that, given $A$ and $ \varepsilon$, outputs a matrix $\widetilde{A}\in \mathbb{R}^{n\times n}$ with at most $\mathcal{O}( n \theta /\varepsilon^2)$ non-zero entries such that $\norm{A - \widetilde{A}} \leq \varepsilon \norm{A}$. Moreover, the algorithm computes $\widetilde{A}$ in deterministic $\widetilde{\mathcal{O}}(\nnz{A}n^2 \theta\log^3 n /\varepsilon^2 + n^4 \theta^2 \log n /\varepsilon^4)$ time.
\end{theorem}
\begin{remark}
The results of~\cite{graph:sparsifiers:twice_ram,phdthesis:Srivastava:2010} imply a deterministic $\mathcal{O}(\nnz{A} \theta n^3 /\varepsilon^2 )$ time algorithm that outputs a matrix $\widetilde{A}$ with at most $ \lceil 19(1+\sqrt{\theta})^2 /\varepsilon^2\rceil n $ non-zero entries such that $\norm{\widetilde{A}-A} \leq \varepsilon\norm{A}$.
\end{remark}
\subsection*{Acknowledgements}
The author would like to thank Mark Braverman for several interesting discussions and comments about this work.
{
|
\section{Introduction}
\label{sec:intro}
General relativity is a theory in which the construction of
exact solutions is not so easy.
In this situation, perturbation theories are powerful techniques
and the developments of perturbation theories lead physically
fruitful results and interpretations of natural phenomena.
For this reason, general relativistic {\it linear} perturbation
theory has been widely used in many area\cite{Bardeen-1980}.
Further, the investigation of {\it higher-order}
general-relativistic perturbations is also necessary due to the
precise observations\cite{Non-Gaussianity-observation} in
cosmology\cite{Tomita-1967-Non-Gaussianity,kouchan-cosmo-second},
and to prepare more precise wave form of gravitational
wave\cite{Gleiser-Nicasio} for the gravitational wave
detection.
As well-known, general relativity is based on the concept of
general covariance.
Due to this general covariance, the ``gauge degree of freedom'',
which is an unphysical degree of freedom, arises in
general-relativistic perturbations.
To obtain physical results, we have to fix this gauge degrees of
freedom or to extract some invariant quantities of
perturbations.
This situation becomes more complicated in higher-order
perturbations.
Since there are so-called {\it gauge-invariant} linear
perturbation theories in some background spacetimes, it is
worthwhile to investigate higher-order gauge-invariant
perturbation theory from a general point of view to avoid this
gauge issues.
According to these motivation, the general framework of
higher-order general-relativistic gauge-invariant perturbation
theory has been discussed in some papers by the present
author\cite{kouchan-gauge-inv}.
In this general framework, we consider the perturbative
expansion of the physical metric
$\bar{g}_{ab}=g_{ab}+\lambda h_{ab}+O(\lambda^{2})$, where
$\lambda$ is an infinitesimal parameter for the perturbation
theory.
$h_{ab}$ is the linear-order metric perturbation.
Further, in our general framework, we assumed the following
conjecture:
\begin{conjecture}
\label{conjecture:decomposition-conjecture}
If there is a tensor field $h_{ab}$ of the second rank, whose
gauge transformation rule is
\begin{eqnarray}
{}_{{\cal Y}}\!h_{ab}
-
{}_{{\cal X}}\!h_{ab}
=
{\pounds}_{\xi}g_{ab},
\label{eq:linear-metric-gauge-trans}
\end{eqnarray}
then there exist a tensor field ${\cal H}_{ab}$ and a vector
field $X^{a}$ such that $h_{ab}$ is decomposed as
\begin{eqnarray}
h_{ab} =: {\cal H}_{ab} + {\pounds}_{X}g_{ab},
\label{eq:linear-metric-decomp}
\end{eqnarray}
where ${\cal H}_{ab}$ and $X^{a}$ are transformed as
\begin{equation}
{}_{{\cal Y}}\!{\cal H}_{ab} - {}_{{\cal X}}\!{\cal H}_{ab} = 0,
\quad
{}_{\quad{\cal Y}}\!X^{a} - {}_{{\cal X}}\!X^{a} = \xi^{a}.
\label{eq:linear-metric-decomp-gauge-trans}
\end{equation}
Here, ${\cal Y}$ and ${\cal X}$ denote different gauge choices.
\end{conjecture}
In the case of cosmological perturbations, we confirmed that
this conjecture is almost true, and then, we developed the
second-order gauge-invariant cosmological
perturbations\cite{kouchan-cosmo-second}.
Since our general framework of higher-order perturbations does
not depend on details of the background metric $g_{ab}$, we will
be able to develop general-relativistic higher-order
gauge-invariant perturbation theory if Conjecture
\ref{conjecture:decomposition-conjecture} is
true\cite{kouchan-gauge-inv}.
Although we recently propose a scenario of the proof of
Conjecture \ref{conjecture:decomposition-conjecture}, in this
article, we give an alternative explanation of this proof.
This is the main purpose of this article.
\section{Construction of gauge-invariant variables}
\label{sec:construction}
Now, we give a scenario of a proof of Conjecture
\ref{conjecture:decomposition-conjecture} on general background
spacetimes.
Here, we assume that the background spacetimes considered in
this paper admit ADM decomposition\cite{Wald-book}.
Therefore, the background spacetime ${\cal M}_{0}$ considered
here is $n+1$-dimensional spacetime which is foliated by
spacelike hypersurfaces $\Sigma(t)$ ($\dim\Sigma = n$).
Each $\Sigma$ may have its boundary $\partial\Sigma$.
The background metric $g_{ab}$ is given by
\begin{eqnarray}
\label{eq:gdb-decomp-dd-minus-main}
g_{ab} &=& - \alpha^{2} (dt)_{a} (dt)_{b}
+ q_{ij}
(dx^{i} + \beta^{i}dt)_{a}
(dx^{j} + \beta^{j}dt)_{b},
\end{eqnarray}
where $\alpha$ is the lapse function, $\beta^{i}$ is the
shift vector, and $q_{ab}=q_{ij}(dx^{i})_{a}(dx^{i})_{b}$ is the
metric on $\Sigma(t)$.
To consider the decomposition (\ref{eq:linear-metric-decomp}) of
$h_{ab}$, we, first, consider the components of the metric
$h_{ab}$ as
\begin{eqnarray}
h_{ab}
=
h_{tt} (dt)_{a}(dt)_{b}
+ 2 h_{ti} (dt)_{(a}(dx^{i})_{b)}
+ h_{ij} (dx^{i})_{a}(dx^{j})_{b}.
\label{eq:K.Nakamura-2010-note-1-2}
\end{eqnarray}
The gauge transformation rule
(\ref{eq:linear-metric-gauge-trans}) gives the transformation
rules for the components $\{h_{tt},h_{ti},h_{ij}\}$, which are
given by
\begin{eqnarray}
{}_{{\cal Y}}h_{tt}
-
{}_{{\cal X}}h_{tt}
&=&
2 \partial_{t}\xi_{t}
- \frac{2}{\alpha}\left(
\partial_{t}\alpha
+ \beta^{i}D_{i}\alpha
- \beta^{j}\beta^{i}K_{ij}
\right) \xi_{t}
\nonumber\\
&&
- \frac{2}{\alpha} \left(\frac{}{}
\beta^{i}\beta^{k}\beta^{j} K_{kj}
- \beta^{i} \partial_{t}\alpha
+ \alpha q^{ij} \partial_{t}\beta_{j}
\right.
\nonumber\\
&& \quad\quad\quad
\left.
+ \alpha^{2} D^{i}\alpha
- \alpha \beta^{k} D^{i} \beta_{k}
- \beta^{i} \beta^{j} D_{j}\alpha
\frac{}{}
\right)\xi_{i}
\label{eq:K.Nakamura-2010-note-3-2}
, \\
{}_{{\cal Y}}h_{ti}
-
{}_{{\cal X}}h_{ti}
&=&
\partial_{t}\xi_{i}
+ D_{i}\xi_{t}
- \frac{2}{\alpha} \left(
D_{i}\alpha
- \beta^{j}K_{ij}
\right) \xi_{t}
\nonumber\\
&&
- \frac{2}{\alpha} \left(
- \alpha^{2} K^{j}_{\;\;i}
+ \beta^{j}\beta^{k} K_{ki}
- \beta^{j} D_{i}\alpha
+ \alpha D_{i}\beta^{j}
\right) \xi_{j}
\label{eq:K.Nakamura-2010-note-4-2}
, \\
{}_{{\cal Y}}h_{ij}
-
{}_{{\cal X}}h_{ij}
&=&
2 D_{(i}\xi_{j)}
+ \frac{2}{\alpha} K_{ij} \xi_{t}
- \frac{2}{\alpha} \beta^{k} K_{ij} \xi_{k}
,
\label{eq:K.Nakamura-2010-note-5-2}
\end{eqnarray}
where $K_{ij}$ is the extrinsic curvature:
\begin{eqnarray}
\label{eq:K.Nakamura-2010-2-generic-alternative-logic-2.21}
K_{ij} = - \frac{1}{2\alpha}
\left[\frac{\partial}{\partial t} q_{ij} -
D_{i}\beta_{j} - D_{j}\beta_{i}\right],
\end{eqnarray}
and $D_{i}$ is the covariant derivative associated with the
spatial metric $q_{ij}$.
Inspecting these gauge transformation rules and assuming the
existence of the Green function of the derivative operator
$\Delta:=D^{i}D_{i}$, we consider the decompositions of the
components $h_{ti}$ and $h_{ij}$ as follows:
\begin{eqnarray}
h_{ti}
&=:&
D_{i}h_{(VL)} + h_{(V)i}
- \frac{2}{\alpha} \left(
D_{i}\alpha
- \beta^{j}K_{ij}
\right) h_{(VL)}
\nonumber\\
&&
- \frac{2}{\alpha} \left(
D_{i}\alpha
- \beta^{j}K_{ij}
\right)
\Delta^{-1}
\left[
\partial_{t}D^{k}h_{(TV)k}
- 2 D_{l}\left\{
h_{(TV)k} \left( \alpha K^{kl} - D^{(k}\beta^{l)} \right)
\right\}
\right.
\nonumber\\
&& \quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad
\left.
+ h_{(TV)k} D^{k}\left( \alpha K - D^{l}\beta_{l} \right)
\frac{}{}
\right]
\nonumber\\
&&
- \frac{2}{\alpha} \left(
- \alpha^{2} K^{j}_{\;\;i}
+ \beta^{j}\beta^{k} K_{ki}
- \beta^{j} D_{i}\alpha
+ \alpha D_{i}\beta^{j}
\right) h_{(TV)j}
\label{eq:K.Nakamura-2010-2-generic-alternative-logic-2.27}
, \\
h_{ij}
&=:&
\frac{1}{n} q_{ij} h_{(L)} + h_{(T)ij}
+ \frac{2}{\alpha} K_{ij} h_{(VL)}
- \frac{2}{\alpha} \beta^{k} K_{ij} h_{(TV)k}
\nonumber\\
&&
- \frac{2}{\alpha} K_{ij} \Delta^{-1}
\left[\frac{}{}
\partial_{t}D^{k}h_{(TV)k}
- 2 D_{l}\left\{
h_{(TV)k} \left( \alpha K^{kl} - D^{(k}\beta^{l)} \right)
\right\}
\right.
\nonumber\\
&& \quad\quad\quad\quad\quad\quad
\left.
+ h_{(TV)k} D^{k}\left( \alpha K - D^{l}\beta_{l} \right)
\frac{}{}
\right]
,
\label{eq:K.Nakamura-2010-2-generic-alternative-logic-2.29}
\\
h_{(T)ij} &=:& D_{i}h_{(TV)j} + D_{j}h_{(TV)i}
- \frac{2}{n} q_{ij} D^{l}h_{(TV)l}
+ h_{(TT)ij},
\label{eq:K.Nakamura-2010-2-generic-alternative-logic-2.30}
\\
&& D^{i}h_{(V)i} = 0, \quad
q^{ij}h_{(TT)ij} = 0, \quad
D^{i}h_{(TT)ij} = 0.
\label{eq:K.Nakamura-2010-2-generic-alternative-logic-2.28}
\end{eqnarray}
From the gauge-transformation rule
(\ref{eq:K.Nakamura-2010-note-3-2}) and the definitions
(\ref{eq:K.Nakamura-2010-2-generic-alternative-logic-2.27}) and
(\ref{eq:K.Nakamura-2010-2-generic-alternative-logic-2.28}) of
the variable $h_{(LV)}$ and $h_{(V)i}$, we can derive the
gauge-transformation rules for these variables by assuming the
existence of the Green function ${\cal F}^{-1}$ of the elliptic
derivative operator
\begin{eqnarray}
\label{eq:K.Nakamura-2010-2-generic-alternative-logic-2.36}
{\cal F}
:=
\Delta
- \frac{2}{\alpha} \left(
D_{i}\alpha
- \beta^{j}K_{ij}
\right)D^{i}
- 2 D^{i}\left\{
\frac{1}{\alpha} \left(
D_{i}\alpha
- \beta^{j}K_{ij}
\right)
\right\}.
\end{eqnarray}
However, in these gauge-transformation rules, the variable
$h_{(TV)i}$ defined in
Eqs.~(\ref{eq:K.Nakamura-2010-2-generic-alternative-logic-2.29})--(\ref{eq:K.Nakamura-2010-2-generic-alternative-logic-2.28})
is also included.
This means that the gauge-transformation rules for variables
$h_{(LV)}$ and $h_{(V)i}$ are specified if we have the
gauge-transformation rule for $h_{(TV)i}$.
From the trace part of the gauge-transformation rule
(\ref{eq:K.Nakamura-2010-note-5-2}), we can derive the
gauge-transformation rule for the variable $h_{(L)}$, which is
defined by
Eqs.~(\ref{eq:K.Nakamura-2010-2-generic-alternative-logic-2.29})--(\ref{eq:K.Nakamura-2010-2-generic-alternative-logic-2.28}).
This gauge-transformation rule also includes the variables
$h_{(LV)}$ and $h_{(TV)i}$.
Since the gauge-transformation rule for the variable $h_{(LV)}$
is determined if the gauge-transformation rule for the variable
$h_{(TV)i}$ is specified, the gauge-transformation rule for
$h_{(L)}$ is also specified if the gauge-transformation rule for
$h_{(TV)i}$ is given.
Thus, the gauge-transformation rules for the variables
$h_{(LV)}$, $h_{(V)i}$, and $h_{(L)}$ are specified if the
gauge-transformation for the variable $h_{(TV)i}$ is specified.
Taking the divergence of the traceless part of the
gauge-transformation rule (\ref{eq:K.Nakamura-2010-note-5-2})
and using
Eqs.~(\ref{eq:K.Nakamura-2010-2-generic-alternative-logic-2.29})--(\ref{eq:K.Nakamura-2010-2-generic-alternative-logic-2.28})
and the gauge-trasformation rule for the variable $h_{(LV)}$
obtained above, we reach to a single equation for a unknown
single vector field
$A_{i}:={}_{{\cal Y}}h_{(TV)i}-{}_{{\cal X}}h_{(TV)i}-\xi_{i}$:
\begin{eqnarray}
{\cal D}^{jl} A_{l}
&=& D_{i}\left[
\frac{2}{\alpha} \widetilde{K}^{ij}
\left\{\frac{}{}
{\cal F}^{-1}\left[\frac{}{}
\partial_{t}D^{k}A_{k}
- 2 D_{l}\left\{
\frac{}{}
A_{k}
\left( \alpha K^{kl} - D^{(k}\beta^{l)} \right)
+
\frac{1}{\alpha} L^{ml} A_{m}
\right\}
\right.
\right.
\right.
\nonumber\\
&& \quad\quad\quad\quad\quad\quad\quad\quad
\left.
\left.
\left.
+ A_{k} D^{k}\left( \alpha K - D^{l}\beta_{l} \right)
\frac{}{}
\right]
\right.
\right.
\nonumber\\
&& \quad\quad\quad\quad\quad
\left.
\left.
- \beta^{k}A_{k}
\frac{}{}
\right\}
\right]
,
\label{eq:Ai-master-eq}
\end{eqnarray}
where the derivative operator ${\cal D}^{ij}$ and the
tensors $\tilde{K}^{ij}$ and $L^{ij}$ are defined by
\begin{eqnarray}
{\cal D}^{ij}
&:=&
q^{ij}\Delta + \left(1 - \frac{2}{n}\right) D^{i}D^{j} +
R^{ij},
\quad
\widetilde{K}^{ij}
:=
K^{ij} - \frac{1}{n} q^{ij} K
,
\quad
K := q^{ij}K_{ij},
\nonumber\\
L^{ij}
&:=&
- \alpha^{2} K^{ij}
+ \beta^{i}\beta_{k} K^{kj}
- \beta^{i} D^{j}\alpha
+ \alpha D^{j}\beta^{i}
.
\label{eq:J.W.York.Jr-1973-1-2-kouchan-3-main}
\end{eqnarray}
As a trivial solution to Eq.~(\ref{eq:Ai-master-eq}), we have
$A_{i}=0$, i.e.,
\begin{eqnarray}
{}_{{\cal Y}}h_{(TV)i} - {}_{{\cal X}}h_{(TV)i} = \xi_{i}
\label{eq:K.Nakamura-2010-2-generic-alternative-logic-2.56}
\end{eqnarray}
Thus, we have obtained the gauge-transformation rule for the
variable $h_{(TV)i}$.
Using the gauge-transformation rule
(\ref{eq:K.Nakamura-2010-2-generic-alternative-logic-2.56}) for
the variable $h_{(TV)i}$, the gauge-transformation rules for
variables $h_{tt}$, $h_{(VL)}$, $h_{(V)i}$, $h_{(L)}$, and $h_{(TT)ij}$
are given by
\begin{eqnarray}
{}_{{\cal Y}}h_{(VL)}
-
{}_{{\cal X}}h_{(VL)}
&=&
\xi_{t}
+ \Delta^{-1} \left[
\partial_{t}D^{k}\xi_{k}
- 2 D_{l}\left\{ \xi_{k} \left( \alpha K^{kl} - D^{(k}\beta^{l)} \right) \right\}
\right.
\nonumber\\
&& \quad\quad\quad\quad\quad
\left.
+ \xi_{k} D^{k}\left( \alpha K - D^{l}\beta_{l} \right)
\frac{}{}
\right]
\label{eq:K.Nakamura-2010-2-generic-alternative-logic-2.62}
, \\
{}_{{\cal Y}}h_{(V)i}
-
{}_{{\cal X}}h_{(V)i}
&=&
\partial_{t}\xi_{i}
- D_{i}\Delta^{-1}\left[
\partial_{t}D^{k}\xi_{k}
- 2 D_{l}\left\{
\xi_{k} \left( \alpha K^{kl} - D^{(k}\beta^{l)} \right)
\right\}
\right.
\nonumber\\
&& \quad\quad\quad\quad\quad\quad\quad
\left.
+ \xi_{k} D^{k}\left( \alpha K - D^{l}\beta_{l} \right)
\frac{}{}
\right]
,
\label{eq:K.Nakamura-2010-2-generic-alternative-logic-2.63}
\\
{}_{{\cal Y}}h_{(L)}
-
{}_{{\cal X}}h_{(L)}
&=&
2 D^{i}\xi_{i}
,
\label{eq:K.Nakamura-2010-2-generic-alternative-logic-2.64}
\\
{}_{{\cal Y}}h_{(TT)ij}
-
{}_{{\cal X}}h_{(TT)ij}
&=&
0.
\label{eq:K.Nakamura-2010-2-generic-alternative-logic-2.66}
\end{eqnarray}
Inspecting the gauge-transformation rules
(\ref{eq:K.Nakamura-2010-2-generic-alternative-logic-2.56}) and
(\ref{eq:K.Nakamura-2010-2-generic-alternative-logic-2.62}), we
first construct the variables $X_{t}$ and $X_{i}$ which satisfy
the properties ${}_{{\cal Y}}X_{t}-{}_{{\cal X}}X_{t}=\xi_{t}$,
${}_{{\cal Y}}X_{i}-{}_{{\cal X}}X_{i}=\xi_{i}$, respectively.
We can easily find these variables as follows:
\begin{eqnarray}
X_{t}
&:=&
h_{(VL)}
- \Delta^{-1} \left[ \frac{}{}
\partial_{t}D^{k}h_{(TV)k}
- 2 D_{l}\left\{
h_{(TV)k} \left( \alpha K^{kl} - D^{(k}\beta^{l)} \right)
\right\}
\right.
\nonumber\\
&& \quad\quad\quad\quad\quad\quad\quad
\left.
+ h_{(TV)k} D^{k}\left( \alpha K - D^{l}\beta_{l} \right)
\frac{}{}
\right]
,
\label{eq:K.Nakamura-2010-2-generic-alternative-logic-2.70}
\\
X_{i} &:=& h_{(TV)i}.
\label{eq:K.Nakamura-2010-2-generic-alternative-logic-2.69}
\end{eqnarray}
These $X_{t}$ and $X_{i}$ satisfy the desired
gauge-transformation rules, respectively.
Through these $X_{t}$ and $X_{i}$ and the gauge-transformation
rules (\ref{eq:K.Nakamura-2010-note-3-2}),
(\ref{eq:K.Nakamura-2010-2-generic-alternative-logic-2.63})--(\ref{eq:K.Nakamura-2010-2-generic-alternative-logic-2.66}),
we can easily define gauge-invariant variables $\Phi$, $\Psi$,
$\nu_{i}$, and $\chi_{ij}$ by
\begin{eqnarray}
- 2 \Phi
&:=&
h_{tt}
- 2 \partial_{t}X_{t}
+ \frac{2}{\alpha}\left(
\partial_{t}\alpha
+ \beta^{i}D_{i}\alpha
- \beta^{j}\beta^{i}K_{ij}
\right) X_{t}
\nonumber\\
&&
+ \frac{2}{\alpha} \left( \frac{}{}
\beta^{i}\beta^{k}\beta^{j} K_{kj}
- \beta^{i} \partial_{t}\alpha
+ \alpha q^{ij} \partial_{t}\beta_{j}
+ \alpha^{2} D^{i}\alpha
\right.
\nonumber\\
&& \quad\quad\quad
\left.
- \alpha \beta^{k} D^{i} \beta_{k}
- \beta^{i} \beta^{j} D_{j}\alpha
\frac{}{}
\right) X_{i}
,
\label{eq:K.Nakamura-2010-2-generic-alternative-logic-2.81}
\\
- 2 n \Psi
&:=&
h_{(L)} - 2 D^{i}X_{i},
\label{eq:K.Nakamura-2010-2-generic-alternative-logic-2.82}
\\
\nu_{i}
&:=&
h_{(V)i}
- \partial_{t}X_{i}
\nonumber\\
&&
+ D_{i}\Delta^{-1}\left[\frac{}{}
\partial_{t}D^{k}X_{k}
- 2 D_{l}\left\{
X_{k} \left( \alpha K^{kl} - D^{(k}\beta^{l)} \right)
\right\}
\right.
\nonumber\\
&& \quad\quad\quad\quad\quad
\left.
+ X_{k} D^{k}\left( \alpha K - D^{l}\beta_{l} \right)
\frac{}{}
\right]
,
\label{eq:K.Nakamura-2010-2-generic-alternative-logic-2.83}
\\
\chi_{ij} &:=& h_{(TT)ij}.
\label{eq:K.Nakamura-2010-2-generic-alternative-logic-2.84}
\end{eqnarray}
Actually, we can easily confirm that these variables are gauge
invariant.
We also note that the variable $\nu_{i}$ and $\chi_{ij}$ satisfy
the properties $D^{i}\nu_{i}=0$, $q^{ij}\chi_{ij}=0$,
$\chi_{[ij]}=0$, and $D^{i}\chi_{ij}=0$.
Through the original components $h_{tt}$, $h_{ti}$, and
$h_{ij}$ in terms of the variables $\Phi$, $\Psi$, $\nu_{i}$,
$\chi_{ij}$, $X_{t}$, and $X_{i}$, the first-order metric
perturbation $h_{ab}$ are given in the form as
Eq.~(\ref{eq:linear-metric-decomp}) by choosing
\begin{eqnarray}
{\cal H}_{ab}
&=&
- 2 \Phi (dt)_{a}(dt)_{b}
+ 2 \nu_{i} (dt)_{(a}(dx^{i})_{b)}
+ \left(
- 2 q_{ij} \Psi + \chi_{ij}
\right) (dx^{i})_{a}(dx^{j})_{b}
\label{eq:K.Nakamura-2010-2-generic-alternative-logic-2.96}
, \\
X_{a} &=& X_{t} (dt)_{a} + X_{i}(dx^{i})_{a}
\label{eq:K.Nakamura-2010-2-generic-alternative-logic-2.98}
.
\end{eqnarray}
Thus, we have confirmed Conjecture
\ref{conjecture:decomposition-conjecture}.
\section{Summary}
\label{sec:discussions}
We have shown an alternative scenario of a proof of Conjecture
\ref{conjecture:decomposition-conjecture}.
In our previous paper\cite{kouchan-decomp-previous}, we
first assumed the existence of the gauge-variant variables
$X_{t}$ and $X_{i}$ and confirmed this existence through the
explicit construction of these variables.
Logically speaking, this is a no-trivial logic and we have
explicitly constructed gauge-invariant variables $\Phi$, $\Psi$,
$\nu_{i}$, and $\chi_{ij}$.
In this previous derivation, we assume the existence of the
Green functions of the Laplacian $\Delta$ and the elliptic
derivative operator ${\cal D}_{ij}$ in
Eqs.~(\ref{eq:J.W.York.Jr-1973-1-2-kouchan-3-main}).
Therefore, special modes which belong to the kernels of these
two derivative operators are excluded in our consideration.
To include these modes, different treatments will be necessary.
We called this problem as {\it zero-mode problem}.
In this paper, we gave an alternative construction of the
variable $X_{t}$ and $X_{i}$ to support our previous
result\cite{kouchan-decomp-previous}.
In this sense, we may say that the results in our previous
paper\cite{kouchan-decomp-previous} are correct.
However, in the approach in this article, we assumed the
existence of the Green function of the elliptic derivative
operator ${\cal F}$, which defined by
Eq.~(\ref{eq:K.Nakamura-2010-2-generic-alternative-logic-2.36}),
instead of ${\cal D}_{ij}$.
Further, the role fo nontrivial solutions to
Eq.~(\ref{eq:Ai-master-eq}) is not clear.
In this sense, the set up of the above ``zero-mode problem'' is
still ambiguous in general case.
To clarify this ``zero-mode problem'' itself, it will be
necessary to discuss this problem through the concrete
background metric $g_{ab}$ and appropriate boundary conditions
at $\partial\Sigma$ at first.
We will leave these issues as future works.
\section*{Acknowledgments}
The author deeply thanks to Prof. Masa-katsu Fujimoto for
various supports.
|
\section{Introduction}
There is tremendous scientific and technological interest in physical systems of
reduced dimensionality.~\cite{Bowler04} Nanowires down to single atom chains are
of particular interest. The Fermi-liquid theory describing conventional bulk
conductors breaks down in such one-dimensional (1D) systems, which are predicted
to behave remarkably differently. The concept of a single particle becomes
obsolete and one has to switch to a description of collective spin and charge
excitation.~\cite{Giamarchi2004} However, the unusual properties of the
non-Fermi liquid counterpart, the Tomanaga-Luttinger liquid,~\cite{Luttinger65}
remain largely untested by experiment due to the lack of suitable practical
models for electronic systems.
The experimental quest for exotic quantum phenomena in reduced dimensions such
as quantized conductance,~\cite{Agrait03} charge density waves~\cite{Gruener94}
and Wigner crystals,~\cite{Schulz93} Peierls distortions~\cite{Peierls01} and
spin-charge separation~\cite{Zacher98} has triggered numerous developments in
the fabrication of nanostructures over the last decade.~\cite{Nitzan03,Barth05}
Optical and electron-beam lithography as well as nanoimprint techniques all
suffer from physical limitations and low fidelity on nanometer length scales.
Scanning probes enable the fabrication of nanostructures down to single-atom
assemblies.~\cite{Lyding94,Foelsch04,Hirjibehedin2006} However, this is a
tedious and not scalable procedure limited to assemblies of few tens of atoms
only. Moreover, their properties correspond to a particle in a box and are not
suitable to probe, for example, transport in 1D.
A very attractive alternative capable of producing single-atom scale structures
in a very efficient and scalable manner is self-assembly. Semiconductor
surfaces, especially silicon, are among the favourite substrates to
self-assemble nanowires. This is motivated by their compatibility with the
current industrial information technology processing,~\cite{Teichert02} and with
state-of-the-art nanoscale processing techniques.~\cite{Ruess2008a} Another
reason is the energy gap at the Fermi level allowing to decouple the nanowire
and substrate electronic states, which is paramount to obtaining 1D isolated
nanowires on a bulk substrate. A range of metal atoms have been self-assembled
into atomic chains on vicinal Si(553) and Si(557)
surfaces,~\cite{Himpsel04,Oncel08} and on Si(111) and Ge(001)
terraces.~\cite{Schaefer08,Schaefer09} In the first instance, the nanolines
assemble along the step edges, hence their spacing (density) is set by the
vicinal angle. In the second instance, the nanolines form a dense array. The
metallic rare earth silicides comprise another interesting family of 1D systems
that has been a focus of attention for many years.~\cite{Owen06a} Their 1D
growth is due to anisotropic epitaxial strain induced by unidirectional lattice
mismatch.~\cite{Bowler04} YSi$_{2}$ nanowires have been found to exhibit 1D
states.~\cite{Zeng08b} However, the major shortcoming of these structures is
their non-constant width and their high reactivity with air.
Here we present a novel and very promising self-assembled silicon only nanoline
structure embedded in the flat terraces of Si(001). These endotaxial (i.e.
epitaxially grown \emph{into} the substrate)~\cite{George91} nanolines, obtained
through hydrogenation of precursor Bi nanolines,~\cite{Owen06a,Owen10} exhibit a
range of remarkable properties. Like other endotaxial structures in Si(001),
such as silicide nanowires,~\cite{Zhian04} they show large aspect ratios with
lengths that can exceed 1\,$\mu$m. Moreover, they offer striking advantages over
all other nanolines mentioned above: they are absolutely straight at a constant
width of exactly four Si dimers (1.54\,nm); they are atomically perfect and
nearly defect free; they are stable upon air exposure for at least 25\,minutes;
they grow on the flat terrace away from any step edges and their density can be
tuned from isolated and non interacting nanolines to dense arrays of possibly
interacting nanolines.
We shall first describe the fabrication of these endotaxial Si nanolines. Then,
we discuss their structural and electronic properties in detail based on
experimental data from scanning tunnelling microscopy (STM), X-ray photoelectron
spectroscopy (XPS), reflection high energy electron diffraction (RHEED) and
density functional theory (DFT) modelling.
\section{Methods}
Endotaxial Si nanolines on Si(001) were fabricated and studied in ultra high
vacuum (UHV – base pressure in the low $10^{-11}$\,mbar) using an Omicron low
temperature STM. The sample preparation chamber is equipped with hot and cold
stages to control the substrate temperature during nanoline growth, several
Knudsen cells and electron beam evaporation sources, a high efficiency hydrogen
cracker and an ion gun. The nanowire growth is monitored in real time using
RHEED, and we further utilise XPS for chemical characterization.
Boron-doped p-type Si(001) substrates with a resistivity of 0.1\,$\Omega$cm were
prepared using a standard \textit{ex-situ} and \textit{in-situ} cleaning
sequence. \textit{Ex-situ} cleaning starts with degreasing the substrates in
ultrasonic acetone followed by isopropanol baths for 5 minutes each. Each
substrate is then rinsed in purified water and cleaned in an oxidising mixture
of H$_2$O$_2$ and H$_2$SO$_4$ (1:1 by volume) for 30 seconds. This surface oxide
is subsequently removed by a 30 second HF (15 wt\%) dip. Finally, the above
oxidation step is repeated to regrow a thin protective oxide film and the
substrate is rinsed in purified water and dried with nitrogen gas.
\textit{In-situ} cleaning starts by degassing the Si(001) substrate for
typically 12 hours at 600$^{\circ}$C while keeping the pressure below
$1\times10^{-9}$\,mbar. To obtain a clean surface, it is then repeatedly
flash-annealed at 1200$^{\circ}$C while making sure the pressure is kept below
$2\times10^{-9}$\,mbar. This step is repeated until the substrate can be
flashed-annealed for 10–15\,s. The last flash-annealing is followed by a
controlled cool-down from 900$^{\circ}$C to the nanowire growth temperature of
570$^{\circ}$C in roughly 1 minute.
To obtain the precursor nanolines (Fig.\,\ref{fig:rheed}~(a)), bismuth is
evaporated
for 10 to 15 minutes onto the clean Si(001) surface kept at 570$^{\circ}$C from
a MBE Komponenten K-cell heated to 470$^{\circ}$C. After the deposition, the
sample is kept at 570$^{\circ}$C for 2 to 3 additional minutes, and then cooled
down to the hydrogenation temperature of 300$^{\circ}$C. The entire deposition
process is monitored in real time with RHEED. The characteristic arc in the
RHEED pattern (Fig.\,\ref{fig:rheed}~(b)) signalling the formation of Bi
nanolines
appears 6 to 8 minutes into the deposition process.
The Bi nanolines thus obtained are exposed to a hydrogen atomic beam from a MBE
Komponenten HABS high efficiency hydrogen cracker heated to
1600$^{\circ}$C.~\cite{Tschersich08} The sample temperature is set at
300$^{\circ}$C during the hydrogenation process. Hydrogen is fed into the UHV
chamber through the cracker itself and the flow rate is controlled by
maintaining a constant chamber pressure of $2.5\times10^{-7}$\,mbar. The Si(001)
surface and the Bi nanolines are exposed to the atomic hydrogen flux for
4–8\,minutes, corresponding to doses of 60--120\,L. We found that this amount of
hydrogen is sufficient to saturate all the Si dangling bonds to form a
monohydride surface.~\cite{Boland90} It also fully strips the bismuth off the
surface, saturating the exposed Si bonds with hydrogen but without affecting the
underlying one dimensional Haiku reconstruction.~\cite{Owen10} This results in a
1D endotaxial hydrogen saturated Si nanoline, called the Haiku
stripe,~\cite{Owen10} which is the subject of detailed investigations in the
following sections.
We performed DFT~\cite{Hohenberg64,Kohn65} calculations using a plane-wave
implementation within the generalised gradient approximation (GGA). All
calculations have been performed with the Vienna \emph{ab initio} simulation
package (VASP).~\cite{Kresse93,Kresse96} Ultrasoft
pseudopotentials~\cite{Vanderbilt90} are used for all the elements considered.
Convergence of forces and energy differences is achieved by selecting a
Monkhorst-pack~\cite{Monkhorst76} mesh of $4\times2\times1$ to sample the
Brillouin zone, and an energy cutoff of 200 eV. We used a unit cell which was
one dimer row wide, ten dimers long and ten layers deep, with the bottom two
layers fixed and the bottom layer terminated with
hydrogen.~\cite{Rodriguez-Prieto09a,Rodriguez-Prieto09b}
\section{Results}
\subsection{Haiku stripes}
\begin{figure}
\includegraphics[width=.9\linewidth]{fig1.png}
\caption{
(a) High resolution STM micrograph of a Bi nanoline on a clean Si(001)
background. (sample bias $V = -3.5$\,V, $I = 200$\,pA, $T = 77$\,K) (b) RHEED
pattern after Bi growth. The arc signifies the Bi nanoline and the bright dots
are due to the 2${\times}$1 reconstruction of the Si surface. (c) High
resolution STM micrograph of the Haiku stripe (4 Si dimers wide diagonal
structure) revealed after the hydrogen induced removal of the Bi atoms. ($V =
-2.8$\,V, $I = 200$\,pA, $T = 77$\,K) (d) RHEED pattern showing the
characteristic arc of a well-ordered 1D structure on the surface, the Haiku
stripe. (e) XPS spectra measured on the Si(001) surface at the different states
of sample preparation. The small peak at 188\, eV is due to B dopant atoms. No
peaks from contamination of C or O are resolved. The top curves are offset by
+100 and +200\,counts/s for clarity.}
\label{fig:rheed}
\end{figure}
While the Bi nanolines appear to STM as a pair of Bi surface dimers, there is a
substantial reconstruction of the underlying Si lattice – a triangular core of
Si embedded in the top five layers of the Si substrate called the Haiku core.
The Bi nanolines, with the unique Haiku structure, have been found to be the
lowest energy reconstruction of Si(001):Bi.~\cite{Bowler02} In DFT calculations,
the Haiku structure is 0.36\,eV/Bi dimer more favourable than the next best
proposed structure,~\cite{Bowler02} and 0.25\,eV/Bi dimer more favourable than
one monolayer of Bi.~\cite{Owen02b} A remarkable result of our work is to find
that this structure remains stable in the hydrogenated Si(001) surface, despite
the removal of the Bi dimers. Both the Bi nanoline (Fig.\,\ref{fig:rheed}~(a,b))
and the Haiku stripe (Fig.\,\ref{fig:rheed}~(c,d)) manifest themselves as an arc
in the RHEED pattern measured along \textless110\textgreater{}, a direct
consequence of their 1D nature and direct evidence of the subsurface
reconstructed core. Annealing experiments on Haiku stripes showed that the arc
feature remains in RHEED up to a temperature of about 400$^{\circ}$C. This
contrasts with previous experiments and simulations on burial of the nanolines
by Si which found that the Haiku core structure spontaneously dissociated during
overgrowth by Si.~\cite{Sakata05} Fig.\,\ref{fig:rheed}~(e) shows XPS spectra of
the Bi-exposed surface before and after exposure to hydrogen. Characteristic Bi
peaks at 464, 440, 157, 159 and 25\,eV, which are clearly resolved after the Bi
nanoline growth, completely disappear upon hydrogenation of the Si(001) surface.
The XPS spectrum of the hydrogenated surface is indistinguishable from the
pristine Si(001) surface. No contaminant species is detected, except for a
minute B signal originating from the bulk dopant atoms. STM micrographs provide
further evidence that the Haiku stripe imaged in Fig.\,\ref{fig:rheed}~(c) is
not just a different appearance of the Bi nanoline due to particular imaging
conditions. The Haiku stripes appear dark independent of bias, which
unquestionably distinguishes them from the Bi nanolines whose STM image contrast
is always bright relative to a H-terminated surface.~\cite{Owen06a}
In Fig.\,\ref{fig:rheed}~(c) the STM micrograph of the hydrogen covered Haiku
core structure shows well resolved dark stripes. The dark stripes observed after
the Bi has been stripped off can be explained by a replacement of each Si-Bi
bond pair by two Si-H bonds.~\cite{Owen10} The proposed structure from DFT of
the resulting H saturated endotaxial Si nanoline is illustrated in the
ball-and-stick model of Fig.\,\ref{fig:model}. Its key feature is the Haiku core
– a major reconstruction of the silicon extending 5 layers below the surface and
composed of 5- and 7-fold Si rings. The four rows of dots indicated by short
solid white lines in the STM micrograph of the Haiku stripe
(Fig.\,\ref{fig:rheed}~(c)) correspond to the four H atoms spanning the Haiku
stripe over exactly 4 Si dimers (1.54\,nm). These rows appear 60 to 70\,pm
deeper than the surrounding Si(001):H surface, in quantitative agreement with
the value expected from the relaxed structure in the DFT simulation
(Fig.\,\ref{fig:model}), and in good agreement with the depth of similar
P-terminated 5-7-5 structures.\cite{McMahon2006a}
The above result suggests that Bi can be fully stripped off from the Bi
nanolines. This is opposite to previous experiments on hydrogenation of Bi
nanolines which have reported that the nanolines were inert,~\cite{Owen06a}
except for a recent indication of Bi nanoline damage with a very large dose
(1000\,L H$_{2}$) by Wang et al.~\cite{Wang08} We explain this controversy on
the basis of the different hydrogenation setups used. All the previous
experiments were done using a hot filament, whereas we are using a high
efficiency hydrogen cracker. The geometry of the H-cracker exposes the sample to
a much greater density of hot hydrogen accelerated towards the sample. The gas
flux impacts the sample at an angle of 25$^{\circ}$. By contrast, a hot filament
sends the gas atoms and molecules in all directions, thus reducing the flux
density. The exact mechanism by which H replaces the Bi atoms is not understood
at present. One possibility is that hot uncracked H$_2$ molecules reach the
surface with sufficient kinetic energy to attack the Bi-Si bonds, thus exposing
Si dangling bonds which are quickly capped by H atoms. To test this idea, we
lowered the temperature of the H cracker, thus greatly reducing the flux of
atomic hydrogen. We observed no reduction of the Bi nanoline stripping rate,
suggesting that hot, high kinetic energy H$_2$ molecules hold the key to the
observed stripping effect.
Since the Si(001) surface is very reactive to water, we took special care to
reduce the exposure to impurity water. We carefully baked the gas line back to
the H-bottle, and further cryopumped the hydrogen gas used in the gas line. In
addition, we have used a dedicated cracker, degassed over many hours at
operating temperature (1600$^{\circ}$C). The chamber pressure with the cracker
at 1600$^{\circ}$C is $8\times10^{-10}$\,mbar. The cracker has a much
higher efficiency than a hot filament.~\cite{Tschersich08} This allows us to
achieve a saturation H coverage with a very small amount of H$_{2}$ gas
(60-120L), thereby greatly reducing the exposure to any impurities. To achieve
saturation of the surface with a hot filament typically H$_{2}$ doses of
500-1000\,L are required~\cite{Hersam01} (15-20 minutes exposure). In our case
any possible carbon or oxygen on the surface after atomic H dosing was below the
XPS detection limit (Fig.\,\ref{fig:rheed}~(e)). Although we cannot measure the
water partial pressure directly, we believe that any impurity water has been
minimised, and that any effect on the nanoline will come from the exposure to
hydrogen.
\begin{figure}
\includegraphics[width=.9\linewidth]{fig2.png}
\caption{Ball-and-stick model of the DFT structure for the Haiku stripe showing
the 5- and 7- fold ring reconstruction of the Haiku core (cross section).
}
\label{fig:model}
\end{figure}
\begin{figure}
\includegraphics[width=.9\linewidth]{fig3.png}
\caption{
(a) STM image of long Haiku stripes. The black arrows denote the start and
end points of a 1.3\,$\mu$m long stripe, while the green arrows are guides for
the eyes. ($V = -2.5$\,V, $I = 180$\,pA, $T = 77$\,K) (b) Magnification of the
1.3\,$\mu$m long stripe denoted by arrows in (a). For clarity the horizontal
axis has been stretched by a factor of two. (c) STM image taken after 25\,min
air exposure without any special treatment showing a Haiku stripe running
diagonally though the image. Neither the surface nor the Haiku stripe have been
attacked by reactants. The dark structures away from the stripe on the
substrate, also present on as grown surfaces, are either missing Si atoms or
places where background Bi has been extracted by H. ($V = -2.2$\,V, $I =
100$\,pA, $T = 77$\,K)
}
\label{fig:air}
\end{figure}
Fig.\,\ref{fig:air}~(a) shows a large scale micrograph where many long stripes
(\textgreater{}0.2\,$\mu$m) can be seen. The longest stripe on this STM
micrograph, whose extremities are identified by black arrows in
Fig.\,\ref{fig:air}~(a,b), reaches 1.3\,$\mu$m. The terrace containing this
stripe is shown in more detail in Fig.\,\ref{fig:air}~(b).
Fig.\,\ref{fig:air}~(a) confirms that the Haiku stripes share a number of
features with the Bi nanolines: they are straight without any kinks and have a
fixed width, they are seen on terraces and terminate at defects and step edges.
These properties have been attributed to the Haiku core. Additionally, the
density of the Haiku stripes on the surface corresponds to the density of the
precursor Bi nanolines, and such stripes have never been seen after
hydrogenation of the clean Si(001) surface. These facts substantiate our
assumption that these lines result from hydrogenation of precursor Bi nanolines,
and the Haiku core is the correct model for this endotaxial Si-in-Si(001)
nanoline.
It has been known for many years that the Si(001) monohydride surface is stable
in air for up to 40\,hours.~\cite{Hersam01} Since the Si-H bond along the Haiku
stripes is of similar strength as the usual Si-H bond on the hydrogenated
Si(001) surface according to DFT, it is reasonable to expect air stability of
the stripes. We checked this hypothesis experimentally, and
Fig.\,\ref{fig:air}~(c) shows a Haiku stripe which is indeed still intact after
25\,min exposure to air. The sample was transferred to the load lock, which was
then slowly vented, while the sample was protected from high velocity molecules
by lying at an angle pointing away from the leak valve as recommended by
ref.~\onlinecite{Hersam01}. Then the sample was stored in air for 25\,min in a
simple plastic box. After the exposure to air, the sample was reinserted into
the load lock and transferred to the STM without any additional treatment.
The length of the Haiku stripes and their air stability offers a unique prospect
to contact them to electrodes enabling standard transport experiments to be
carried out. The air stability allows to consider common \emph{ex-situ}
semiconductor lithographic techniques unlike the precursor Bi nanolines, or
Rare-Earth silicide nanowires. In UHV, this can also be realised either by
depositing metal on top of the surface using shadow-mask technology, or by STM
lithography using the H layer as a mask. Recent developments in the latter
area\cite{Ruess04a} provide the ability to perform more complex contacting and
gating of the stripes on a few-nm length scale.
\subsection{Electronic structure}
STM is ideally suited to probe some of the quantum phenomena emerging in 1D
systems, in particular charge density waves, local spin phases and van Hove
singularities. Bias dependent STM imaging is a very effective way to distinguish
between purely electronic contrast due to, for example, localised states, and
structural contrast reflecting atomic positions. In the following, we compare
topographic images measured at different sample bias voltages and tunnelling
current with DFT simulations. The experiment and model are in excellent
agreement, and reveal the existence of an extended 1D state at the surface along
the Haiku core.
\begin{figure}
\includegraphics[width=.9\linewidth]{fig4.png}
\caption{
Filled state high resolution STM micrographs of the Haiku stripe with the
corresponding DFT simulations shown within the dashed rectangles. (a) Sample
bias $V = -2.5$\,V, $I = 80$\,pA, $T = 77$\,K, DFT at -1.5\,eV for low current.
(b) $V = -2.8$\,V, $I = 200$\,pA, $T = 77$\,K, DFT at -1.5\,eV for high current.
}
\label{fig:fstate}
\end{figure}
High resolution STM micrographs (positive sample bias) reveal the Haiku stripe
as a row of four dark atoms. In Fig.\,\ref{fig:fstate}~(a) recorded at a low
current, they appear regularly spaced and all at the same elevation about 70\,pm
lower (darker color) than the surrounding Si(001) surface. At a higher
tunnelling current (Fig.\,\ref{fig:fstate}~(b)), the middle two atomic sites
appear raised by 5–10\,pm compared to the two outer atomic sites of the Haiku
stripe. STM simulations (inset outlined by dashed rectangles in
Fig.\,\ref{fig:fstate}~(a,b)) based on the DFT relaxed structure of the Haiku
stripe (Fig.\,\ref{fig:model}) perfectly reproduce all of the above
characteristics. They were calculated for the same tunnelling bias, but
different electron density isosurfaces to simulate different tunnelling
currents. Fig.\,\ref{fig:model} shows that the four atomic sites spanning the
Haiku stripe are physically at the same elevation, suggesting that the height
contrast within the four atoms wide stripes is of purely electronic origin. The
DFT simulation further reproduces very accurately the lateral position of the
atomic sites, with a clear shift of the two outer sites towards the edges of the
Haiku stripe when observed at the higher tunnelling current
(Fig.\,\ref{fig:fstate}~(b)).
\begin{figure*}
\includegraphics[width=.9\linewidth]{fig5.png}
\caption{
(a) Positive sample bias (empty states) STM micrographs of two Haiku stripes.
Note the 1D state observed in the upper half of the image recorded at
$V=+2.0$\,V, which disappears when increasing the bias to $+2.5$\,V in the lower
half of the micrograph ($ I = 150$\,pA, $T = 77$\,K). (b) A magnification of the
central state of the Haiku stripe seen at $V = +2.0$\,V and (c) corresponding
DFT model calculated for an energy of +1.0\,eV. (d) A magnification of the Haiku
stripe at $V = +2.5$\,V and (e) the corresponding DFT model calculated for an
energy of +2.5\,eV.
}
\label{fig:estate}
\end{figure*}
Imaging the Si(001) at a positive tunnelling bias, which probes the empty states
of the sample also reveals the four atomic sites across the Haiku stripe as
shown in Fig.\,\ref{fig:estate}. The top half and the bottom half in panel (a)
were taken one after the other with the same tip while only changing the bias
magnitude. In contrast to the negative bias images (Fig.\,\ref{fig:fstate}), the
two central atoms along the Haiku stripe appear shifted away from each other,
while the outer two are shifted towards the centre of the structure
(Fig.\,\ref{fig:estate}). The central atoms now appear lower than the outer ones
— the apparent height difference is reversed compared to the filled state
images. All these features illustrated in the magnified STM image sections
(Fig.\,\ref{fig:estate}~(b,d)) are perfectly reproduced in the corresponding DFT
simulations (Fig.\,\ref{fig:estate}~(c,e)).
The most remarkable feature in the empty state STM images is a bright linear
feature extending along the centre of the Haiku stripe
(Fig.\,\ref{fig:estate}~(a) top). It is only seen in a narrow energy window
around +2.0\,V (Fig.\,\ref{fig:estate}~(a)), and does not match any atomic
position in the Haiku model (Fig.\,\ref{fig:model}). Note that the background
Si(001) is the same in the top and bottom sections of the micrograph shown in
Fig.\,\ref{fig:estate}~(a). Again, all these characteristics are perfectly
reproduced by DFT; the simulation at +1.0\,V (Fig.\,\ref{fig:estate}~(c)) shows
a linear electronic feature in the middle of the structure, whereas it is absent
from the simulation at higher energy (Fig.\,\ref{fig:estate}~(e)).
\begin{figure}
\includegraphics[width=.95\linewidth]{fig6.png}
\caption{(a)–(d) Contour plots of constant charge densities on the Haiku stripe
cross section for the four bands above the Conduction Band Minimum (CBM). The
isosurface colour represents the height within the structure. (a) The CBM (b)
+0.02\,eV (c) +0.04\,eV (d) +0.10\,eV above the CBM. In (d), the charge density
is located in the middle of the Haiku stripe. A cut through the charge density
taken at the position of the dashed line reveals a delocalised state along the
Haiku stripe. On the cut, the colour corresponds to the charge densities: the
blue is a low density and red is a high charge density.}
\label{fig:densities}
\end{figure}
In order to identify the origin of this 1D state, we computed the charge
densities in the vicinity of a Haiku stripe. Fig.\,\ref{fig:densities} shows DFT
isosurfaces of constant charge density for four bands starting at the conduction
band minimum (CBM) towards higher energy states corresponding to the states
probed primarily by STM at positive sample bias. The isosurfaces demonstrate the
substantially different electronic structure below the surface of the Haiku
core, compared to pure Si. The first three bands above the lowest unoccupied
molecular orbital (LUMO) show little to no charge density within the Haiku core
and none of them are continuous along the stripes. On the other hand, the fourth
band in Fig.\,\ref{fig:densities}~(d) has its maximal charge density in the
middle of the Si reconstruction. A cross-section through the charge density
along the centre of the stripe (Fig.\,\ref{fig:densities}~(d), inset) shows that
this fourth band is delocalised along the endotaxial nanoline, in excellent
agreement with the STM observation (Fig.\,\ref{fig:estate}~(a), top).
\subsection{Discussion}
The Haiku stripes are extremely high aspect ratio Si-only 1D structures of
unprecedented perfection. They offer a range of very attractive features towards
the experimental exploration of 1D quantum physics. There is the experimental
opportunity to attach contacts, which is paramount to study transport in a truly
isolated 1D model system. The prime avenue will be the 1D state confined to the
Haiku core, but delocalized along it, which we have identified in STM and
confirmed by DFT. Because it is located just above the minimum of the conduction
band, one challenge will be to dope it to a suitable level. Local removal of the
H atoms on the top of the stripe, followed by adsorption of PH$_{3}$
(ref.~\onlinecite{Ruess04a}) could lead to local doping of this state, for
example. The physical and electronic structure of the stripes is well
understood, so that experiments can enjoy strong theoretical support, while the
micrometre length gives access to the infinite length regime addressed by
theory. Moreover, the degree of interaction between neighbouring nanowires is
tunable. Adjustable nanoline density is achieved by means of the growth
temperature and Bi flux during the self-assembly of the Bi nanoline precursor.
This is distinctly different from other self-assembled nano-scale wires, whose
density can either not be adjusted\cite{Wang2004a} or is set by the step density
of a vicinal surface.~\cite{Himpsel04}
As well as the intrinsic properties of the Haiku stripes, they can also be used
as a template for self assembling 1D chains of other atomic species. Previous
experiments have shown templating properties for the Bi nanolines.\cite{Owen06b}
According to modelling, the stripes host favourable adsorption sites for a range
of metal atoms. For example, Cu may diffuse to subsurface sites, forming
encapsulated single-atom chains.~\cite{Rodriguez-Prieto09b} We are investigating
the possibility that Mn atoms may also move into subsurface sites, producing
atomic chains with strong magnetic properties. The H-termination of the Haiku
stripes offers the exciting possibility that such subsurface metal atomic chains
would be stable in air.
Looking more broadly in the direction of future device fabrication, the H
termination of these nanolines and consequent air stability, their micrometre
length and their stability to an elevated temperature make these nanolines
compatible with a wide variety of semiconductor processing technologies.
\section{Conclusions}
We have described the properties of a novel endotaxial Si nanoline in Si(001)
identified as the Haiku stripe. Its core holds a delocalised state just above
the conduction band minimum, which combined with the length and the perfection
of the Haiku stripe makes it a very promising and appealing structure to explore
quantum properties emerging in 1D. Moreover, the stability to air and to higher
temperature, which we have demonstrated here, offer exciting experimental
opportunities. The detailed matching of DFT simulations and STM micrographs
confirms that the structure of the stripes is very well understood and settles
the Haiku core model on firm ground. Thus the discovery of this unique Si-in-Si
nanoline opens possibilities for exploring 1D physics and has a great potential
for the fabrication of nanoscale electronic devices on a technologically
relevant substrate.
\begin{acknowledgments}
The authors are very grateful to G. Manfrini for his expert technical support,
and to A. Rodriguez-Prieto for scientific discussions. We thank the Swiss
National Science Foundation for financial support through the National Center of
Competence in Research MaNEP (Materials with Novel Electronic Properties) and
division II. D.R. Bowler was supported by the Royal Society, and acknowledges
computational facilities in the London Centre for Nanotechnology.
\end{acknowledgments}
|
\section{Introduction}
In mathmatics, algebriac strucutres usually represent the axiomatization of known objects, hide in the dephts of a unsolved problem. The strucutre under consideration in this work, was inspired in a problem aparentley trivial: if we have a graph defined as the usual like the 2-tuple (Berge \cite{Berge})
\begin{center}
$G=(V,E),$
\end{center}
where V is the set of vertexes and E is the set of the unordered pair of vertexes (in a non-oriented graph) of edges, then we assume a graph G with his subgraph G', and we want to find the graph X in the graph G, this is
\begin{center}
$G=G'\cup X.$
\end{center}
If we consider only the asumptions on graph theory so far, which is bassiclay based on ZF set theory axiom system, we have no solution to the problem. If we look, instead, at the extension made in the works of Carroll \cite{Carroll}(2009) and Gatica \cite{Gatica}(2010), considering the existence of an antiset B where
\begin{center}
$A \cup B = \emptyset,$
\end{center}
with $B=\tilde{A}$ the solution would be
$$G \cup \tilde{G'}= X,$$
where naturally emerges the concept of antigraph as a solution for a graph boolean equation system with one variable. This way of working graphs will help to bring simplest solutions to apparently complex problems.
Before we start to define our algebraic structure, we will introduce some preliminar concepts, like graphs, the definition of set union al link, and the basic of set/antiset theory.
\section{Introduction to Link Algebra}
\subsection{Graph Theory}In graph theory, the concepts of vertex and edge are essential: is the source of thousands of objects and forms, but the problem lies in one thing: it is still atached to combinatorial analisis and arithmetic formulations. Is not a propper object with an algebraic background: it is still a mathmatical bricolage of external mathematicas sub-areas.
For many years, graphs has been consider as a mere exentricity born in the mind of Euler to solve the Koenigsberg bridge problem; in the XX century, Claude Berge made a more formal definition based on the definition by Konig, giving emphasis in the vertex and edges.In this new century, the view of graph has lightly changed: they still remains as 2-tuples only used when the ocation appears; but if we take a closer look of some basic concepts in their theory, we will be able to see the strucure that lies beneath the surface.
In order to make a better background, lets remind the following concepts: complete bipartite graph, null graph and graph union.
A complete bipartite graph is ussually defined as
\begin{center}
$K_{n,m}:=(V_1 \cup V_2, E),$
\end{center}
where E is the set of the pairs formed by the conection of the edges v in $V_1$ and w in $V_2$, with n and m the number of vertex contained in the set $V_1$ and $V_2$ respectively.
The next concept is the null graph, which is defined as
$$N_n:=(V, \{\emptyset\})$$
having vertexes but no edges, where n is the number of edges.
Finally, we have graph union which is definded as
$$G:=(V \cup V', E \cup E'),$$
where V and V' are the vertexes of two diferent graphs, being E and E' their edges. With this three concepts, we are now able to construct the main operations of Link Algebra.
\subsection{Basic Definitions} From the concepts view of graph theory, we can construct the main definitios of our work.
\begin{definition}
We define the union betwen two graphs G and G' as
$$G \dashv \vdash G':=(V \cup V', E \cup E')$$
\end{definition}
\begin{definition}
We define the linking betwen two graphs G and G' as
\begin{enumerate}
\item $N_n \dot{\Lambda} N'_m :=K_{n,m}$, if they are null graphs;
\item $G \dot{\Lambda} G':=G \dashv \vdash G' \dashv \vdash K_{n,m}$, where $N_n$ and $N'_m$ are the null graphs of G and G' respectively. \footnote{If the reader is interested, use as an example $G=(\{a,b\},\{\{a,b\}\})$ and $G'=(\{c,d\},\{\{c,d\}\})$.}
\end{enumerate}
\end{definition}
Once defined this operations, we will construct the fundamental objects of graph theory: Vertexes and Edges, objects that will allow us to create the laws that will support our algebraic structure.
Back in the definition of null graph, it is not to difficult to define a single vertex, we only need n=1 and the definition is done. In the other case, the edge, we will need the definition of conection betwen two null graphs of one vertex. Hence, the concepts could be defined as follows:
\begin{definition}
We define a \textbf{Vertex} v from the single graph G as
$$v_G:=N_1,$$ where $N_1$ is the null graph of the single graph G.
\end{definition}
\begin{definition}
We define a \textbf{Edge} e from the edge graph G as
$$e_G:=N_1 \dot{\Lambda} N'_1,$$ where $N_1$ and $N'_1$ are the null graphs of the single subgraph H and H' of G respectively.
\end{definition}
Before we start defining a Conective Algebra an its laws, lets make a little introduction to set/antiset theory, for prepare the ground to the propper concept of antivertex and antiedge.
\section{A step back in to the land of sets}
\subsection{Sets and Antisets}When Zermelo and Fraenkel made the axioms of set theory, they consider sets as an unique unit in the set universe. After the appearance of the antiparticle in phisics, many mathematicians started to search a possible analogy in the land of sets. The problem was that the axiomatization made was unable to accept such antielements or antisets. In the 90's some mathematicians started to ask if there was a posibility to make an extension to accept such idea. Then in this century, works like those from Caroll or Gatica made de propper needed axiomatization. In this brief section, we will introduce the basis of set/antiset theory for using in graph theory to define antigraph, taking the notation from the work of Carroll.
As we defined before, let be A and B sets which
$$A \cup B = \emptyset,$$
where $B= \tilde{A}$, definded as antiset of the set A.
In both works, they start by defining union in to a element view, so, union, in this form is defined for a set A and B which
$$ A \cup B:= \bigcup^{}_{} \{A,B\},$$
where A could be $A=\{a\}$ and $B=\{b\}$ where naturally
\begin{tabular}{llll}
\\
& $A \cup B$ & $= \{a\} \cup \{b\}$ \\
& & $= \{a,b\}$, \\
\end{tabular}
\\
which is also called axiom of fusion. If A and B are sets of subsets, the subsets interact as they were just elements, this is, there are no fusion inside fusions. Is also important to mention that is linear and the sign of the set also keeps in their elements, this is
\begin{tabular}{llll}
\\
& (a) $\widetilde{A \cup B}$ & $= \tilde{A} \cup \tilde{B}$ \\
& (b) $ \tilde{A}$ & $= \{\tilde{a}\},$
\end{tabular}
\\
in Carroll's work \cite{Carroll} , this is clearly defined, and derived naturally from the ZF axiomatic extention of the fusion and the union of elements.
with this, it is posible to define the Boolean Algebra with an extention of antielements. For that reason, the new universe set will be defined as a set with both positive and negative properties. The boolean algebraic properties remains as ussually for the operations union and intersection, now extended to antisets.
\subsection{Consecuences in Graph theory} Using this extension in graphs, gives the naturally the concept of antigraph, this is
$$\tilde{G}:=(\tilde{V}, \tilde{E}),$$
which joined to the graph G (now extended) it result to be
$$ G \cup \tilde{G}:=(\{\emptyset\}, \{\emptyset\}),$$
where $(\{\emptyset\}, \{\emptyset\})$ is definded as the empty graph $\phi_G$.
In the same way, we will define an antivertex and an antiedge as
$$\tilde{v_G}:=\tilde{N_1},$$
$$\tilde{e_G}:=\widetilde{N_1 \dot{\Lambda} N'_1},$$
Other important consecuence, is the solution of the equation $G = G' \cup X$ seen in the introduction as $X =G \cup \tilde{G'}$.
Showed this concepts, we are finally able to define our algebraic structure.
\section{Link Algebra}
\subsection{Definition of Link Algebra} In abstract algebra, in general, the algebraic structure represents the true goal beneath the scrutine of those hide patterns in the behave of a mathematical object. Graphs has been seen as diagrams, even confussed with Graphics, missing their true nature in the world of mathmatics.
The structure, finnaly introduced in this section, has developed from the concepts in graph theory redefinded by set/antiset theory and the new definiton of vertex by using the concept of null graphs.
\begin{definition}
A \textbf{Link Algebra} is a set G (extended to antigraphs) of vertexes v with the operations $\dashv \vdash$ and $\dot{\Lambda}$ satisfing the following axioms
\begin{enumerate}
\item Closure
$$v \dashv \vdash w \in G,$$
$$v \dot{\Lambda} w \in G.$$
\item Idempotency
$$v \dashv \vdash v = v,$$
$$v \dot{\Lambda} v = v.$$
\item Asociativity \footnote{Exept for $v=w$ and $t=\tilde{v}$, to avoid the same problem in set/antiset theory. If the reader is interested, see Gatica's work.}
$$v \dashv \vdash (w \dashv \vdash t) = (v \dashv \vdash w) \dashv \vdash t,$$
$$v \dot{\Lambda} (w \dot{\Lambda} t) = (v \dot{\Lambda} w) \dot{\Lambda} t.$$
\item Identity element
There exist an element $\phi$, that for every elemet of G the equation
$$v \dashv \vdash \phi = v,$$
$$v \dot{\Lambda} \phi = v.$$
\item Inverse element
For every element v of G there exist an element $\tilde{v}$ in G that
$$v \dashv \vdash \tilde{v} = \tilde{v} \dashv \vdash v = \phi,$$
$$v \dot{\Lambda} \tilde{v} = \tilde{v} \dot{\Lambda} v = \phi.$$
\item Conmutativity
$$v \dashv \vdash w = w \dashv \vdash v,$$
$$v \dot{\Lambda} w = w \dot{\Lambda} v.$$
\item Distributivity
\begin{enumerate}
\item $v \dot{\Lambda} (w \dashv \vdash t) = (v \dot{\Lambda} w) \dashv \vdash (v \dot{\Lambda} t).$
\item $(w \dashv \vdash t) \dot{\Lambda} v = (w \dot{\Lambda} v) \dashv \vdash (t \dot{\Lambda} v).$
\end{enumerate}
\item Conectivity
For every equation $\Gamma$ and $\Gamma'$ of G, where
$$\Gamma = V \dashv \vdash E,$$
being V and E defined as
\begin{enumerate}
\item $V= v_1 \dashv \vdash \ldots \dashv \vdash v_p$
\item $E= w_1 \dot{\Lambda} t_1 \dashv \vdash \ldots \dashv \vdash w_q \dot{\Lambda} t_q,$
\item $N_n = V \dashv \vdash w_1 \dashv \vdash t_1 \dashv \vdash \ldots \dashv \vdash w_q \dashv \vdash t_q $
\end{enumerate}
with $N_n$ (where $n=p + 2q$) and $N'_m$ their vertex equations
$$\Gamma \dot{\Lambda} \Gamma':= \Gamma \dashv \vdash \Gamma' \dashv \vdash N_n \dot{\Lambda} N'_m.$$
\end{enumerate}
\end{definition}
Some of the axioms within the structure, bring the open posiblity to construct some important objects in graph theory, such as Paths, Cycles, Stars and complete graphs; but before, we will look at the fundamental properties of the Link Algebra to continue with the consrtuction of main objects such K, P, C and S equations.
\subsection{Theorems of Link Algebra} The following theorems, except few, have proofs using just the basic of group theory. If the reader is interested, could see this proofs in the work of Kaufman and Percigout \cite{Kauffman}, in the sections of groups.
\begin{theorem}
if v,w and t $\in$ G and $v \dot{\Lambda} t = w \dot{\Lambda} t$, then $v=w$
\end{theorem}
\begin{proof}
One way to solve the equation $v \dot{\Lambda} t = w \dot{\Lambda} t$ is by linking $\tilde{t}$ and using the axiom (8), this is
\begin{tabular}{llll}
& $v \dot{\Lambda} t$ & $= w \dot{\Lambda} t$ & {} \\
& $(v \dot{\Lambda} t) \dot{\Lambda} \tilde{t}$ & $= (w \dot{\Lambda} t) \dot{\Lambda} \tilde{t}$ \\
& $v \dot{\Lambda} t \dashv \vdash \tilde{t} \dashv \vdash (v \dashv \vdash t) \dot{\Lambda} \tilde{t}$ & $=w \dot{\Lambda} t \dashv \vdash \tilde{t} \dashv \vdash (w \dot{\Lambda} t) \dot{\Lambda} \tilde{t}$ \\
& $v \dot{\Lambda} t \dashv \vdash \tilde{t} \dashv \vdash (v \dot{\Lambda} \tilde{t}) \dashv \vdash (t \dot{\Lambda} \tilde{t})$ & $=w \dot{\Lambda} t \dashv \vdash \tilde{t} \dashv \vdash (w \dot{\Lambda} \tilde{t}) \dashv \vdash (t \dot{\Lambda} \tilde{t})$ \\
& $v \dot{\Lambda} t \dashv \vdash \tilde{t} \dashv \vdash (v \dot{\Lambda} \tilde{t}) \dashv \vdash \phi$ & $=w \dot{\Lambda} t \dashv \vdash \tilde{t} \dashv \vdash (w \dot{\Lambda} \tilde{t}) \dashv \vdash \phi$ \\
& $v \dot{\Lambda} t \dashv \vdash \tilde{t} \dashv \vdash (v \dot{\Lambda} \tilde{t})$ & $=w \dot{\Lambda} t \dashv \vdash \tilde{t} \dashv \vdash (w \dot{\Lambda} \tilde{t})$ \\
& $v \dot{\Lambda} t \dashv \vdash \tilde{t} \dot{\Lambda} (\phi \dashv \vdash v)$ & $=w \dot{\Lambda} t \dashv \vdash \tilde{t} \dot{\Lambda} (\phi \dashv \vdash w)$ \\
& $v \dot{\Lambda} t \dashv \vdash \tilde{t} \dot{\Lambda} v$ & $=w \dot{\Lambda} t \dashv \vdash \tilde{t} \dot{\Lambda} w$ \\
& $v \dot{\Lambda} (t \dashv \vdash \tilde{t})$ & $=w \dot{\Lambda} (t \dashv \vdash \tilde{t})$ \\
& $v \dot{\Lambda} \phi$ & $w \dot{\Lambda} \phi$ \\
& $v$ & $=w$ \\
\end{tabular}
\\
The other way, is by using the axiom (3), (5) and (4).
\end{proof}
\begin{theorem}
For every v in G, $v \dot{\Lambda} \phi = \phi \dot{\Lambda} v$
\end{theorem}
\begin{theorem}
A link Algebra, has a single neutral element.
\end{theorem}
\begin{proof}
Lets supose that there is a $\phi'$ that $\phi \dot{\Lambda} \phi' = \phi$ and $\phi' \dot{\Lambda} \phi = \phi',$ occurs.
Using the axiom (6), the consequence is immediate. The same occurs for $\dashv \vdash$.
\end{proof}
\begin{theorem}
For every v of G, $\tilde{v} \dot{\Lambda} v = v \dot{\Lambda} \tilde{v}$
\end{theorem}
\begin{theorem}\label{theo alg rght}
if v,w and t $\in$ and $t \dot{\Lambda} v = t \dot{\Lambda} w$, then $v=w$
\end{theorem}
\begin{theorem}
Every element of a link algebra have a single inverse element.
\end{theorem}
\begin{theorem}
If v $\in$ G, then $\widetilde{\tilde{v}} = v$
\end{theorem}
\begin{proof}
We want to know if there is some $\widetilde{\tilde{v}}$ that
$$\tilde{v} \dashv \vdash \widetilde{\tilde{v}} = \phi$$
lets take the axiom (5) for $\dashv \vdash$
$$\tilde{v} \dashv \vdash v = \phi,$$
using the theorem \ref{theo alg rght}, we proof that $\widetilde{\tilde{v}} = v$
\end{proof}
\begin{theorem}
If v and w $\in$ G, exists single x and y of G that $v \dot{\Lambda} x = B$ and $y \dot{\Lambda} v = B$
\end{theorem}
\begin{theorem}
(G, $\dashv \vdash$) obey group theorems.
\end{theorem}
\subsection{Objects in Link Algebra} In Link algebra, same as graph theory, there are fundamental objects like Paths, Cycles, Stars and Complete Conective forms. In this section we will present the definition of those objects.
\begin{definition}\label{def path}
$\forall$ $v_1, \dots, v_n \in$ G, an equation is called \textbf{Path} if
$$ P_n = v_1 \dot{\Lambda} v_2 \dashv \vdash \dots \dashv \vdash v_{n-1} \dot{\Lambda} v_n$$
\end{definition}
\begin{definition}
$\forall$ $v_1, \dots, v_n \in$ G, an equation is called \textbf{Cycle} if
$$C_n = P_n \dashv \vdash v_1 \dot{\Lambda} v_n$$
\end{definition}
\begin{definition}\label{def star}
$\forall$ $v_1, \dots, v_n$ and w $\in$ G, an equation is called \textbf{Star} if
$$ S_n = w \dot{\Lambda} (v_1 \dashv \vdash \dots \dashv \vdash v_n)$$
\end{definition}
\begin{definition}
$\forall$ $v_1, \dots, v_n \in$ G, an equation is called \textbf{Complete Conective Form} if
$$K_n = v_1 \dot{\Lambda} \dots \dot{\Lambda} v_n.$$
\end{definition}
\begin{definition}
$\forall$ v and w $\in$ G, an Edge is definded as
$$\bar{\varepsilon} = v \dot{\Lambda} w.$$
\end{definition}
\begin{definition}
$\forall$ $v_1, \dots, v_n \in$ G, an equation is called Null Form if
$$N_n = v_1 \dashv \vdash \dots \dashv \vdash v_n.$$
\end{definition}
\subsection{Aplication on Graph Theory} As we said in the introduction, we will show how Link Algebra is useful on Graph Theory. Therefore, we will give some main theorems derived from the figures in the earlier section. Some will be demonstrated, the rest are easy to proof from the other theorems.
\begin{theorem}
if $V=v_1 \dashv \vdash \dots \dashv \vdash v_n$, then $K_n = V \dot{\Lambda} V.$
\end{theorem}
\begin{proof}
Using the definition of $K_n$ and using recursively the axiom (8)
\begin{tabular}{llll}
& $K_n$ & $= v_1 \dot{\Lambda} \dots \dot{\Lambda} v_n$ & {}\\
& {} & $= v_1 \dashv \vdash \dots \dashv \vdash v_n \dashv \vdash v_{n-1} \dot{\Lambda} v_n \dashv \vdash \dots \dashv \vdash v_1 \dot{\Lambda} (v_2 \dashv \vdash \dots \dashv \vdash v_n)$ \\
& {} & $=v_1 \dot{\Lambda} v_1 \dashv \vdash \dots \dashv \vdash v_n \dot{\Lambda} v_n \dashv \vdash S_1 \dashv \vdash \dot \dashv \vdash S_n$ \\
& {} & $=v_1 \dot{\Lambda} v_1 \dashv \vdash \dots \dashv \vdash v_n \dot{\Lambda} v_n \dashv \vdash S$ \\
& {} & $=v_1 \dot{\Lambda} v_1 \dashv \vdash \dots \dashv \vdash v_n \dot{\Lambda} v_n \dashv \vdash S \dashv \vdash S$ \\
& {} & $=v_1 \dot{\Lambda} (v_1 \dashv \vdash \dots \dashv \vdash v_n) \dashv \vdash \dots \dashv \vdash v_n \dot{\Lambda} (v_1 \dashv \vdash \dots \dashv \vdash v_n \dot{\Lambda} v_n)$ \\
& {} & $=V \dot{\Lambda} V$ \\
\end{tabular}
\\
Where $S=S_1 \dashv \vdash \dots \dashv \vdash S_n$ And the proof is done.
\end{proof}
\begin{theorem}\label{star v theo}
if $V= v_1 \dashv \vdash \dots \dashv \vdash v_n$, then $S_n \dashv \vdash V = S_n$
\end{theorem}
\begin{proof}
using the definition of star
\begin{tabular}{llll}
& $S_n \dashv \vdash V$ & $= w \dot{\Lambda} (v_1 \dashv \vdash \dots \dashv \vdash v_n) \dashv \vdash V$ & {} \\
& {} & $=w \dot{\Lambda} V \dashv \vdash V$ & (using the hipotesis) \\
& {} & $=w \dot{\Lambda} V \dashv \vdash V \dot{\Lambda} \phi$ & (using (4)) \\
& {} & $=w \dot{\Lambda} (V \dashv \vdash \phi)$ & (using (7)) \\
& {} & $=S_n$ & (using (4) and the def \ref{def star})
\end{tabular}
\\
which is, as a matter of fact, a Star.
\end{proof}
\begin{theorem}
$P_n$ is inductive.
\end{theorem}
\begin{proof}
Using Peano Axioms, the case n=1 is trivial. Now, we need to prove that $P_n$ implies $P_{n+1}$.
\begin{tabular}{llll}
& $P_{n+1}$ & $=v_1 \dot{\Lambda} v_2 \dashv \vdash \dots \dashv \vdash v_{n-1} \dot{\Lambda} v_n \dashv \vdash v_n \dot{\Lambda} v_{n+1}$ & (using def \ref{def path}) \\
& {} & $=P_n \dashv \vdash v_n \dot{\Lambda} v_{n+1}$ & (using def \ref{def path} again) \\
\end{tabular}
\\
so, $P_n$ is inductive.
\end{proof}
\begin{corollary}\label{cor cy}
$C_n$ is inductive.
\end{corollary}
\begin{proof}
the case n=1, as in the theorem, is trivial. We need to prove that $C_n$ implies $C_{n+1}$.
\begin{tabular}{llll}
& $C_{n+1}$ & $= P_{n+1} \dashv \vdash v_1 \dot{\Lambda} v_{n+1}$\\
& {} & $= P_n \dashv \vdash v_n \dot{\Lambda} v_{n+1} \dashv \vdash v_1 \dot{\Lambda} v_{n+1}$\\
& {} & $= P_n \dashv \vdash R \dashv \vdash v_1 \dot{\Lambda} v_n \dashv \vdash \widetilde{v_1 \dot{\Lambda} v_n}$ \\
& {} & $= P_n \dashv \vdash v_1 \dot{\Lambda} v_n \dashv \vdash R \dashv \vdash \widetilde{v_1 \dot{\Lambda} v_n}$ \\
& {} & $=C_n \dashv \vdash R \dashv \vdash \widetilde{v_1 \dot{\Lambda} v_n},$ \\
\end{tabular}
\\
calling $R=v_n \dot{\Lambda} v_{n+1} \dashv \vdash v_1 \dot{\Lambda} v_{n+1}$, $C_n$ is inductive.
\end{proof}
\begin{corollary}
$S_n$ is inductive.
\end{corollary}
\begin{proof}
Same as corollary \ref{cor cy} .
\end{proof}
\begin{theorem}
if $S_n= S_{n-1} \dashv \vdash \bar{\varepsilon}$ and $S'_m=S'_{m-1} \dashv \vdash \bar{\varepsilon}$, then
$$S_n \dashv \vdash S'_m \dashv \vdash \tilde{\bar{\varepsilon}}=S_{n-1} \dashv \vdash S'_{m-1}.$$
\end{theorem}
\begin{proof}
first we use the definitions
\begin{tabular}{llll}
& $S_n \dashv \vdash S'_m \dashv \vdash \tilde{\bar{\varepsilon}}$ & $=(S_{n-1} \dashv \vdash \bar{\varepsilon}) \dashv \vdash (S'_{m-1} \dashv \vdash \bar{\varepsilon}) \dashv \vdash \tilde{\bar{\varepsilon}}$ \\
& {} & $=(S_{n-1} \dashv \vdash S'_{m-1}) \dashv \vdash (\bar{\varepsilon} \dashv \vdash \bar{\varepsilon}) \dashv \vdash \tilde{\bar{\varepsilon}}$ \\
& {} & $=S_{n-1} \dashv \vdash S_{m-1} \dashv \vdash (\bar{\varepsilon} \dashv \vdash \tilde{\bar{\varepsilon}})$ \\
& {} & $=S_{n-1} \dashv \vdash S_{m-1} \dashv \vdash \phi$ \\
& {} & $=S_{n-1} \dashv \vdash S_{m-1}$ \\
\end{tabular}
\\
then, we only use idempotency on edges and the axiom of inverse element.
\end{proof}
\begin{theorem}
If $V=V' \dashv \vdash V''$ where $V'=v_1 \dashv \vdash \dots \dashv \vdash v_j$, $V''=v_j \dashv \vdash \dots \dashv \vdash v_n$ where $j \in \{1, \dots, n \}$, then $S_n=S'_j \dashv \vdash S''_{n-j+1}$.
\end{theorem}
\begin{proof}
Using the axiom of distributivity and asociativity, and the hipothesis, the proof is obvious.
\end{proof}
Before we present the next definition, lets define the following concepts
\begin{definition}
If $f(v_1, \dots, v_n)=\Gamma$ is a conective form where $f:G^n \rightarrow G$, we will call Conective Form Complement to
$$\Gamma^c= K_n \dashv \vdash \tilde{\Gamma}$$
\end{definition}
\begin{definition}\label{def cf}
If $\Gamma$ is a Conective Form, we will call a Star Composed Form to
$$\Gamma_s= N_n \dot{\Lambda} N_n \dashv \vdash \widetilde{\Gamma^c}$$
\end{definition}
\begin{example}
Lets say that we have a conective form
$$\Gamma = v_1 \dot{\Lambda} v_2 \dot{\Lambda} v_3 \dashv \vdash v_3 \dot{\Lambda} v_4.$$
The complement of $\Gamma$, using the definition will be:
\begin{tabular}{llll}
& $\Gamma^c$ & $= N_4 \dot{\Lambda} N_4 \dashv \vdash \widetilde{(v_1 \dot{\Lambda} v_2 \dot{\Lambda} v_3 \dashv \vdash v_3 \dot{\Lambda} v_4)}$ \\
& {} & $=v_1 \dot{\Lambda} (v_2 \dashv \vdash v_3 \dashv \vdash v_4 \dashv \vdash v_5) \dashv \vdash \dots \dashv \vdash v_4 \dot{\Lambda} v_5 \dashv \vdash \widetilde{v_1 \dot{\Lambda} v_2 \dot{\Lambda} v_3} \dashv \vdash \widetilde{v_3 \dot{\Lambda} v_4}$ \\
& {} & $=v_1 \dot{\Lambda} v_4 \dashv \vdash v_4 \dot{\Lambda} v_2.$ \\
\end{tabular}
\\
Hence, the SCF:
\begin{tabular}{llll}
& $\Gamma_s$ & $= N_4 \dot{\Lambda} N_4 \dashv \vdash \widetilde{(v_1 \dot{\Lambda} v_4 \dashv \vdash v_4 \dot{\Lambda} v_2)}$ \\
& {} & $=(v_1 \dashv \vdash v_2 \dashv \vdash v_3 \dashv \vdash v_4) \dot{\Lambda} (v_1 \dashv \vdash v_2 \dashv \vdash v_3 \dashv \vdash v_4) \dashv \vdash \widetilde{v_1 \dot{\Lambda} v_4} \dashv \vdash \widetilde{v_4 \dot{\Lambda} v_2}$ \\
& {} & $=v_1 \dot{\Lambda} (v_2 \dashv \vdash v_3) \dashv \vdash v_2 \dot{\Lambda} (v_1 \dashv \vdash v_3) \dashv \vdash v_3 \dot{\Lambda} (v_1 \dashv \vdash v_2 \dashv \vdash v_4) \dashv \vdash v_4 \dot{\Lambda} v_3.$ \\
\end{tabular}
\\
\end{example}
\begin{definition}\label{theo card star}
If $V=v_1 \dashv \vdash \dots \dashv \vdash v_n$, with $S_n$ defined as usual, for $f:G \rightarrow G$, $f(S_n) = S'_m$, if
$$card(V) = card(V'),$$
With card, defined as the cardinality of the star.
\end{definition}
\begin{definition}
The graphs $\Gamma$ and $\Gamma'$ will be isomorphic (ordered form maximum to minimum star by cardinality) if there is some $f:G \rightarrow G$, lineal to $\dashv \vdash$ that
$$f(\Gamma_s)= \Gamma'_s.$$
\end{definition}
\begin{example}
Lets supose we have the graphs $\Gamma = v_1 \dot{\Lambda} v_2 \dot{\Lambda} v_3 \dashv \vdash v_3 \dot{\Lambda} v_4$ and $\Gamma'= w_1 \dot{\Lambda} (w_2 \dashv \vdash w_3 \dashv \vdash w_4) \dashv \vdash w_2 \dot{\Lambda} w_4$.
Applying the definition \ref{def cf} in both, we have
$$\Gamma_s= v_1 \dot{\Lambda} (v_2 \dashv \vdash v_3) \dashv \vdash v_2 \dot{\Lambda} (v_1 \dashv \vdash v_3) \dashv \vdash v_3 \dot{\Lambda} (v_1 \dashv \vdash v_2 \dashv \vdash v_4) \dashv \vdash v_4 \dot{\Lambda} v_3$$
$$\Gamma'_s= w_1 \dot{\Lambda} (w_2 \dashv \vdash w_3 \dashv \vdash w_4) \dashv \vdash w_2 \dot{\Lambda} (w_3 \dashv \vdash w_4) \dashv \vdash w_3 \dot{\Lambda} w_1 \dashv \vdash w_4 \dot{\Lambda} (w_1 \dashv \vdash w_2),$$
as we see, ordering by maximum to minimum star and applying a function to $\Gamma$ (lineal to $\dashv \vdash$),the stars have equal cardinality, so, we acctually have $f(\Gamma_s) = \Gamma'_s$.
\end{example}
\begin{theorem}
if $\Gamma^c$ is the complement of $\Gamma$ with n vertex both, then
$$\Gamma^c \dashv \vdash \Gamma= K_n.$$
\end{theorem}
\begin{proof}
Using the defintion of complementary graph
\begin{tabular}{llll}
& $\Gamma^c \dashv \vdash \Gamma$ & $=(K_n \dashv \vdash \tilde{\Gamma}) \dashv \vdash \Gamma$ \\
& {} & $=K_n \dashv \vdash (\tilde{\Gamma} \dashv \vdash \Gamma)$ \\
& {} & $=K_n \dashv \vdash \phi$ \\
& {} & $=K_n.$ \\
\end{tabular}
\\
the graph and the complement make the complete graph.
\end{proof}
\section{A brief view of applications on multi, pseudo an oriented Graphs}
\subsection{A reminder of the definitions}\label{subsect} In many mathematical bibliography, graphs seems to be confused with other kind of objects like multi graphs, pseudo graphs and oriented graps. The truth is, there are conceptual differences, of which we are going to make a short remind to the reader.
\begin{definition}
A graph $G=(V,E)$ is called \textbf{multigraph} if some edges could be equal; between two nodes, can exist many edges.
\end{definition}
\begin{definition}
A graph is called \textbf{pseudograph} if there are edges with the form $\{a,a\}$, but no multiedges of the canonical form.
\end{definition}
\begin{definition}
A graph with the form $G=(V,E)$ is called \textbf{oriented} if the edges of E have the form $e=(v,w)$, which is an ordered pair of vertexes.
\end{definition}
\subsection{Arithmetical Graphs} Wolfram, in an article referred at Pseudographs \cite{Wolfram}, referred as a graph with loops an multiedges. For some authors, pseudographs and graphs are deeply connected by the idea of multiset. But the problem was that the definition of those graphs only consider multiedges but no multivertexes, by that reason there is no posibility to define precisely the operation of union.
In 2003, Wildberger \cite{Wildberger} presented a work proposing an alternative form to treat multisets by using linear notation. In that work, he propose a third operation on sets: the sum, were elements of two sets A and B just add in even if elements of both repeat. Let's give a simple example. Let's define the multisets A and B as
$$A= \{a,a,b\}$$
$$B=\{b,b,c\}$$
Then
$$A+B=\{a,a,b,b,b,c\}$$
If we add antisets, we will have something very close to the arithmetical structure of $\mathbb{Z}$ with the operation sum, defined as an abelian group. In the work of Carroll, there is an extension of the concept of negative sets to negative numbers that acctually shows how group theory can be constructed using the new extension to ZF axiomatic.
Multisets are linear, this is, for an external operation product in $\mathbb{Z}$
\begin{enumerate}
\item $(n+m)S=nS+mS$
\item $(nm)S=n(mS)$
\item $n(S+S')=nS+nS'$
\item $1S=S$
\end{enumerate}
If the reader is interested, a deepest analisis of this structure of multisets is avaible on the mencioned work.
Following this concept closer, we will observe that if we have a graph $G=(V,E)$ the definition of graph sum extend not only to the edges but to the vertexes.Now, if we consider the definition on Wildbergers work, then we will have the folowing definition.
\begin{definition}
Lets say that an \textbf{arithmetical graph} is such graph where V and E are multisets, with V multiset or vertexes and E multiset of the non ordered pairs of vertexes or edges.
\end{definition}
Now, lets continue with the change of our previous algebra.
\begin{definition}
A set $\Pi$ with the operations $\dashv \mid \vdash$ and $\mathring{\Lambda}$ is called \textbf{Arithmetical link}
if is a link algebra, but where axiom (4) is $v \dashv \mid \vdash \phi=v, v \mathring{\Lambda} \phi = \phi$, and both operations are not idempotent, with an external operation (multiplication) on $\mathbb{Z}$ (is linear).
\end{definition}
the theorems remain equal, the only variation of this algebra is the introduction of the concept fo Vertex and Edge Engrosure , which includes the properties of numbers in graphs.
\begin{definition}
we will call the relation $\dot{\prec}$ \textbf{Engrosure}, where $\forall n,m \in \mathbb{Z}$ and v $\in \Pi$ we have $n(v) \dot{\prec} m(v)$ if $n < m$.
\end{definition}
is not too difficult to see that if $(\Pi, \dashv \mid \vdash)$ is a group, $(\Pi, \dot{\prec})$ is an equivalence relation. The reader can easily prove the reflexivity, symmetry and transitivity.
\begin{definition}
we will call \textbf{Inverse vertex} to
$$(-1)v := \tilde{v}.$$
\end{definition}
\begin{definition}
We will call a \textbf{Loop} to
$$v^\sigma := v \mathring{\Lambda} v.$$
\end{definition}
there is a theorem we want to add to this structure, probably it is obvious to demonstrate, but it is worth to be written.
\begin{theorem}
For an edge $e=v \mathring{\Lambda} w$ with $n \in \mathbb{Z}$
$$ne=(nv) \mathring{\Lambda} w = v \mathring{\Lambda} (nw).$$
\end{theorem}
\begin{proof}
we have
\begin{tabular}{llll}
& $n(v \mathring{\Lambda} w)$ & $= v \mathring{\Lambda} w \dashv \mid \vdash \dots \dashv \mid \vdash v \mathring{\Lambda} w$ \\
& {} & $=v \mathring{\Lambda} (w \dashv \mid \vdash \dots \dashv \mid \vdash w)$ \\
& {} & $=v \mathring{\Lambda} (nv)$ \\
\end{tabular}
\\
the other case is analogous.
\end{proof}
\begin{example}
Let's say we have the multigraph
$$\Pi = 2(v_1 \mathring{\Lambda} v_2) \dashv \mid \vdash 2(v_2 \mathring{\Lambda} v_3) \dashv \mid \vdash v_4 \mathring{\Lambda} (v_1\dashv \mid \vdash v_2 \dashv \mid \vdash v_3).$$
Using the linearity he have:
$$\Pi = 2(v_1 \mathring{\Lambda} v_2 \dashv \mid \vdash v_2 \mathring{\Lambda} v_3) \dashv \mid \vdash v_4 \mathring{\Lambda} (v_1\dashv \mid \vdash v_2 \dashv \mid \vdash v_3),$$
Using the defintion of Path and star:
$$\Pi = 2(P_3) \dashv \mid \vdash S_3.$$
This example is an homage to Bernhard Euler, to his famous solution of the Koenisberg bridge problem presented on his work of 1736 Solutio problematis ad geometriam situs pertinentis.
\end{example}
\subsection{Oriented and mixed graphs} This subsection will be briefest than others, the reason is that the definition introduced is short.
As we said in the subsection \ref{subsect}, an oriented graph is defined as a graph $G=(V,E)$ where E is an ordered pair of vertexes. Therefore, the only thing we will define before to show the definition of oriented link algebra, is the following
\begin{definition}
Let's have (x,y) with Kuratowski's defintion
$$(x,y):=\{\{x\},\{x,y\}\},$$
We will call \textbf{Oriented Twist} to
$$(x,y)^\circlearrowleft=(y,x).$$
\end{definition}
A more precise definition with Set/antiset theory would be
$$(x,y)^\circlearrowleft=(x,y) \cup \{\tilde{\{x\}},\{y\}\},$$
where $\{x\}=A$ and $\tilde{\{x\}}= \tilde{A}$.
Now, lets define An Oriented Link Algebra
\begin{definition}
An \textbf{Oriented Link Algebra} is a set $\varGamma$ with operations $\dashv \vdash$, $\vec{\Lambda}$ and ${}^\circlearrowleft$ (unitary), that satisfies the axioms of a Link Algebra except commutativity for the operation $\vec{\Lambda}$ satisfying instead
\begin{enumerate}
\item $v \vec{\Lambda} w = w \overleftarrow{\Lambda} v$
\item $(v \vec{\Lambda} w)^\circlearrowleft=v \overleftarrow{\Lambda} w$
\end{enumerate}
Basically, it works same as a link algebra, with the same theorems and definitions.
\end{definition}
For a mixed graphs the definition is as follow
\begin{definition}
An \textbf{Mixed Link Algebra} is a set $\varGamma$ with the operations $\dashv \vdash$, $\dot{\Lambda}$, $\vec{\Lambda}$, ${}^\circlearrowleft$ where
\begin{enumerate}
\item ($\varGamma$, $\dashv \vdash$, $\dot{\Lambda}$) is a Link Algebra.
\item ($\varGamma$, $\dashv \vdash$, $\vec{\Lambda}$, ${}^\circlearrowleft$) is a Oriented Link Algebra.
\end{enumerate}
\end{definition}
To finish the exposition of this concept we will give two last examples.
\begin{example}
Let's have the oriented graph
$$\Gamma = w_1 \vec{\Lambda} v_1 \dashv \vdash v_1 \vec{\Lambda} v_2 \dashv \vdash v_2 \vec{\Lambda} v_3 \dashv \vdash v_1 \overleftarrow{\Lambda} v_4 \dashv \vdash v_4 \overleftarrow{\Lambda} v_3$$
using the concept of antisimetric conmutativity
$$\Gamma = w_1 \vec{\Lambda} v_1 \dashv \vdash v_1 \vec{\Lambda} v_2 \dashv \vdash v_2 \vec{\Lambda} v_3 \dashv \vdash v_4 \vec{\Lambda} v_1 \dashv \vdash v_3 \vec{\Lambda} v_4$$
\end{example}
\begin{example}
Let's have the mixed graph
$$\Gamma = v_1 \vec{\Lambda} v_2 \dashv \vdash v_2 \vec{\Lambda} v_1 \dashv \vdash v_2 \dot{\Lambda} v_3.$$
Where we can identify a circular vinculation and a simple vinculation.
\end{example}
\subsection{As an Epilogue} For almost 200 years, graphs had a minor status in the world of mathmatics, being just the shadow of combinatorics, being only consulted when combinatorial arithmetic fails. George Boole, the father of modern logic, once said that in a near future, mathematic will be not mathmatics of numbers, but of abstract objects beyond numbers, opening the view to a new world where there is no center in the mathematical universe: just objects, operations, lattices and who knows what other concepts to be defined and discover.
Boole as same as Sassure (father of modern semiotics), observe that Mathematical simbols \textbf{are} if there is a meaning that support them, an essence that allow to work the abstraction as it were concrete, like the Demiurgos of Plato's Timeo, who commiserating of our material universe, decided to printed material copies from the universe of ideas. \textbf{Demiurgos} means \textbf{Artisan} in greek, and that is the essence of the work of a mathematitian: be the artisan of a never ending sculpture, the diachronically undefined but synchronously defined mathematical Rodin's Dante: thinking of the \textit{commedia della ragione} in the top of the gate of hell.
Le Graphe est mort, vive le Graphe !
|
\section{Introduction}
Let
$$
P=\{(\xi,\eta)\mid \xi\geq0,\eta\geq0\}-\{(0,0)\}.$$
A nonsingular flow $\{\Phi^t\}$ on $P$ defined by
$$
\Phi^t(\xi,\eta)=(e^t\xi,e^{-t}\eta)$$
is called the {\em standard Reeb flow}. In this note the oriented
foliation $\RR$ whose leaves are the orbits of $\{\Phi^t\}$
with the orientation given by the time direction
is called the {\em Reeb foliation}. A continuous flow on $P$
with orbit foliation
$\RR$ is called an $\RR$-flow.
The topological conjugacy classes of $\RR$-flows $\{\phi^t\}$
are classified in \cite{L}
in the following way. Let
$\gamma_1:[0,\infty)\to P$ (resp.\ $\gamma_2:[0,\infty)\to P$)
be a continuous path such that $\gamma_1(0)\in\{\xi=0\}$
(resp.\ $\gamma_2(0)\in\{\eta=0\}$) which intersects every interior leaf of $\RR$
at exactly one point.
Then one can define a continuous function
$$f_{\{\phi^t\},\gamma_1,\gamma_2}:(0,\infty)\to \R$$
by setting that $f_{\{\phi^t\},\gamma_1,\gamma_2}(x)$
is the time needed for the flow $\{\phi^t\}$ to
move from the point $\gamma_1(x)$ until it
reaches a point on the curve $\gamma_2$.
Then $f_{\{\phi^t\},\gamma_1,\gamma_2}$ belongs to the
following space
$$E=\{f:(0,\infty)\to\R\mid f\ \ \mbox{is continuous and}
\ \ \lim_{x\to 0}f(x)=\infty\}.$$
Of course $f_{\{\phi^t\},\gamma_1,\gamma_2}$ depends upon the
choices of $\gamma_1$ and $\gamma_2$. There are two umbiguities,
one coming from the parametrization of $\gamma_1$, and the other
coming from the positions of $\gamma_1$ and $\gamma_2$.
Let $H$ be the space of homemorphisms of $[0,\infty)$ and $C$
the space of continuous functions on $[0,\infty)$.
Define an equivalence relation $\sim$ on $E$ by
$$
f\sim f'\Longleftrightarrow f'=f\circ h+k,\ \ \exists h\in H,\
\ \exists k\in C.$$
Then clearly the equivalence class of $f_{\{\phi^t\},\gamma_1,\gamma_2}$
does not depend on the choice of $\gamma_1$ and $\gamma_2$. Moreover
it is an invariant of the topological conjugacy classes of
$\RR$-flows. Thus if we denote
by $\mathcal E$ the set of the topological conjugacy classes of
the $\RR$-flows, then there is a well defined map
$$
\iota:\mathcal E\to E/\sim.$$
The main result of \cite{L} states that
$\iota$ is a bijection. In particular any $f\in E$ is obtained
as $f=f_{\{\phi^t\},\gamma_1,\gamma_2}$ for some $\RR$-flow $\{\phi^t\}$
and paths $\gamma_i$.
Clearly any strictly monotone function of $E$ belongs to a single
equivalence class, and this corresponds to the standard Reeb
flow $\{\Phi^t\}$. The purpose of this note is to show the following
characterization of the standard Reeb flow.
\begin{Theorem} \label{t}
An $\RR$-flow $\{\phi^t\}$ is topologically conjugate to
the standard Reeb flow $\{\Phi^t\}$ if and only if $\{\phi^{\lambda t}\}$
is topologically conjugate to $\{\phi^t\}$ for any $\lambda>0$.
\end{Theorem}
Of course the only if part is immediate. We shall show the if part
in the next section.
\begin{remark} \label{r10}
A single $\lambda$ is not enough for Theorem \ref{t}.
In fact there is an $\RR$-flow $\{\phi^t\}$ not
topologically conjugate to $\{\Phi^t\}$ such that
$\{\phi^{2t}\}$ is topologically conjugate to $\{\phi^t\}$.
This will be given in Example \ref{example} below.
\end{remark}
The author wishes to express his hearty thanks to the anonymous
referees whose valuable comments are indeed helpful for the improvement
of the paper
\section{Proof of the if part}
The equivalence class of $f\in E$ is determined by how
$f(x)$ oscilates while it tends to $\infty$ as $x\to 0$.
So to measure the degree of oscilation of
$f\in E$, define a nonnegative valued continuous function
$f^*$ defined on $(0,1]$ by
$$
f^*(x)=\max(f\vert_{[x,1]})-f(x).
$$
Then we have the following lemma.
\begin{lemma} \label{easy}
(1) If $\lambda>0$, then $(\lambda f)^*=\lambda f^*$.
(2) If $c$ is a constant, then $(f+c)^*=f^*$.
(3) If $h\in H$, then there is $0<a<1$ such that
$(f\circ h)^*=f^*\circ h$ on $(0,a)$.
(4) If $k\in C$ and $x\to 0$, then $(f+k)^*(x)-f^*(x)\to 0$.
(5) There is a sequence $\{x_n\}$ tending to $0$
such that $f^*(x_n)=0.$
\end{lemma}
{\sc Proof}. Points (1) and (2) are immediate.
To show (3) notice that
\begin{eqnarray*}
&(f\circ h)^*(x)=\max(f\vert_{[h(x),h(1)]})-f(h(x))\ \mbox{ and }\\
&f^*\circ h(x)=\max(f\vert_{[h(x),1]})-f(h(x)).
\end{eqnarray*}
Since $f(x)\to\infty$ ($x\to0$), both maxima coincide for small $x$.
Let us show
(4). By (2) we only need to show (4) assuming that
$k(0)=0$. Now given $\epsilon>0$, there is $\delta>0$
such that if $0<x<\delta$, then $\abs{k(x)}<\epsilon$.
Choose $\eta>0$ small enough so that if $0<x<\eta$,
then we have
$$
f(x)\geq\max(f\vert_{[\delta,1]})\ \ \mbox{and}\ \
(f+k)(x)\geq\max((f+k)\vert_{[\delta,1]}).
$$
This implies that for $x\in(0,\eta)$,
$$
\abs{f^*(x)-(f+k)^*(x)}
\leq\abs{f(x)-(f+k)(x)}+\abs{\max((f+k)\vert_{[x,\delta]})
-\max(f\vert_{[x,\delta]})}<2\epsilon.
$$
This shows (4). Finally (5) follows from the assumption $f(x)\to\infty$
as $x\to0$.
\qed
\bigskip
For $f\in E$ define an invariant $\sigma(f)=\limsup_{x\to 0}f^*(x)$
which takes value in $[0,\infty]$. In fact $\sigma(f)$ coincides
with the invariant $\mathcal A(f)$ defined in \cite{L} and used
to show that $\mathcal E$ is uncountable.
\begin{lemma}\label{sigma}
Assume $f,f'\in E$ and $\lambda>0$.
(1) We have $\sigma(\lambda f)=\lambda\sigma(f)$.
(2) If $f\sim f'$, then $\sigma(f)=\sigma(f')$.
In particular $f$ corresponds to the standard Reeb flow
if and only if $\sigma(f)=0$.
\end{lemma}
{\sc Proof}. Clearly (1) follows from Lemma \ref{easy} (1),
while the first statement of (2) is an easy consequence of Lemma
\ref{easy}
(3) and (4).
To show the last statement, assume $\sigma(f)=0.$ Extend the function
$f^*$ defined on $(0,1]$ to $[0,\infty)$ by letting
$$f^*=0\ \mbox{ on }\ \{0\}\cup(1,\infty).$$
Since $\sigma(f)=0$, $f^*$ is continuous, i.\ e.\ $f^*\in C$.
Thus $f\sim f+f^*$, and the latter is (weakly) monotone near $0$.
Still adding a suitable function, $f$ is seen to be equivalent to
$g$ which is strictly monotone on the whole $(0,\infty)$ such that
$g(x)\to0$ ($x\to\infty$).
Clearly such functions are mutully equivalent by
a pre-composition of some $h\in H$, and correspond to the standard
Reeb flow $\{\Phi^t\}$.
\qed
\bigskip
Now since
\begin{equation} \label{e101}
f_{\{\phi^{\lambda t}\},\gamma_1,\gamma_2}=
\lambda^{-1} f_{\{\phi^t\},\gamma_1,\gamma_2},
\end{equation}
for $\lambda>0$,
Theorem \ref{t} reduces to the following proposition.
\begin{proposition} \label{p}
If $f\in E$ and $f\sim\lambda f$ for any $\lambda>0$, then $\sigma(f)=0$.
\end{proposition}
The rest of the paper is devoted to the proof of Proposition \ref{p}.
But before starting, let us mention an example
for Remark \ref{r10}.
\begin{example} \label{example}
By (\ref{e101}) and the main result of \cite{L},
it suffices to construct a function $f\in E$ such that $f(x/2)=2f(x)$ and
that $\sigma(f)=\infty$. Set for example
$$
f(x)=\frac{1}{x}2^{\sin(2\pi\log_2x)}.$$
\end{example}
The following lemma, roughly the same thing as
the linearization in one dimensional local
dynamics, plays a crucial role in what follows.
\begin{lemma} \label{key}
Assume $f\in E$ satisfies $\lambda f=f\circ h+k$ for some $h\in H$,
$k\in C$ and $\lambda>1$. Then
$0$ is an attracting fixed point of $h$
and there exists $f_\infty\in E$ such that
$f_\infty-f\in C$, $\lambda f_\infty=f_\infty\circ h$
and $f_\infty(x)\to 0$ ($x\to\infty$).
\end{lemma}
{\sc Proof}.
Any equivalence class of $E$ has a representative $f$
such that
\begin{equation} \label{e1}
f\vert_{[1,\infty)}\ \ \mbox{is bounded}.
\end{equation}
So it is no loss of generality to assume that the function $f$ in
the lemma satisfies (\ref{e1}). We can also assume that $k(0)=0$,
by adding a suitable constant to $f$ if necessary.
Choose $a'\in(0,1)$ so that if $a\in(0,a')$,
$$
f(a)>\frac{2}{\lambda-1}\max(\abs{k}\vert_{[0,1]}).$$
Then we have
\begin{equation} \label{e100}
f\circ h(a)>\frac{\lambda+1}{2}f(a),\ \ \forall a\in(0,a').
\end{equation}
If $a$ is sufficiently
near $0$, we have
$$
f(a)>\sup(f\vert_{[1,\infty)}).$$
If furthermore $f^*(a)=0$, then
$$
\{x\mid f(x)>f(a)\}\subset(0,a).$$
Thus (\ref{e100})
implies
$h(a)<a$ for such $a$.
But this allows us to use (\ref{e100}) repeatedly for $h^n(a)$
($n=1,2,\cdots$) instead of $a$,
showing that $f\circ h^n(a)\to\infty$ as $n\to\infty$.
Clearly this implies that
$[0,a]$ is contained in the attracting domain
of an attractor $0$ of the homeomorphism $h$, showing the first point
of Lemma \ref{key}.
For the rest of the proof, let us divide the argument into
two cases according to the dynamics of $h$.
First assume that the whole line $[0,\infty)$
is the attracting domain of $0$.
Let
$$
f_n(x)=\lambda^{-n}f(h^n(x)).$$
Then we have
$$
f_{n+1}(x)-f_n(x)=-\lambda^{-n-1}k(h^n(x)),$$
showing that $f_n\to f_\infty$ uniformly on compact
subsets of $(0,\infty)$ for some
continuous function $f_\infty$.
Now since
$$\lambda f_{n+1}(x)=f_n(h(x)),$$
we have
$$
\lambda f_\infty=f_\infty\circ h.$$
We also have
$$\abs{f(x)-f_\infty(x)}\leq\sum_
{n=0}^\infty \lambda^{-n-1}\abs{k(h^n(x))}. $$
The continuity of $k$, together with the assumption $k(0)=0$,
implies that
$$\lim_{x\to0}\abs{f(x)-f_{\infty}(x)}=0,$$
showing that $f_\infty-f\in C$.
Finally since $h^{-n}(x)\to\infty$
($n\to\infty$) and
$$f_\infty\circ h^{-n}(x)=\lambda^{-n}f_\infty(x), \ \ \forall
x\in(0,\infty),
$$
we have $f_\infty(x)\to 0$ ($x\to\infty$).
Next assume there is a fixed point $b$ of $h$ such that
$(0,b)$ is an attracting domain of $0$.
Thus we have $h^{-n}(x)\to b$ ($n\to \infty$) for any
$x\in (0,b)$.
The same argument
as above shows the existence of a continuous function
$f_\infty$ on $(0,b)$.
Since
$$
f_\infty\circ h^{-n}(x)=\lambda^{-n}f_\infty(x),\ \ \forall x\in(0,b),
$$
we have
$$
\lim_{x\uparrow b}f_\infty(x)=0.$$
Now extend $f_\infty$ by setting $f_\infty=0$ on $[b,\infty)$.
\qed
\bigskip
Let us start the proof of Proposition \ref{p}.
Assume $f\in E$ satisfies $f\sim 2^{1/N}f$ for any $N\in\N$.
Applying Lemma \ref{key}, $f$ can be changed
within the equivalence class to one which satisfies the condition
of $f_\infty$ for $\lambda=2$.
We also assume for contradiction that
$\sigma(f)>0$.
Thus the proof of Proposition \ref{p} reduces to showing
that there is no $f\in E$ which
satisfies the following assumption.
\begin{assumption} \label{a1}
A function $f\in E$ satisfies
\begin{eqnarray}
2f=f\circ h,\ \ \exists h\in H, \
& f(x)\to 0\ \ (x\to\infty),\label{first}\\
2^{1/N}f-f\circ h_N\in C\ \ & \exists h_N\in H, \label{second}\ \
\forall N\geq2\ \ \mbox{and}\\
\sigma(f)>0. & \label{third}
\end{eqnarray}
\end{assumption}
Define
$$E_0=\{f\in E\mid f(x)\to0\ \ (x\to\infty)\}.$$
Henceforth all the functions dealt with will be in $E_0$,
and the following definition is more convenient.
For $f\in E_0$ define
$$
f^\sharp(x)=\max(f\vert_{[x,\infty)})-f(x).$$
Clearly $f^\sharp$ and $f^*$ are the same near $0$ and Lemma
\ref{easy} (1), (4) and (5) hold also for $f^\sharp$, while (3)
becomes stronger. In summary we have:
\begin{lemma} \label{sharp}
Assume $f, f'\in E_0$.
(1) If $\lambda>0$, then $(\lambda f)^\sharp=\lambda f^\sharp$.
(3) If $h\in H$, then $(f\circ h)^\sharp=f^\sharp\circ h$.
(4) If $f'-f\in C$ and $x\to0$, then $f^\sharp(x)-(f')^\sharp(x)\to0$.
(5) There is a sequence $\{x_n\}$ tending to $0$ such that
$f^\sharp(x_n)=0$.
\qed
\end{lemma}
Hereafter $f$ is always to be a function satisfying Assumption \ref{a1}.
Thus we have
\begin{equation} \label{no}
2f^\sharp=f^\sharp\circ h.
\end{equation}
Fix $N$ for a while and let $h_1=h_N^N$.
Notice that by Lemma \ref{key} both
$h$ and $h_1$ have $0$ as their attractors
and that
$$f\circ h-f\circ h_1=2f-f\circ h_1
=\sum_{\nu=0}^{N-1}2^{\frac{N-\nu-1}{N}}(2^{\frac{1}{N}}f\circ
h_N^\nu
-f\circ h_N^{\nu+1})\in C.$$
The following is an easy corollary of Lemma \ref{sharp}.
\begin{corollary} \label{small}
We have
$$
\lim_{x\to 0}\abs{f^\sharp\circ h(x)-f^\sharp\circ h_1(x)}=0.$$
\qed
\end{corollary}
Our overall strategy is to show that $f^\sharp$ is too much
oscilating in a fundamental domain of $h$, thanks to condition
(\ref{second}).
For that purpose first of all
we have to compare the dynamics of $h$ and $h_1$ near the common
attractor $0$ and to show that they have more or less the same
fundamental domains.
\begin{lemma} \label{final}
Either there exists a sequence $\{a_n\}$ such that
$a_n\to 0$ and that $h^2(a_n)\leq h_1(a_n)\leq h(a_n)$
or there exists a sequence $\{a_n\}$ such that
$a_n\to 0$ and that $h^2_1(a_n)\leq h(a_n)\leq h_1(a_n)$.
\end{lemma}
{\sc Proof}.
If there is a sequence $\{a_n\}$ such that $a_n\to 0$
and that $h(a_{n})=h_1(a_n)$, there is nothing to prove.
So there are two cases to consider.
One is when $h_1(x)<h(x)$ for any small $x$, and the other
$h_1(x)>h(x)$.
For the moment assume the former.
In way of contradiction assume the contrary of the assertion
of the lemma.
This is equivalent to saying that $h_1(x)< h^2(x)$
for {\em any small} $x$.
For small $x$,
let $y=y(x)\in[h_1(x),x]$ be any point which gives
$\max(f^\sharp\vert_{[h_1(x),x]})$.
Notice that $f^\sharp(y)$ can be as large as we wish by choosing
$x$ even smaller.
Then since $f^\sharp(h^2(y))=4f^\sharp(y)>f^\sharp(y)$,
the point $h^2(y)$ is contained in
$$[h^2\circ h_1(x),h^2(x)]- (h_1(x),x]=[h^2\circ h_1(x),h_1(x)]
\subset [h^2_1(x),h_1(x)].
$$
The last inclusion follows from the assumption for a contradiction.
Put $h^2(y)=h_1(z)$ for some $z=z(x)\in[h_1(x),x]$. Then we
have
\begin{equation} \label{e31}
f^\sharp\circ h_1(z)=4f^\sharp(y)\geq 4f^\sharp(z)\ \ \mbox{ and }\ \
f^\sharp\circ h(z)=2f^\sharp(z).
\end{equation}
If we choose $x$ near enough to $0$, then
the associated $z=z(x)$ is also near, and thus
$$\abs{2f^\sharp(z)
-f^\sharp\circ h_1(z)}
=\abs{f^\sharp\circ h(z)
-f^\sharp\circ h_1(z)}
$$
can be arbitrarily small by Corollary \ref{small}.
Then we have
$$f^\sharp(z)\approx \frac{1}{2}f^\sharp\circ h_1(z)=
2f^\sharp(y)\geq 1
$$
for any such $z=z(x)$.
On the other hand $z(x)$ can be arbitrarily near to $0$, and
thus (\ref{e31}) contradicts Corollary \ref{small}.
The opposite case where $h(x)<h_1(x)$ for any small $x$
can be dealt with similarly by considering $f'\in E$,
equivalent to $f$, such that $2f'=f'\circ h_1$ and $f'(x)\to0$
($x\to\infty$), instead of $f$.
\qed
\bigskip
Now fix a large number $N$ and choose $f_1\in E_0$ such that
$$f_1-f\in C,\ \ 2^{1/N}f_1=f_1\circ h_N.
$$
The existence of such $f_1$ is
guaranteed by Lemma \ref{key} applied to $\lambda=2^{1/N}$.
We have then
\begin{equation}\label{yes}
2^{1/N}f_1^\sharp=f_1^\sharp\circ h_N.
\end{equation}
Together with Lemma \ref{final} which asserts that
the fundamental domain of $h_N^N$ is more or
less comparable with that of $h$, this implies that $f_1^\sharp$ is oscilating
in an extremely high frequency for $N$ big. We are going to get
a contradiction from this.
We still assume (\ref{first}) for $f$.
According to Lemma \ref{final}, there are two cases to consider.
One is when there is a sequence $a_n\to 0$ such that
$h^2(a_n)\leq h^N_N(a_n)\leq h(a_n)$, the other being
$h^{2N}_N(a_n)\leq h(a_n)\leq h^N_N(a_n)$.
Assume for the moment that the former holds for infinitely many $N$.
Let $x_n^1$ be the largest point such that $x^1_n\leq a_n$ and
$f_1^\sharp(x_n^1)=0$.
Notice that by Lemma \ref{sharp} (5) and the equation (\ref{yes}), we have
\begin{equation}\label{24}
x_n^1\in(h_N(a_n),a_n].
\end{equation}
Then again by (\ref{yes})
$f_1^\sharp$ vanishes at the points $x_n^\nu=h_N^{\nu-1}(x^1_n)$
for any $1\leq \nu\leq N$. Let
$y_n^1$ be any point in $[x_n^2,x_n^1]$ at which $f_1^\sharp$
takes the maximal value
and let $y_n^\nu=h_N^{\nu-1}(y^1_n)$
for $1\leq \nu \leq N-1$.
By (\ref{24}) the order of these points are as follows.
$$
h^2(a_n)<h_N^N(a_n)\leq x^{N}_n < y^{N-1}_n<\cdots<y_n^\nu<x_n^\nu<\cdots
<y^1_n<x^1_n\leq a_n.$$
Notice that $y_n^\nu$ is a point in $[x_n^{\nu+1},x_n^\nu]$
at which $f_1^\sharp$ takes the maximal value,
and
$$f_1^\sharp(y_n^\nu)=2^{(\nu-1)/N}f_1^\sharp(y_n^1).$$
We also have
\begin{equation}\label{1/2}
f_1^\sharp(y_n^\nu)\geq \frac{1}{2}
\max(f_1^\sharp\vert_{[h_N^N(a_n),a_n]}).
\end{equation}
In fact on one hand
$$\max(f_1^\sharp\vert_{[x_n^N,a_n]})
=f_1^\sharp(y_n^{N-1})= 2^{(N-2)/N}f_1^\sharp(y_n^1)\leq
2f_1^\sharp(y_n^1).
$$
On the other hand
$$\max(f_1^\sharp\vert_{[h_N^N(a_n),x_n^N]})\leq 2^{(N-1)/N}\max
(f_1^\sharp\vert_{[x_n^2,x_n^1]})
\leq 2f_1^\sharp(y_n^1),$$
because
$$h_N^{-N+1}[h_N^N(a_n)), x_n^N]=[h_N(a_n),x_n^1]
\subset [x_n^2,x_n^1].
$$
Henceforth we focus our attention to the other homeomorphism
$h\in H$.
There is a sequence $\{m_n\}$ of integers such that
the points $h^{-m_n}(a_n)$ belong to a fixed fundamental domain
in the basin of $0$ for $h$. Notice that $m_n\to\infty$ since
$a_n\to0$.
Passing to a subsequence if necessary, we may assume that
$$h^{-m_n}(a_n)\to a,\ \ \
h^{-m_n}(x^\nu_n)\to x^\nu\ \mbox{and}\ h^{-m_n}(y^\nu_n)\to y^\nu,$$
for some points $a$, $x^\nu$ and $y^\nu$.
There
is an ordering
$$h^2(a)\leq x^{N} \leq y^{N-1}\leq\cdots \leq y^\nu\leq x^\nu\leq\cdots
\leq y^1\leq x^1\leq a.
$$
We shall show that $f^\sharp(x^\nu)=0$ and that $f^\sharp(y^\nu)$ is bounded
away from 0 with a bound {\em independent of $N$}.
Since these points can be taken in the same compact interval $[h^2(a),a]$,
this will contradict the continuity of $f^\sharp$.
By Lemma \ref{sharp} (4), $f_1^\sharp(x_n^\nu)=0$
implies $f^\sharp(x_n^\nu)\leq 1$ for any
large $n$. Therefore by(\ref{no})
$$
f^\sharp(h^{-m_n}(x_n^\nu))\leq 2^{-m_n},$$
showing that $f^\sharp(x^\nu)=0.$
On the other hand since $h_N^N(a_n)\leq h(a_n)$, we have by (\ref{1/2})
$$
f_1^\sharp(y_n^\nu)\geq \frac{1}{2}\max(f_1^\sharp\vert_{[h_N^N(a_n),a_n]})
\geq \frac{1}{2}\max(f_1^\sharp\vert_{[h(a_n),a_n]}),
$$
and therefore again by Lemma \ref{sharp} (4), for any large $n$,
$$f^\sharp(y_n^\nu)\geq\frac{1}{2}\max(f^\sharp\vert_{[h(a_n),a_n]})-1.$$
Let $M=\max(f^\sharp\vert_{[h(a),a]})$ and notice that $M>0$ since
$\sigma(f)>0$ (\ref{third}) and by (\ref{no}).
For any large $n$, the interval $h^{-m_n}[h(a_n),a_n]$
is near $[h(a),a]$, and is composed of a subinterval of $[h(a),a]$
and the iterate by $h^{\pm1}$ of the complementary subinterval, and
therefore
$$
\max(f^\sharp\vert_{h^{-m_n}[h(a_n),a_n]})\geq M/2.$$
This implies by (\ref{no})
$$
\max(f^\sharp\vert_{[h(a_n),a_n]})\geq \frac{1}{2}M2^{m_n},$$
showing that for any large $n$
$$f^\sharp(y_n^\nu)\geq\frac{1}{4}M2^{m_n}-1.$$
This concludes that
$$f^\sharp(y^\nu)\geq\frac{1}{4}M,$$
as is desired.
The opposite case where
$h^{2N}_N(a_n)\leq h(a_n)\leq h_N^N(a_n)$ ($\exists a_n\to 0$)
holds for infinitely many $N$
can be dealt with in a similar way, although the argument is not
completely symmetric.
|
\section{Introduction}
The theory of shadowing of approximate trajectories
(pseudotrajectories) of dynamical systems is now a well developed
part of the global theory of dynamical systems (see, for example,
the monographs \cite{Palmer1}, \cite{Pilyugin}). This theory is
closely related to the classical theory of structural stability (the
basic definitions of structural stability and $\Omega$-stability for
flows can be found, for example, in the monograph \cite{PilSSBook}).
It is well known that a diffeomorphism has the shadowing property in
a neighborhood of a hyperbolic set \cite{Anosov}, \cite{Bowen} and a
structurally stable diffeomorphism has the shadowing property on the
whole manifold \cite{Robinson}, \cite{Morimoto}, \cite{Sawada}.
Analyzing the proofs of the first shadowing results by Anosov
\cite{Anosov} and Bowen \cite{Bowen}, it is easy to see that, in a
neighborhood of a hyperbolic set, the shadowing property is
Lipschitz (and the same holds in the case of a structurally stable
diffeomorphism, see \cite{Pilyugin}). At the same time, it is easy
to give an example of a diffeomorphism that is not structurally
stable but has the shadowing property (see \cite{Pilyugin2}, for
example). Thus, structural stability is not equivalent to shadowing.
However it was shown in \cite{Pilyugin1} that structural stability
of a diffeomorphism is equivalent to Lipschitz shadowing.
Turning to flows, it is well known that a flow has the shadowing
property in a neighborhood of a hyperbolic set \cite{Palmer1},
\cite{Pilyugin} and a structurally stable flow has the shadowing
property on the whole manifold \cite{Pilyugin}, \cite{Pilyugin4}. In
fact, in a neighborhood of a hyperbolic set, the shadowing property
is Lipschitz and the same holds in the case of a structurally stable
flow, see \cite{Pilyugin}. At the same time, it is easy to give an
example of a flow that is not structurally stable but has the
shadowing property (to construct such an example, one can use almost
the same idea as in \cite{Pilyugin2}). Thus, as with
diffeomorphisms, structural stability is not equivalent to
shadowing. However it is our purpose in this article to show that
structural stability of a flow is equivalent to Lipschitz shadowing.
Let us note that the proof for the flow case is a nontrivial
modification of the proof for the diffeomorphism case.
One of the previously used approaches to compare shadowing property
and structural stability is passing to $C^1-$interiors. Sakai
\cite{Sak} showed that the $C^1-$interior of the set of
diffeomorphisms with the shadowing property coincides with the set
of structurally stable diffeomorphisms. See also \cite{PilRodSak}
for the generalization of this result to other types of shadowing
properties. For vector fields the situation is different. There is
an example of a vector field with the robust shadowing property
which is not structurally stable \cite{PilTikh}. See also
\cite{LeeSak,PilTikh2008,Tikh} for some positive results in this
direction.
In this paper, we also study vector fields having the Lipschitz
periodic shadowing property. Diffeomorphisms having the Lipschitz
periodic shadowing property were studied in \cite{OPT}, where it was
shown that this property is equivalent to $\Omega$-stability. We
prove a similar statement for vector fields.
\section{Preliminaries}
Let $M$ be a smooth closed manifold with Riemannian metric
dist$(\cdot,\cdot)$ and let $X$ be a vector field on $M$ of class
$C^1$. Let $\phi(t,x)$ be the flow on $M$ generated by $X$.
\medskip
\begin{defin} A (not necessarily continuous) function
$y:I\to M$ (where $I$ is an interval in $\mathds{R}$) is called a
$d$-pseudotrajectory if
$$
\dist(y(\tau+t),\phi(t,y(\tau))\leq d,\quad 0\leq t \leq 1, \quad
\tau, \tau+t\in I.
$$
Mostly we work with pseudotrajectories defined on $I=\mathds{R}$.
\end{defin}
\begin{defin} We say that the vector field $X$ has the {\it Lipschitz shadowing property}
($X \in \LipSh$) if there exist $d_0$ and ${\cal L}>0$ such that if
$y:\mathds{R}\mapsto M$ is a $d$-{\it pseudotrajectory} for $d\le d_0$,
then $y(t)$ is ${\cal L}d$-{\it shadowed by a trajectory}, that is,
there exists a trajectory $x(t)$ of $X$ and an increasing
homeomorphism (\textit{reparametrization}) $\alpha(t)$ of the real
line satisfying
\begin{equation}
\label{eqdef2.1} \alpha(0) = 0, \quad
\left|\frac{\alpha(t_2)-\alpha(t_1)}{t_2-t_1}-1\right|\le {\cal L}d
\end{equation}
for $t_2\neq t_1$ and
\begin{equation}\label{eqdef2.2}
\dist(y(t),x(\alpha(t))\le {\cal L}d
\end{equation}
for all $t$.
\end{defin}
\begin{defin} We say that the vector field $X$ has the
\textit{Lipschitz periodic shadowing property} ($X \in \LipPerSh$)
if there exist $d_0$ and ${\cal L} > 0$ such that if $y:\mathds{R}\mapsto
M$ is a periodic $d$-{\it pseudotrajectory} for $d\le d_0$, then
$y(t)$ is ${\cal L}d-${\it shadowed by a periodic trajectory}, that
is, there exists a trajectory $x(t)$ of $X$ and an increasing
homeomorphism $\alpha(t)$ of the real line satisfying inequalities
\sref{eqdef2.1} and \sref{eqdef2.2} and such that
\begin{equation}
\notag x(t+\omega)=x(t)
\end{equation}
for some $\omega > 0$.
The last equality implies that $x(t)$ is either a closed trajectory
or a rest point of the flow $\phi$.
\end{defin}
The main results of the paper are the following theorems.
\begin{thm}
\label{thmLipSh} A vector field $X$ satisfies the Lipschitz
shadowing property if and only if $X$ is structurally stable.
\end{thm}
\begin{thm}
\label{thmLipPerSh} A vector field $X$ satisfies the Lipschitz
periodic shadowing property if and only if $X$ is $\Omega$-stable.
\end{thm}
It is known that expansive diffeomorphisms having the Lipschitz
shadowing property are Anosov (see \cite{Pilyugin1}).
We show, as a consequence of Theorem \ref{thmLipSh}, that expansive
vector fields having the Lipschitz shadowing property are Anosov.
Let us recall the definition of expansivity for vector fields.
\begin{defin} We say that a vector field $X$
and the corresponding flow $\phi(t,x)$ are \textit{expansive} if
there exist constants $a, \delta > 0$ such that if
$$
\mbox{dist}(\phi(t, x), \phi(\alpha(t), y)) < a, \quad t \in\mathds{R},
$$
for points $x, y \in M$ and an increasing homeomorphism $\alpha$ of the
real line, then $y = \phi(\tau, x)$ for some $|\tau| < \delta$.
\end{defin}
\begin{thm}
An expansive vector field $X$ having the Lipschitz shadowing
property is Anosov.
\end{thm}
\begin{proof} By Theorem \ref{thmLipSh}, a vector field $X$ having the Lipschitz
shadowing property is structurally stable. Hence, there exists a
neighborhood ${\cal N}$ of $X$ in the $C^1$-topology such that any
vector field in ${\cal N}$ is expansive (this property of $X$ is
sometimes called robust expansivity).
By Theorem B of \cite{MorSakSun}, robustly expansive vector fields
having the shadowing property are Anosov.
\end{proof}
In Sec.~\ref{secLipSh} we prove Theorem~\ref{thmLipSh} and in
Sec.~\ref{secLipPerSh} we prove Theorem~\ref{thmLipPerSh}. Both
proofs are long so that each section is divided into several
subsections.
\section{The Lipschitz shadowing property}\label{secLipSh}
As was mentioned above in \cite{Pilyugin4} it was proved that
structurally stable vector fields have the Lipschitz shadowing
property. Our goal here is to show that vector fields satisfying
Lipschitz shadowing are structurally stable. It is well known (see
\cite{Robinson1}) that for this purpose it is enough to show that
such a vector field satisfies Axiom A$'$ and the strong
transversality condition.
First we show that Lipschitz shadowing implies discrete Lipschitz
shadowing. Define a diffeomorphism $f$ on $M$ by setting
$f(x)=\phi(1,x)$.
\medskip
\begin{defin}\label{defDLSh} The vector field $X$ has the {\it discrete Lipschitz shadowing
property} if there exist $d_0$, $L>0$ such that if $y_k\in M$ is a
sequence with
$$ \dist(y_{k+1},f(y_k)) \le d, \quad k\in \mathds{Z}$$
for $d\le d_0$, then there exist sequences $x_k\in M$ and $t_k\in
\mathds{R}$ satisfying
$$ |t_k-1|\le Ld,\quad \dist(x_k, y_k) \le Ld, \quad x_{k+1}=\phi(t_k,x_k)$$
for all $k$.
\end{defin}
\medskip
\begin{lem}\label{lemPalm1} Lipschitz shadowing implies discrete Lipschitz
shadowing.
\end{lem}
\begin{proof} Let $y_k$ be a sequence with
$$ {\rm dist}(y_{k+1},f(y_k))={\rm dist}(y_{k+1},\phi(1,y_k))\le d, \quad k\in \mathds{Z}.$$
Then we define
$$ y(t)=\phi(t-k,y_k)\quad k\le t<k+1,\quad k\in \mathds{Z}.$$
Assume that $k\le\tau<k+1$. If $0\le t\le 1$ and $\tau+t<k+1$, then
$$ {\rm dist}(y(\tau+t),\phi(t,y(\tau))
={\rm dist}(\phi(\tau+t-k,y_k),\phi(t,\phi(\tau-k,y_k)))=0$$
and if $k+1\le \tau+t$, then
$$ \begin{array}{rl}
&{\rm dist}(y(\tau+t),\phi(t,y(\tau)))\\
&={\rm dist}(\phi(\tau+t-k-1,y_{k+1}),\phi(t+\tau-k,y_k))\\
&={\rm dist}(\phi(\tau+t-k-1,y_{k+1}),\phi(\tau+t-k-1,\phi(1,y_k)))\\
&\le \nu d,\end{array}$$
where $\nu$ is a constant such that
\begin{equation}\label{new5}{\rm dist}(\phi(t,x),\phi(t,y))\le \nu\,{\rm dist}(x,y)\quad{\rm for}\quad
x,y \in M,\; 0\le t\le 1.\end{equation}
Then if $d\le d_0/\nu$, there exists a trajectory
$x(t)$ of $X$ and a function $\alpha(t)$ satisfying
$$ \left|\frac{\alpha(t_2)-\alpha(t_1)}{t_2-t_1}-1\right|\le {\cal L}\nu d $$
for $t_2\neq t_1$ and
$$ {\rm dist}(y(t),x(\alpha(t)))\le {\cal L}\nu d$$
for all $t$. Then if we define
$$ x_k=x(\alpha(k)),\quad t_k=\alpha(k+1)-\alpha(k),$$
we see that
$$ x_{k+1}=x(\alpha(k+1))=\phi(\alpha(k+1)-\alpha(k),x(\alpha(k)))
=\phi(t_k,x_k),$$
$$ \dist(x_k, y_k)=\dist(x(\alpha(k)), y(k)) \le {\cal L}\nu d$$
and
$$ |t_k-1|=\left|\frac{\alpha(k+1)-\alpha(k)}{k+1-k}-1\right|
\le {\cal L}\nu d.$$
Taking $L = {\cal L}\nu$ and $d_0$ in Definition \ref{defDLSh} as
$d_0/\nu$, we complete the proof of the lemma.
\end{proof}
Our main tool in the proof is the following lemma which
relates the shadowing problem to the problem of existence
of bounded solutions of certain difference equations.
To ``linearize'' our problem, we apply the standard technique of exponential mappings.
Denote by $T_xM$ the tangent space to $M$ at a point $x$; let $|v|$ be the norm of $v$ corresponding to the metric dist$(\cdot,\cdot)$.
Let $\exp:TM\mapsto M$ be the standard exponential mapping on the
tangent bundle of $M$ and let $\exp_x$ be the corresponding mapping
$T_xM\mapsto M$.
Denote by $B(r,x)$ the ball in $M$ of radius $r$ centered at a point
$x$ and by $B_T(r,x)$ the ball in $T_xM$ of radius $r$ centered at
the origin.
There exists $r>0$ such that, for any $x\in M$, $\exp_x$ is a
diffeomorphism of $B_T(r,x)$ onto its image, and $\exp_x^{-1}$ is a
diffeomorphism of $B(r,x)$ onto its image. In addition, we may
assume that $r$ has the following property:
If $v$, $w\in B_T(r,x)$, then
\begin{equation}\label{new3}
{\rm dist}(\exp_x(v), \exp_x(w)) \le 2 |v-w|;
\end{equation}
if $y$, $z\in B(r,x)$, then
\begin{equation}\label{new4}
|\exp_x^{-1}(y)-\exp_x^{-1}(z)| \le 2 \dist(y, z).
\end{equation}
Let $x(t)$ be a trajectory of $X$; set $p_k = x(k)$ for $k \in \mathds{Z}$.
Denote $A_k=Df(p_k)$ and ${\cal M}_k=T_{p_k}M$. Clearly, $A_k$ is a
linear isomorphism between ${\cal M}_k$ and ${\cal M}_{k+1}$.
In the sequel whenever we construct $d$-pseudotrajectories of the
diffeomorphism $f$, we always take $d$ so small that the points of
the pseudotrajectories under consideration, the points of the
associated shadowing trajectories, their lifts to tangent spaces,
etc. belong to the corresponding balls $B(r,p_k)$ and $B_T(r,p_k)$.
We consider the mappings
\begin{equation}\label{new1}
F_k=\exp_{p_{k+1}}^{-1}\circ f\circ \exp_{p_{k}}: B_T(\rho,p_k)\to
{\cal M}_{k+1}
\end{equation}
with $\rho\in(0,r)$ small enough, so that
$$
f\circ \exp_{p_{k}}(B_T(\rho,p_k))\subset B(r,p_{k+1}).
$$
It follows from standard properties of the exponential mapping that
$D\exp_x(0)= \Id$; hence,
$$ DF_k(0)=A_k.$$
Since $M$ is compact, for any $\mu>0$ we can find
$\delta=\delta(\mu) >0$ such that if $|v|\le\delta$, then
\begin{equation}\label{new2}|F_k(v)-A_kv|\le\mu|v|.
\end{equation}
\begin{lem}\label{lemPalm2} {\it Assume that $X$ has the discrete Lipschitz shadowing
property with constant $L$. Let $x(t)$ be an arbitrary trajectory of
$X$, let $p_k = x(k)$, $A_k=Df(p_k)$ and let $b_k\in{\cal M}_k$ be a
bounded sequence (denote $b=\|b\|_{\infty}$). Then there exists a
sequence $s_k$ of scalars with $|s_k|\le b'=L(2b+1)$ such that the
difference equation
$$ v_{k+1}=A_kv_k+X(p_{k+1})s_k+ b_{k+1} $$
has a solution $v_k$ such that}
$$ \|v\|_{\infty}\le 2b'.$$
\end{lem}
\begin{proof} Fix a natural number $N$ and define $\Delta_k\in{\cal M}_k$ as
the solution of
$$ v_{k+1}=A_kv_k+b_{k+1},\quad k=-N,\ldots,N-1 $$
with $\Delta_{-N}=0$. Then
\begin{equation}\label{4} |\Delta_k|\le C,\end{equation}
where $C$ depends on $N$, $b$ and an upper bound on $|A_k|$.
Fix a small number $d>0$ and fix $\mu$ in (\ref{new2}) so that $\mu
< 1/(2C)$. Then consider the sequence of points $y_k\in M$, $k\in
\mathds{Z}$, defined as follows: $y_k = p_k$ for $k\le -N$, $y_k =
\exp_{p_k}(d\Delta_k)$ for $-N + 1\le k \le N$, and $y_{N +k} =
f^k(y_N)$ for $k > 0$.
By definition, $y_{k+1} = f(y_k)$ for $k\le -N$ and $k\ge N$. If
$-N-1 \le k \le N - 1$, then
$$y_{k+1} = \exp_{p_{k+1}}(d\Delta_{k+1})= \exp_{p_{k+1}}(dA_k\Delta_k + db_{k+1}),$$
and it follows from estimate (\ref{new3}) that if $d$ is small enough, then
\begin{equation}\label{6} {\rm dist}\left(y_{k+1}, \exp_{p_{k+1}}(dA_k\Delta_k )\right) \le 2d|b_{k+1}| \le 2db. \end{equation}
On the other hand,
$$ f(y_k) = \exp_{p_{k+1}}(F_k(d\Delta_k ))$$
(see the definition (\ref{new1}) of the mapping $F_k$), and we deduce from (\ref{new3}),
(\ref{new2}) and (\ref{4}) that if $Cd\le\delta(\mu)$
\begin{equation}\label{7} \begin{array}{rl}
{\rm dist}\left(f(y_k), \exp_{p_{k+1}}(dA_k\Delta_k)\right)
&\le 2|F_k(d\Delta_k)-dA_k\Delta_k|\\
&\le 2\mu|d\Delta_k|\\
&\le 2C\mu d\\
&< d.
\end{array}\end{equation}
Estimates (\ref{6}) and (\ref{7}) imply that
$$
{\rm dist}(y_{k+1}, f(y_k )) < d(2b+1),\quad k \in \mathds{Z},
$$
if $d$ is small enough (let us emphasize here that the required
smallness of $d$ depends on $b$, $N$ and estimates on $A_k$). By
hypothesis, there exist sequences $x_k$ and $t_k$ such that
$$ |t_k-1| \le b'd,\; {\rm dist}(x_k, y_k) \le b'd,\;
x_{k+1} = \phi(t_k ,x_k ),\quad k \in \mathds{Z}.$$ If we write
$$ x_k = \exp_{p_k}(dc_k),\; t_k = 1 + ds_k,$$
then it follows from estimate (\ref{new4}) that
$$|dc_k-d\Delta_k | \le 2\,{\rm dist}(x_k, y_k) \le 2b'd.$$
Thus,
\begin{equation}\label{9}|c_k-\Delta_k| \le 2b',\; k \in \mathds{Z}.
\end{equation}
Clearly,
\begin{equation}\label{10} |s_k| \le b',\; k \in \mathds{Z}.
\end{equation}
We may assume that the value $\rho$ fixed above is small enough, so
that the mappings
$$
G_k:(-\rho,\rho)\times B_T(\rho,p_k)\to {\cal M}_{k+1}
$$
given by
$$
G_k(t,v) = \exp^{-1}_{p_{k+1}}(\phi(1+t,\exp_{p_k}(v))).
$$
are defined. Then $G_k(0,0)=0$,
\begin{equation}\label{11} D_tG_k(t,v)|_{t=0,v=0} = X(p_{k+1}),\quad
D_vG_k(t,v)|_{t=0,v=0} = A_k.\end{equation}
We can write the equality
$$ x_{k+1} = \phi(1 + ds_k, x_k)$$
in the form
$$ \exp_{p_{k+1}}(dc_{k+1}) = \phi(1 + ds_k, \exp_{p_k}(dc_k )),$$
which is equivalent to
\begin{equation}\label{12} dc_{k+1} = G_k(ds_k,dc_k).\end{equation}
Now let $d = d_m$, where $d_m \to 0$. Note that the corresponding
$c_k = c_k^{(m)}$, $t_k = t_k^{(m)}$, and $s_k = s_k^{(m)}$ depend
on $m$.
Since $|c_k^{(m)}| \le 2b' +C$ and $|s_k^{(m)}| \le b'$ for all $m\ge 1$ and $-N \le k \le N-1$, by taking a subsequence if necessary, we can assume that $c_k^{(m)}\to \tilde c_k$, $t_k^{(m)}\to \tilde t_k$,
and $s_k^{(m)}\to \tilde s_k$ for $-N \le k \le N - 1$ as $m\to\infty$.
Applying relations (\ref{12}) and (\ref{11}), we can write
$$ d_mc_{k+1}^{(m)} = G_k(d_ms_k^{(m)},d_mc_k)
= A_kd_mc_k^{(m)}+ X(p_{k+1})d_ms_k^{(m)}+ o(d_m).$$
Dividing by $d_m$, we get the relations
$$ c_{k+1}^{(m)} = A_kc_k^{(m)}+X(p_{k+1})s_k^{(m)} + o(1),\;
-N \le k \le N - 1.$$
Letting $m\to\infty$, we arrive at
$$\tilde c_{k+1} =A_k\tilde c_k + X(p_{k+1})\tilde s_k,\; -N \le k \le N - 1,$$
where
$$ |\Delta_k -\tilde c_k| \le 2b',\quad |\tilde s_k|\le b',\quad-N \le k \le N - 1$$
due to (\ref{9}) and (\ref{10}).
Denote the obtained $\tilde s_k$ by $s_k^{(N)}$. Then $v_k^{(N)} = \Delta_k - \tilde c_k$ is a solution of the system
$$ v^{(N)}_{k+1} = A_kv^{(N)}_k + X(p_{k+1})s_k^{(N)}+ b_{k+1},\;
-N \le k \le N - 1,$$
such that $|v^{(N)}_k| \le 2b'$.
There exist subsequences $s_k^{(j_N)}\to s_k'$ and $v_k^{(j_N)}\to v_k'$ as $N\to\infty$ (we do
not assume uniform convergence) such that $|s_k'| \le b'$, $|v_k'|\le 2b'$, and
$$ v_{k+1}' = A_kv_k' + X(p_{k+1})s_k' + b_{k+1},\; k\in \mathds{Z}.$$
Thus, the lemma is proved.
\end{proof}
Further, we have to refer to two known statements. It is convenient to state them as lemmas.
First we make a definition.
\begin{defin}\label{defPalm4} Consider a sequence of linear isomorphisms
$$ C = \{C_k: \mathds{R}^n \mapsto \mathds{R}^n, k\in \mathds{Z}\}$$
such that $\sup_k(\|C_k\|, \|C_k^{-1}\|) <\infty$. The associated
{\it transition operator} is defined for indices $k$, $l\in \mathds{Z}$ by
$$ \Phi(k,l)=\begin{cases}C_{k-1}\circ\cdots\circ C_l, & l < k,\cr
\Id, & l=k,\cr
C_{k}^{-1}\circ\cdots\circ C^{-1}_{l-1}, & l >
k.
\end{cases}
$$
The sequence $C$ is called {\it hyperbolic on} $\mathds{Z}_+$ (has an {\it
exponential dichotomy} on $\mathds{Z}_+$) if there exist constants $K > 0$,
$\lambda \in (0,1)$, and families of linear subspaces $S_k$, $U_k$
of $\mathds{R}^n$ for $k\in \mathds{Z}_+$ such that
\begin{description}
\item{(1)} $S_k\oplus U_k = \mathds{R}^n$, $k \in \mathds{Z}_+$;
\item{(2)} $C_k(S_k) = S_{k+1}$ and $C_k(U_k) = U_{k+1}$ for $k\in \mathds{Z}_+$;
\item{(3)} $|\Phi(k,l)v|\le K\lambda^{k-l}|v|$ for $v \in S_l$, $k \ge l \ge 0$;
\item{(4)} $|\Phi(k, l)v|\le K\lambda^{l-k}|v|$ for $v \in U_l$, $0\le k\le l$.
\end{description}
\end{defin}
The following result was shown by Maizel$'$ \cite{Maizel} (see also
Coppel \cite{Coppel}).
\medskip
\begin{lem}\label{lemPalm3} If the system
$$ v_{k+1} = C_kv_k + b_{k+1},\quad k \ge 0,$$
has a bounded solution $v_k$ for any bounded sequence $b_k$, then
the sequence $C$ is hyperbolic on $\mathds{Z}_+$ (and a similar statement
holds for $\mathds{Z}_-$).
\end{lem}
The second of the results which we need was proved by Pliss \cite{Pliss}. An analogous statement was proved later by
Palmer \cite{Palmer2, Palmer3}; he also described the Fredholm properties of the corresponding operator
$$ \{v_k\in \mathds{R}^m: k\in \mathds{Z}\}\mapsto \{v_k-A_{k-1}v_{k-1}\}.$$
\medskip
\begin{lem}\label{lemPalm4} Set
$$ B^+(C) = \{v \in \mathds{R}^n: |\Phi(k, 0)v| \to 0,\; k \to +\infty\}$$
and
$$ B^-(C) = \{v \in \mathds{R}^n: |\Phi(k, 0)v| \to 0,\; k \to -\infty\}.$$
Then the following two statements are equivalent:
\begin{description}
\item{{\rm (a)}} for any bounded sequence $\{b_k \in \mathds{R}^n,\; k \in \mathds{Z}\}$ there exists a bounded sequence $\{v_k \in \mathds{R}^n,\; k \in \mathds{Z}\}$ such that
$$ v_{k+1} =C_kv_k + b_{k+1},\quad k \in \mathds{Z};$$
\item{{\rm (b)}} the sequence $C$ is hyperbolic on each of the rays $\mathds{Z}_+$ and $\mathds{Z}_-$, and the subspaces $B^+(C)$ and $B^-(C)$ are transverse.
\end{description}
\end{lem}
\begin{rem}\label{remPalm1} Both Lemmas \ref{lemPalm3} and \ref{lemPalm4} were proved for linear systems of
differential equations, but they hold as well (in the form stated
above) for sequences of linear isomorphisms of Euclidean spaces and
for sequences of linear isomorphisms of arbitrary linear spaces of
the same dimension (we apply them to the isomorphisms $A_k$ of the
spaces ${\cal M}_k$ in Section \ref{secPalm4} and to the
isomorphisms $B_k$ of the spaces $V_k$ in Sections \ref{secPalm5}
and \ref{secPalm6}). For further discussion of this point, see
\cite{Pilyugin3}.
\end{rem}
In the following three sections we assume that $X$ has the Lipschitz
shadowing property (and, consequently, the discrete Lipschitz
shadowing property).
\subsection{Hyperbolicity of the rest points}
\label{secPalm3}
Let $x_0$ be a rest point. We apply Lemma \ref{lemPalm2} with
$p_k=x_0$. Noting that $X(p_k)=0$, we conclude that the difference
equation
$$ v_{k+1}=Df(x_0)v_k+b_{k+1}$$
has a bounded solution $v_k$ for all bounded sequences $b_k\in{\cal
M}_{x_0}$. Then it follows from Lemma \ref{lemPalm4} that
$$ v_{k+1}=Df(x_0)v_k $$
is hyperbolic on both $\mathds{Z}_+$ and $\mathds{Z}_-$. In particular, this
implies that any solution bounded on $\mathds{Z}_+$ tends to $0$ as
$k\to\infty$. However if $Df(x_0)$ had an eigenvalue on the unit
circle, the equation would have a nonzero solution with constant
norm. Hence the eigenvalues of $Df(x_0)$ lie off the unit circle. So
$x_0$ is hyperbolic.
\subsection{The rest points are isolated in the chain recurrent
set}\label{secPalm4}
\begin{lem}\label{lemPalm5} If a rest point $x_0$ is not isolated in the chain
recurrent set ${\cal CR}$, then there is a homoclinic orbit $x(t)$
associated with it.
\end{lem}
\begin{proof} We choose $d>0$ so small that ${\rm dist}(\phi(t,y),x_0)\le
{\cal L}d$ for $|t|$ large implies that $\phi(t,y)\to x_0$ as
$|t|\to\infty$.
Assume that there exists a point $y\in {\cal CR}$ such that $y\neq
x_0$ is arbitrarily close to $x_0$. Since $y$ is chain recurrent,
given any $\varepsilon_0$ and $\theta>0$ we can find points $y_1,\dots,y_N$
and numbers $T_0,\dots,T_N>\theta$ such that
$$
\dist(\phi(T_0,y),y_1)<\varepsilon_0,
$$
$$
\dist(\phi(T_i,y_i),y_{i+1})<\varepsilon_0,\quad i=1,\dots,N,
$$
$$
\dist(\phi(T_N,y_N),y)<\varepsilon_0.
$$
Set $T=T_0+\dots+T_N$ and define $g^*$ on $[0,T]$ by
$$
g^*(t) =
\begin{cases}
\phi(t,y),&\quad 0\leq t< T_0,\\
\phi(t,y_i),&\quad T_0+\dots+T_{i-1}\leq t <T_0+\dots+T_{i},\\
y,&\quad t=T.
\end{cases}
$$
Clearly, for any $\varepsilon>0$ we can find $\varepsilon_0$ depending only on $\varepsilon$
and $\nu$ (see \sref{new5}) such that $g^*(t)$ is an
$\varepsilon$-pseudotrajectory on $[0,T]$.
Then we define
$$
g(t)= \begin{cases}x_0, & t\le 0, \cr
g^*(t), & 0<t\le T, \cr
x_0, & t>T.
\end{cases}
$$
We want to choose $y$ and $\varepsilon$ in such a way that $g(t)$ is
a $d$-pseudotrajectory. We need to show that for all $\tau$ and
$0\le t\le 1$
\begin{equation}\label{ST1}{\rm dist}(\phi(t,g(\tau)),g(t+\tau))\le d.\end{equation}
Clearly this holds for (i) $\tau\le -1$, (ii) $\tau\ge T$, (iii) $\tau$, $\tau+t\in [-1,0]$, and
(iv) $\tau$, $\tau+t\in [0,T]$.
If $-1\le \tau\le 0$, $\tau+t>0$, then with $\nu$ as in \sref{new5}
$$
\begin{array}{rl}
{\rm dist}(\phi(t,g(\tau)),g(\tau+t))
&={\rm dist}(x_0,g^*(\tau+t))\\
&\le {\rm dist}(x_0,\phi(\tau+t,y))+{\rm dist}(\phi(\tau+t,y),g^*(\tau+t))\\
&\le \nu{\rm dist}(x_0,y)+\varepsilon\\
&\le d,\end{array}
$$
if $\dist(y,x_0)$ and $\varepsilon$ are
sufficiently small. Note that, for the fixed $y$, we can decrease
$\varepsilon$ and increase $N,T_0,\dots, T_N$ arbitrarily so that $g(t)$
remains a $d$-pseudotrajectory.
Similarly, (\ref{ST1}) holds if $\tau\in [0,T]$ and $\tau+t>T$.
Thus $g(t)$ is ${\cal L}d-$shadowed by a trajectory $x(t)$ so that in particular\newline
dist$(x(t),x_0) \le{\cal L}d$ if $|t|$ is sufficiently large so that $x(t)\to x_0$ as $|t|\to\infty$.
\medskip
We must also be sure that $x(t)\neq x_0$. If $y$ is not on the local stable manifold of $x_0$,
then there exists $\varepsilon_1>0$ independent of $y$ such that
${\rm dist}(\phi(t_0,y),x_0)\ge\varepsilon_1$ for some $t_0>0$. We
can choose $T_0>t_0$. Now we know that ${\rm
dist}(x(t),\phi(t_0,y))\le {\cal L}d$. So provided ${\cal
L}d<\varepsilon_1$, we have $x(t_0)\neq x_0$.
If $y$ is on the local stable manifold of $x_0$, then provided ${\rm dist}(y,x_0)$ is sufficiently
small, it is not on the local unstable manifold of $x_0$.
Then, applying the same argument to the flow with time reversed noting that the chain recurrent set is also the chain recurrent set for the reversed flow and also that the reversed flow will have the Lipschitz shadowing property also, we show that $x(t)\neq x_0$.
\medskip
Now we show the existence of this homoclinic orbit $x(t)$ leads to a
contradiction. Set $p_k=x(k)$. Since $A_kX(p_k)=X(p_{k+1})$, it is
easily verified that if
$$ \beta_{k+1}=\beta_k+s_k,\quad k\in \mathds{Z}$$
then $v_k=\beta_kX(p_k)$ is a solution of
\begin{equation}\label{27} v_{k+1}=A_kv_k +X(p_{k+1})s_k,\quad k\in \mathds{Z}. \end{equation}
Also if $s_k$ is bounded then $\beta_kX(p_k)$ is also bounded, since
$X(p_k)\to 0$ exponentially as $|k|\to\infty$ and $|\beta_k|/|k|$ is
bounded.
By Lemma \ref{lemPalm2}, for all bounded $b_k\in{\cal M}_k$ there
exists a bounded scalar sequence $s_k$ such that
$$ v_{k+1}=A_kv_k+X(p_{k+1})s_k+b_{k+1}$$
has a bounded solution. But we know (\ref{27}) has a bounded solution. It follows that
$$ v_{k+1}=A_kv_k+b_{k+1}$$
has a bounded solution for arbitrary $b_k\in{\cal M}_k$. Then it
follows from Lemma \ref{lemPalm4} that
$$ v_{k+1}=A_kv_k$$
is hyperbolic on both $\mathds{Z}_+$ and $\mathds{Z}_-$ and that the spaces
$B^+(A)$ and $B^-(A)$ are transverse. This is a contradiction since
dim\,$B^+(A)+$ dim\,$B^-(A)=n$ (because $B^+(A)$ has the same
dimension as the stable manifold of $x_0$ and $B^-(A)$ has the same
dimension as the unstable manifold of $x_0$) but they contain
$X(p_0)\neq 0$ in their intersection.
So we conclude that the rest points are isolated in the chain recurrent set.
\end{proof}
\subsection{Hyperbolicity of the chain recurrent
set}\label{secPalm5}
We have shown that the rest points of $X$ are hyperbolic and form a finite,
isolated subset of the chain recurrent set ${\cal C}{\cal R}$. Let $\Sigma$ be the chain recurrent
set minus the rest points. We want to show this set is hyperbolic. To this
end we use the following lemma. Let us first introduce some notation.
Let $x(t)$ be a trajectory of $X$ in $\Sigma$. Put $p_k = x(k)$ and denote by $P_k$
the orthogonal projection in ${\cal M}_k$ with kernel spanned by $X(p_k)$ and by $V_k$
the orthogonal complement to $X(p_k)$ in ${\cal M}_k$. Introduce the operators $B_k=
P_{k+1}A_k:V_k\mapsto V_{k+1}$.
\medskip
\begin{lem}\label{lemPalm6} For every bounded sequence $b_k\in V_k$
(denote $b=\|b\|_{\infty}$) there exists a solution $v_k\in V_k$ of
the system
\begin{equation}\label{14} v_{k+1}=B_kv_k+b_{k+1},\quad k\in \mathds{Z}, \end{equation}
such that for all $k$,
$$ |v_k| \le 2L(2b + 1).$$
\end{lem}
\begin{proof} By Lemma \ref{lemPalm2}, there exists a bounded sequence $s_k$ such that
the system
\begin{equation}\label{15} w_{k+1}=A_kw_k+X(p_{k+1})s_k+b_{k+1},\quad k\in \mathds{Z} \end{equation}
has a solution $w_k$ with $|w_k|\le 2L(2b + 1)$.
Note that $A_kX(p_k)=X(p_{k+1})$. Since $(\Id-P_k)v\in \{X(p_k)\}$
for $v\in {\cal M}_k$, we see that $P_{k+1}A_k(\Id-P_k)=0$, which
gives us the equality
\begin{equation}\label{16} P_{k+1}A_k=P_{k+1}A_kP_k. \end{equation}
Multiplying (\ref{15}) by $P_{k+1}$, taking into account the equalities $P_{k+1}X(p_{k+1})=0$
and $P_{k+1}b_{k+1}=b_{k+1}$, and applying (\ref{16}), we see that $v_k=P_kw_k$ is the required
solution.
Thus, the lemma is proved.
\end{proof}
Now we prove $\Sigma$ is hyperbolic. Let $x(t)$ be a trajectory in
$\Sigma$ with the same notation as given before Lemma
\ref{lemPalm6}. Then by Lemmas \ref{lemPalm6} and \ref{lemPalm4},
\begin{equation}\label{17} v_{k+1}=B_kv_k,\quad v_k\in V_k \end{equation}
is hyperbolic on both $\mathds{Z}_+$ and $\mathds{Z}_-$ and $B^+(B)$ and $B^-(B)$
are transverse. It follows that the adjoint system
$$ v_{k+1}=(B_k)^{*-1}v_k,\quad v_k\in V_k $$
is hyperbolic on both $\mathds{Z}_+$ and $\mathds{Z}_-$ and has no nontrivial
bounded solution.
Now we consider the discrete linear skew product flow on the normal
bundle $V$ over $\Sigma$ generated by the map defined for
$p\in\Sigma$, $v\in V_p$ (where $V_p$ is the orthogonal complement
to $X(p)$ in $T_pM$) by
\begin{equation}\label{Text11.1}
(p,v) \mapsto (\phi(1,p),B_pv),
\end{equation}
where $B_p=P_{\phi(1,p)}D\phi(1,p)$, $P_p$ being the orthogonal
projection of $T_pM$ onto $V_p$. Its adjoint flow is generated by
the map defined by
$$ (p,v) \mapsto (\phi(1,p),(B^*_p)^{-1}v).$$
Now we want to apply the Corollary on page 492 in Sacker and Sell \cite{SS1}.
What we have shown above is that the adjoint flow has the
no nontrivial bounded solution property. It follows from the Sacker and Sell corollary that
the adjoint flow is hyperbolic and hence the original
skew product flow
$$ (p,v) \mapsto (\phi(1,p),B_pv)$$
is also. However then it follows from Theorem 3 in Sacker and Sell \cite{SS2} that $\Sigma$ is hyperbolic.
\subsection{Strong Transversality}\label{secPalm6}
To verify strong transversality, let $x(t)$ be a trajectory that
belongs to the intersection of the stable and unstable manifolds of
two trajectories, $x_+(t)$ and $x_-(t)$, respectively, lying in the
chain recurrent set. Denote $p_0=x(0)$ and $p_k=x(k),\;k\in\mathds{Z}$; let
$W^s(p_0)$ and $W^u(p_0)$ denote the stable manifold of $x_+(t)$ and
the unstable manifold of $x_-(t)$, respectively. Denote by $E^s$ and
$E^u$ the tangent spaces of $W^s(p_0)$ and $W^u(p_0)$ at $p_0$.
By Lemma \ref{lemPalm6} (using the same notation as in the previous
section), for all bounded $b_k\in V_k$, there exists a bounded
solution $v_k\in V_k$ of (\ref{14}). By Lemma \ref{lemPalm4} again,
this implies that
\begin{equation}\label{STC1}
{\cal E}^s+{\cal E}^u=V_0,
\end{equation}
where
$$
\begin{array}{rl}
{\cal E}^s&=\{w_0: w_{k+1}=B_kw_k,\; \sup_{k\ge 0}|w_k|<\infty\},\\ \\
{\cal E}^u&=\{w_0: w_{k+1}=B_kw_k,\; \sup_{k\le
0}|w_k|<\infty\}\end{array}
$$
Moreover (\ref{17}) is hyperbolic on both $\mathds{Z}_+$ and $\mathds{Z}_-$.
We are going to use the following folklore result, which for
completeness we prove after showing it implies the strong
transversality:
\begin{equation}\label{STC2}
{\cal E}^s\subset E^s,\quad {\cal E}^u\subset E^u.
\end{equation}
Combining equality\ (\ref{STC1}) with the inclusions (\ref{STC2})
and the trivial relations
$$
E^s = V_0 \cap E^s + \{X(p_0)\},\quad E^u = V_0 \cap E^u +
\{X(p_0)\},
$$
we conclude that
$$
E^s + E^u = T_{p_0}M,
$$
and so the strong transversality holds.
Let us now prove the first relation in (\ref{STC2}); the second one
can be proved in a similar way.
{\it Case 1}: The limit trajectory in ${\cal C}{\cal R}$ is a
rest point. In this case, the stable manifold of the rest point
coincides with its stable
manifold as a fixed point of the time-one map $f(x)=\phi(1,x)$. By
the theory for diffeomorphisms, if $p_k$ is a trajectory on the
stable manifold, the tangent space to the stable manifold at $p_0$
is the subspace $E^s$ of initial values of bounded solutions of
\begin{equation}
\label{STC3}
v_{k+1} = A_k v_k, \quad k \geq 0.
\end{equation}
Let us prove that ${\cal E}^s \subset E^s$. Fix an arbitrary sequence
$w_k$ satisfying
$w_{k+1} = B_k w_k$ with $w_0 \in {\cal E}^s$.
Consider the sequence
$$
v_k = \lambda_kX(p_k)/|X(p_k)|+w_k
$$
with
$\lambda_k$ satisfying
\begin{equation}\label{21}
\lambda_{k+1}=\displaystyle\frac{|X(p_{k+1})|}{|X(p_k)|}\lambda_k-
\frac{X(p_{k+1})^*}{|X(p_{k+1})|}A_kw_k
\end{equation}
and $\lambda_0 = 0$. It is easy to see that $v_k$ satisfy
(\ref{STC3}).
Since $x(t)$ is on the stable manifold of a hyperbolic rest point,
there are positive constants $K$ and $\alpha$ such that
$$
|\dot x(t)|\le Ke^{-\alpha(t-s)}|\dot x(s)|
$$
for $0\le s\le t$. From this it follows that
$$
|X(p_k)|\le Ke^{-\alpha(k-m)}|X(p_m)|
$$
for $0\le m\le k$ so that the scalar difference equation
$$
\lambda_{k+1}=\frac{|X(p_{k+1})|}{|X(p_k)|}\lambda_k
$$
is hyperbolic on $\mathds{Z}_+$ and is, in fact, stable. Since the second
term on the right-hand side of equation (\ref{21}) is bounded as
$k\to\infty$, it follows that $\lambda_k$ are bounded for any choice
of $\lambda_0$. This fact implies that $v_k$ is a bounded solution
of (\ref{STC3}), and we conclude that $v_0 = w_0 \in E^s$, hence
${\cal E}^s \subset E^s$.
The proof in Case 1 is complete.
\medskip
{\it Case 2}: Assume that the limit trajectory is in $\Sigma$, the
chain recurrent set minus the fixed points which we know to be
hyperbolic. We want to find the intersection of its stable manifold
near $p_0=x(0)$ with the cross-section at $p_0$ orthogonal to the
vector field (in local coordinates generated by the exponential
mapping). To do this, we discretize the problem and note that there
exists a number $\sigma>0$ such that a point $p\in M$ close to $p_0$
certainly belongs to $W^s(p_0)$ if and only if the distances of
consecutive points of intersections of the positive semitrajectory
of $p$ with the sets $\exp_{p_k}({\cal M}_k)$ to the points $p_k$ do
not exceed $\sigma$.
\medskip
For suitably small $\mu>0$, we find all sequences $t_k$ and $z_k\in
V_k$, the subspace of $T_{p_k}M$ orthogonal to $X(p_k)$, such that
for $k\ge 0$
$$
|t_k-1|\le \mu,\quad |z_k|\le\mu,\quad y_{k+1}=\phi(t_k,y_k),
$$
where $y_k=\exp_{p_k}(z_k)$. Thus we have to solve the equation
$$
\exp_{p_{k+1}}(z_{k+1})=\phi(t_k,\exp_{p_k}(z_k)),\quad k\ge 0
$$
for $t_k$ and $z_k\in V_k$ such that $|t_k-1|\le \mu$ and
$|z_k|\le\mu$.
\medskip
We set it up as a problem in Banach spaces. By lemmas \ref{lemPalm4}
and \ref{lemPalm6} the difference equation
$$
z_{k+1}=B_kz_k,\quad
z_k\in V_k
$$
(recall that $B_k=P_{k+1}A_k$ and $P_k$ is the orthogonal projection
on ${\cal M}_k$ with range $V_k$), has an exponential dichotomy on
$\mathds{Z}_+$ with projection (say) $Q_k:V_k\mapsto V_k$. Denote by ${\cal
R}(Q_0)$ the range of $Q_0$ and note that ${\cal R}(Q_0)={\cal
E}^s$. Fix a positive number $\mu_0$ and denote by $V$ the space of
sequences
$$
\{z_k\in V_k, |z_k|\leq \mu_0,k\in\mathds{Z}_+\}
$$
and by $l^{\infty}(\mathds{Z}_+, \{{\cal M}_{k+1}\})$ the space of
sequences $\{\zeta_k \in {\cal M}_{k+1}, \; k \in \mathds{Z}_+ \}$ with the
usual norm.
Then the $C^1$ function
$$
G:[1-\mu_0,1+\mu_0]^{\mathds{Z}_+}\times V\times {\cal R}(Q_0)\mapsto
\ell^{\infty}(\mathds{Z}_+,\{{\cal M}_{k+1}\})\times {\cal R}(Q_0)
$$
given by
$$ G(t,z,\eta)=(\{ z_{k+1}-\exp^{-1}_{p_{k+1}}(\phi(t_k,\exp_{p_k}(z_k))\}_{k\ge 0},
Q_0z_0-\eta)
$$
is defined if $\mu_0$ is small enough.
We want to solve the equation
$$
G(t,z,\eta)=0
$$
for $(t,z)$ as a function of $\eta$. It is clear that
$$
G(1,0,0)=0,
$$
where the first argument of $G$ is $\{1,1,\ldots\}$, the second
argument is $\{0,0,\ldots\}$ and the right-hand side is
$(\{0,0,\ldots\},0)$.
To apply the implicit function theorem, we must verify that
$$
T=\frac{\partial G}{\partial (t,z)}(1,0,0)
$$
is invertible. Note that if $(s,w)\in \ell^{\infty}(\mathds{Z}_+,\mathds{R})\times
V$, then
$$
T(s,w)=(\{ w_{k+1}-X(p_{k+1})s_k-A_kw_k\}_{k\ge 0}, Q_0w_0).
$$
To show that $T$ is invertible, we must show that
$$
T(s,w)=(g,\eta)
$$
has a unique solution $(s,w)$ for all $(g,\eta)\in
l^\infty(\mathds{Z}_+,\{{\cal M}_{k+1}\})\times {\cal R}(Q_0)$. So we need
to solve the equations
$$
w_{k+1}=A_kw_k+X(p_{k+1})s_k+g_k,\quad k\ge 0
$$
subject to
$$
Q_0w_0=\eta.
$$
If we multiply the difference equation by $X(p_{k+1})^*$ and solve
for $s_k$, we obtain
$$
s_k=-\frac{X(p_{k+1})^*}{|X(p_{k+1})|^2}[A_kw_k+g_k],\quad k\ge 0
$$
and if we multiply it by $P_{k+1}$, we obtain
$$
w_{k+1}=P_{k+1}A_kw_k+P_{k+1}g_k=B_kw_k+P_{k+1}g_k,\quad k\ge 0.
$$
Now we know this last equation has a unique bounded solution $w_k\in
V_k$, $k \geq 0$, satisfying $Q_0w_0=\eta$. Then the invertibility
of $T$ follows.
\medskip
Thus we can apply the implicit function theorem to show that there
exists $\mu>0$ such that provided $|\eta|$ is sufficiently small,
the equation $G(t,z,\eta)=0$ has a unique solution
$(t(\eta),z(\eta))$ such that $\|t-1\|_{\infty}\le\mu$,
$\|z\|_{\infty}\le\mu$. Moreover, $t(0)=1$, $z(0)=0$ and the
functions $t(\eta)$ and $z(\eta)$ are $C^1$.
The points $\exp_{p_0}(z_0(\eta))$ with small $|\eta|$ form a
submanifold containing $p_0$ and contained in $W^s(p_0)$. Thus, the
range of the derivative $z_0'(0)$ is contained in $E^s$.
Take an arbitrary vector $\xi\in {\cal E}^s$ and consider
$\eta=\tau\xi,\xi\in\mathds{R}$. Differentiating the equalities
$$
z_{k+1}(\tau\xi)=\exp_{p_{k+1}}^{-1}(\phi(t_k(\tau\xi),
\exp_{p_{k}}(z_k(\tau\xi)))),\quad k\geq 0,
$$
and
$$
Q_0z_0(\tau\xi)=\tau\xi
$$
with respect to $\tau$ at $\tau=0$, we see that
$$
s_k=\frac{\partial t_k}{\partial \eta}|_{\eta=0}\xi,\quad
w_k=\frac{\partial z_k}{\partial \eta}|_{\eta=0}\xi\in V_k
$$
are bounded sequences satisfying
$$
w_{k+1}=A_kw_k+X(p_{k+1})s_k, \quad Q_0w_0=\xi.
$$
Multiplying by $P_{k+1}$, we conclude that
$$
w_{k+1}=B_kw_k, k\geq 0,\quad Q_0w_0=\xi.
$$
It follows that $w_0\in{\cal E}^s={\cal R}(Q_0)$. Then
$w_0=Q_0w_0=\xi$. We have shown that the range of $z_0'(0)$ is
exactly ${\cal E}^s$, and thus ${\cal E}^s\subset E^s$.
\section{Lipschitz periodic shadowing} \label{secLipPerSh}
It is known that a vector field $X$ is $\Omega$-stable if and only
$X$ satisfies Axiom A$'$ and the no-cycle condition (see
\cite{PughShub} and \cite{Hayashi}). Thus, to prove Theorem
\ref{thmLipPerSh}, we prove the following two lemmas.
\begin{lem}\label{lemLipPer1}
If a vector field $X$ has the Lipschitz periodic shadowing property,
then $X$ satisfies Axiom A$'$ and the no-cycle condition.
\end{lem}
\begin{lem}\label{lemLipPer2}
If $X$ satisfies Axiom A$'$ and the no-cycle condition, then $X$ has
the Lipschitz periodic shadowing property.
\end{lem}
Lemma \ref{lemLipPer1} is proved in Secs.
\ref{secLipPer1}-\ref{secLipPer5}; Lemma \ref{lemLipPer2} is proved
in Sec. \ref{secLipPer6}.
The proof of Lemma \ref{lemLipPer1} is divided into several steps.
We assume that $X$ has the Lipschitz periodic shadowing property and
establish the following statements.
\begin{enumerate}
\item Closed trajectories are uniformly hyperbolic.
\item Rest points are hyperbolic.
\item The chain-recurrent set coincides with the closure of the set of rest
points and closed trajectories; rest points are separated from the
remaining part of the chain-recurrent set.
\item The hyperbolic structure on the set of closed trajectories can be extended
to the chain-recurrent set.
\item The no-cycle condition holds.
\end{enumerate}
\subsection{Uniform hyperbolicity of closed trajectories}
\label{secLipPer1}
Without loss of generality we can assume that ${\cal L} > 1$.
Let $x(t)$ be a nontrivial closed trajectory of period $\omega$.
Choose $n_1,n \in \mathds{N}$ such that $\tau = n_1\omega/n \in [1/2, 1]$.
Let $x_k = x(k\tau)$, $f(x) = \phi(\tau, x)$, and $A_k = D f(x_k)$.
Note that $A_{k+n} = A_k$. Below we prove a statement similar to
Lemma \ref{lemPalm2}.
\begin{lem}
\label{LP2} If $X \in \LipPerSh$, then for any $b>0$ there exists a
constant $K$ (the same for all closed trajectories $x(t)$ of $X$)
such that for any sequence $b_k \in T_{x_k} M$ with $|b_k| < b$
there exist sequences $s_k \in \mathds{R}$ and $v_k \in T_{x_k} M $ with
the following properties:
\begin{equation}
\label{2.1} v_{k+1} = A_k v_k + X(x_{k+1})s_{k+1} + b_{k+1}
\end{equation}
and
\begin{equation}
\label{2.2} |s_k|, |v_k| \leq K.
\end{equation}
\end{lem}
Before we go to the proof of Lemma \ref{LP2}, we need to generalize
the notion of discrete Lipschitz shadowing property. Let $d,\tau >
0$; we say that a sequence $y_k$ is a $\tau$-discrete
$d$-pseudotrajectory if $\dist(y_{k+1}, \phi(\tau, y_k)) < d$.
Let $\varepsilon>0$; we say that a sequence $x_k$ $\varepsilon$-shadows $y_k$ if
there exists a sequence $t_k > 0$ such that
$$
\dist(x_k, y_k)< \varepsilon, \quad |t_k - \tau| < \varepsilon, \quad x_{k+1}
=\phi(t_k, x_k).
$$
The following lemma can be proved similarly to Lemma \ref{lemPalm1}.
\begin{lem} If $X \in \LipPerSh$, then there exist constants $d_0, L > 0$ such
that for any $\tau \in [1/2, 1]$ and $d>0$ and any periodic
$\tau$-discrete $d$-pseudotrajectory $y_k$ with $d\leq d_0$ there
exists a sequence $x_k$ (not necessarily periodic) that $Ld$-shadows
$y_k$.
\end{lem}
In the proof of Lemma \ref{LP2}, we use the following technical
statement which is well-known in control theory
\cite{Sontag},\cite{Heemels}.
\begin{lem}
\label{LP2.5} Let $B:\mathds{R}^m \to \mathds{R}^m$ be a linear operator such that
the absolute values of its eigenvalues equal 1. Then for any
$\Delta_0 \in \mathds{R}^m$ and $\delta > 0$ there exists a number $R \in
\mathds{N}$ and a sequence $\delta_k \in \mathds{R}^m$, \;$k \in [1, R]$, such
that $|\delta_k| < \delta $ and the sequence $\Delta_k \in \mathds{R}^m$
defined by
\begin{equation}
\label{AC1.0.4} \Delta_{k+1} = B\Delta_k + \delta_{k+1}, \quad k \in
[0, R-1],
\end{equation}
satisfies $\Delta_R = 0$.
\end{lem}
\begin{proof}[Proof of Lemma \ref{LP2}]
Fix an arbitrary sequence $b_k$ with $|b_k| < b$ and a number $l \in
\mathds{N}$.
First we will find a number $l_1 > l$ and sequences $c_k$ and
$\Delta_k$ defined for $k \in [-ln, l_1n]$ such that $|c_k| < b$ and
$$
c_k = b_k, \quad k \in [-ln, ln],$$
\begin{equation}
\label{3.0.5} \Delta_{k+1} = A_k \Delta_k + c_{k+1}, \quad k \in
[-ln, l_1 n - 1],
\end{equation}
$$
\Delta_{-ln} = \Delta_{l_1n}.
$$
Consider the operator $A: T_{x_0} \to T_{x_0}$ defined by $A =
A_{n-1}\cdots A_0$.
The tangent space $T_{x_0}$ can be represented in the form
\begin{equation}
\label{3.1} T_{x_0} = E^s_0 \oplus E^c_0 \oplus E^u_0
\end{equation}
so that the subspace $E^s_0$ corresponds to the eigenvalues
$\lambda_j$ of $A$ such that $|\lambda_j|<1$, the subspace $E^c_0$
corresponds to the eigenvalues $\lambda_j$ such that
$|\lambda_j|=1$, and the subspace $E^u_0$ corresponds to the
eigenvalues $\lambda_j$ such that $|\lambda_j|>1$.
For any index $k$ consider the decomposition $T_{x_k} = E^s_k \oplus
E^c_k \oplus E^u_k$ as the image of decomposition \sref{3.1} under
the mapping $A_{k-1}\cdots A_0$.
In the coordinates corresponding to these decompositions, the
matrices $A_k$ can be represented in the following form:
$$
A_k = \diag(A^s_k, A^c_k, A^u_k).
$$
Set $A_\sigma=A^\sigma_{n-1}\cdots A^\sigma_{0}$ for $\sigma=s,c,u$.
Consider the corresponding coordinate representations $b_k = (b_k^s,
b_k^c, b_k^u)$, $c_k = (c_k^s, c_k^c, c_k^u)$, and $\Delta_k =
(\Delta_k^s, \Delta_k^c, \Delta_k^u)$ (and note that the values
$|b_k^s|, |b_k^c|, |b_k^u|$ are not necessarily less than $b$).
Equations \sref{3.0.5} are equivalent to the system
\begin{equation}
\label{4.2} \Delta^s_{k+1} = A_k^s \Delta^s_{k} + c_{k+1}^s,
\end{equation}
\begin{equation}
\notag\Delta^c_{k+1} = A_k^c \Delta^c_{k} + c_{k+1}^c,
\end{equation}
\begin{equation}
\label{4.3} \Delta^u_{k+1} = A_k^u \Delta^u_{k} + c_{k+1}^u.
\end{equation}
Set $c_k = b_k$ for $k \in [-ln, ln - 1]$.
Consider the sequence satisfying \sref{4.2} with initial data
$\Delta^s_{-ln} = 0$ and denote $\Delta^s_{ln}$ by $a^s$; Consider
the sequence satisfying \sref{4.3} with initial data $\Delta^u_{ln}
= 0$ and denote $\Delta^u_{-ln}$ by $a^u$.
There exist numbers $l_s, l_u > 1$ such that
$$
|A^{-l} a^u| < b, \; l \geq l_u, \quad |A^{l} a^s| < b, \; l \geq
l_s.
$$
Set $\Delta_{-ln} = (0, 0, a^u)$; then the definition of $a^s$ and
$a^u$ implies that $\Delta_{ln} = (a^s, C_1, 0)$ for some $C_1 \in
E^c_0$.
Set $c_k = 0$ for $k \in [ln+1, (l+l_s)n]$; then $\Delta_{(l+l_s)n}
= (A_s^{l_s}a^s, C_2, 0)$ for some $C_2 \in E^c_0$.
Set $c_k = 0$ for $k \in [(l+l_s)n+1, (l+l_s+1)n-1]$ and $c_k
=(-A_s^{l_s+1}a^s, 0, 0)$ for $k = (l+l_s+1)n$. Then
$\Delta_{(l+l_s+1)n} = (0, C_3, 0)$ for some $C_3 \in E^c_0$.
Applying Lemma~\ref{LP2.5} to $A^c:E^c_0\to E^c_0$, we find a number
$R$ and a sequence $\delta_k$ with $|\delta_k|\leq b$ such that if
$$
x_{i+1}=A^c x_i+\delta_{i+1},\quad x_0=\Delta^c_{(l+l_s+1)n},
$$
then $x_R=0$. Then if we set for $i=0,\dots,R-1$, $c_k=0$ for
$(l+l_s+i+1)n+1\leq k\leq (l+l_s+i+2)n-1$ and
$c_{(l+l_s+i+2)n}=(0,\delta_{i+1},0)$, we see that
$$
\Delta^c_{(l+l_s+i+2)n}=A^c\Delta^c_{(l+l_s+i+1)n}+\delta_{i+1},
\quad i=0,\dots,R-1,
$$
so that $\Delta^c_{(l+l_s+R+1)n}=0$; of course, the other two
components of $\Delta_{(l+l_s+R+1)n}$ remain zero.
Set $c_k = 0$ for $k \in [(l+l_s+R+1)n +1, (l+l_s+R+2)n -1]$ and
$c_k = (0, 0, A_u^{-l_u}a^u)$ for $k = (l+l_s+R+2)n$; then
$\Delta_{(l+l_s+R+2)n} = (0, 0, A_u^{-l_u}a^u)$. Finally, we set
$c_k =0$ for $k \in [(l+l_s+R+2)n +1, (l+l_s+R+2+l_u)n]$ and see
that $\Delta_{(l+l_s+2+R+l_u)n} = (0, 0, a^u) = \Delta_{-ln}$. Thus,
we have constructed the sequences mentioned in the beginning of the
proof.
Taking $d$ small enough, considering the periodic $\tau$-discrete
pseudotrajectory $y_k = \exp_{x_k}(d\Delta_k)$, and repeating the
reasoning similar to that in the proof of Lemma \ref{lemPalm2}, we
can prove that relations \sref{2.1} and \sref{2.2} hold with
$K=L(2b+1)$ for $k \in [-ln, ln-1]$.
After that, we repeat the reasoning used in the last two paragraphs
of the proof of Lemma \ref{lemPalm2} to complete the proof of Lemma
\ref{LP2}.
\end{proof}
As in Sec. \ref{secPalm5}, we define ${\cal M}_k, V_k, P_k$, and
$B_k = P_{k+1}A_k: V_k \to V_{k+1}$. Note that $B_k^{-1} =
P_kA_{k}^{-1}$. Since $M$ is compact, there exists a constant $N>0$
such that $\| D \phi(\tau, x) \| < N$ for any $\tau \in [-1, 1]$
and $x \in M$. Hence, $\|A_k\|, \|A_k^{-1}\| < N$, and
\begin{equation}
\label{8.1} \|B_k\|, \|B_k^{-1}\| < N.
\end{equation}
The same reasoning as in the proof of Lemma~6 establishes the
following statement.
\begin{lem}
\label{LP8} There exists a constant $K > 0$ (the same for all closed
trajectories $x(t)$) such that for every sequence $b_k \in V_k$ with
$|b_k| \leq 1$ there exists a solution $v_k \in V_k$ of the system
$$
v_{k+1} = B_k v_k + b_{k+1}
$$
such that
$$
\|v_k\| \leq K.$$
\end{lem}
A remark on page 26 of \cite{Coppel}, Lemma \ref{LP8} and the
inequalities \sref{8.1} imply that there exist constants $C_1 > 0$
and $\lambda_1 \in (0, 1)$ (the same for all closed trajectories) and a
representation $V_k = E^s(x_k) \oplus E^u(x_k)$ such that
$$
B_k E^s(x_k) = E^s(x_{k+1}), \quad B_k E^u(x_k) = E^u(x_{k+1}),
$$
$$
|B_{l+k}\cdots B_k v^s| \leq C_1 \lambda_1^l|v_s|, \quad v^s\in
E^s(x_k), \; l>0, k \in \mathds{Z}
$$
$$
|B^{-1}_{-l+k}\cdots B_k^{-1} v^u| \leq C_1\lambda_1^l|v_u|, \quad v^u
\in E^u(x_k), \; l>0, k \in \mathds{Z}.
$$
\begin{rem}
In fact in \cite{Coppel} exponential dichotomy with uniform
constants was proved only on $\mathds{Z}^+$. However we can extend the
corresponding inequalities to the whole of $\mathds{Z}$ by the periodicity
of $B_k$.
\end{rem}
Since $\tau \in [1/2, 1]$ and $\| D \phi(\tau, x)\| \leq N$ the
above conditions imply that there exist constants $C_2
> 0$ and $\lambda_2\in (0, 1)$ such that if $x(t)$ is a closed
trajectory, then
\begin{equation}
\label{8.5.1} |P_{\phi(t, x_0)} D \phi(t, x(t_0)) v^s| \leq C_2
\lambda_2^t |v^s|, \quad v^s \in E^s(x(t_0)), \; t>0, \; t_0 \in \mathds{R},
\end{equation}
\begin{equation}
\label{8.5.2} |P_{\phi(-t, x_0)} D \phi(-t, x(t_0)) v^u| \leq C_2
\lambda_2^t |v^u|, \quad v^u \in E^u(x(t_0)), \; t>0, \; t_0 \in \mathds{R},
\end{equation}
where $P_{y\in M}$ is the orthogonal projection of $T_yM$ with
kernel $X(y)$, $E^{s, u}(x(t_0)) = P_{\phi(t_0, x)} D \phi(t_0, x)
E^{s, u}(x_0)$.
\begin{rem}\label{LPR} In particular, the above
inequalities imply that $x(t)$ is a hyperbolic closed trajectory.
\end{rem}
\subsection{Hyperbolicity of the rest points}
Let $x_0$ be a rest point. As in subsection \ref{secPalm3} (using
Lemma \ref{LP2}),
we conclude that $D \phi(1, x_0)$ is hyperbolic; hence, $x_0$ is a hyperbolic rest point.
\subsection{The rest points are separated from the remaining part of the
chain-recurrent set}
Denote by $\Per(X)$ the set of rest points and points belonging to
closed trajectories of a vector field $X$; let ${\cal C}{\cal R}(X)$ be the set
of its chain-recurrent points. For a set $A \subset M$ denote by
$\Cl A$ the closure of $A$ and by $B(a,A)$ its $a$-neighborhood.
\begin{lem}
\label{LP7} If $X \in \LipPerSh$, then $\Cl \Per(X) = {\cal C}{\cal R}(X)$.
\end{lem}
\begin{proof} If $y_0 \in {\cal C}{\cal R}(X)$, then for any $d>0$
there exists a periodic $d$-pseudotrajectory $g(t)$ such that $g(0)
=y_0$.
Since $X\in\LipPerSh$, there exists a point $x_d \in \Per(X)$ such
that $\dist(x_d, y_0) < \mbox{$\cal{L}$} d$. Hence, $B(\mbox{$\cal{L}$} d, y_0) \cap \Per(X)
\ne \emptyset$ for arbitrary $d > 0$, which proves our lemma.
\end{proof}
\begin{lem}Let $X \in \LipPerSh$ and let $p$ be a rest point of $X$.
Then $p \notin \Cl({\cal C}{\cal R}(X)\setminus p)$.
\end{lem}
\begin{proof}
It has already been proved that all rest points of a vector field $X
\in \LipPerSh$ are hyperbolic; hence the set of rest points is
finite. Assume that $p \in \Cl({\cal C}{\cal R}(X)\setminus p)$. Then Lemma
\ref{LP7} implies that $p \in \Cl(\Per(X)\setminus p)$.
Denote by $W^s_{loc,a}(p)$ and $W^u_{loc,a}(p)$ the local stable and
unstable manifolds of size $a$.
Since the rest point $p$ is hyperbolic, there exists $\varepsilon \in (0,
1/2)$ such that if $x \in M$ and $\phi(t,x)\subset B(4\varepsilon,
p),\;t\geq 0$, then $x \in W^s_{loc,4\varepsilon}(p)$; if $\phi(t,x)\subset
B(4\varepsilon, p),\;t\leq 0$, then $x \in W^u_{loc,4\varepsilon}(p)$; and if
$\phi(t,x)\subset B(4\varepsilon, p),\;t\in\mathds{R}$, then $x=p$.
Let $d_1 = \min(d_0, \varepsilon/\mbox{$\cal{L}$})$, where $d_0$ and $\mbox{$\cal{L}$}$ are the
constants from the definition of $\LipPerSh$. Take a point $x_0 \in
\Per(X)$ (let the period of the trajectory of $x_0$ equal $\omega$)
and a number $T>0$ and define the mapping
$$
g_{x_0, T}(t) =
\begin{cases} p, & \quad t \in [-T, T],\\
\phi(t - T, x_0), & \quad t \in (T, T+\omega),\\
\end{cases}
$$
for $t \in [-T, T+\omega)$. Continue this mapping periodically to
the line $\mathds{R}$.
There exists $d_2 < d_1$ depending only on $d_1$ and $\nu$ (see
\sref{new5}) such that if $x_0 \in B(d_2, p)$, then $g_{x_0, T}(t)$
is a $d_1$-pseudotrajectory for any $T > 0$. We fix such a point
$x_0\in B(d_2,p)$ and consider below pseudotrajectories $g_{x_0,T}$
with this fixed $x_0$ and with increasing numbers $T$.
By our assumptions, the pseudotrajectory $g_{x_0, T}$ can be
$\varepsilon$-shadowed by the trajectory of a point $z_T \in \Per(X)$ with
reparametrization $\alpha_T(t)$:
\begin{equation}
\label{11.1} \dist(g_{x_0, T}(t), \phi(\alpha_T(t), z_T)) < \varepsilon.
\end{equation}
Our choice of $\varepsilon$ implies that there exist times $t_1, t_2 > 0$
such that
$$
\dist(p, \phi(t_1, x_0)) \in [2\varepsilon, 3\varepsilon], \quad \phi(t, x_0) \in
B(4\varepsilon, p), \quad t \in [0, t_1],
$$
$$
\dist(p, \phi(-t_2, x_0)) \in [2\varepsilon, 3\varepsilon], \quad \phi(t, x_0) \in
B(4\varepsilon, p), \quad t \in [-t_2, 0].
$$
We emphasize that the numbers $t_1,t_2$ depend on our choice of the
point $x_0$ but not on our choice of $T$. Let
$$
r_T = \phi(\alpha_T(T+t_1), z_T), \quad q_T = \phi(\alpha_T(-T-t_2), z_T).
$$
Inequalities \sref{11.1} and the following two relations imply that
\begin{equation}
\label{11.2.1} \phi(t, q_T) \in B(5\varepsilon, p), \quad t \in [0,
-\alpha_T(-T-t_2)],
\end{equation}
\begin{equation}
\label{11.2.2} \phi(t, r_T) \in B(5\varepsilon, p), \quad t \in [ -
\alpha_T(T+t_1), 0].
\end{equation}
Since $\mbox{$\cal{L}$} d_2\leq \varepsilon <1/2$ and $t_1,t_2$ are fixed, inequality
\sref{eqdef2.1} implies that if $T$ is large enough, then
\begin{equation}
\label{11.3} - \alpha_T(-T-t_2) \geq T/2, \quad \alpha_T(T+t_1) \geq
T/2.
\end{equation}
Since \sref{11.2.1}-\sref{11.3} imply that $\dist(\phi(t, q_T), p)
\leq 4\varepsilon$ for $0 \leq t \leq T/2$ and $\dist(\phi(t, r_T), p) \leq
4 \varepsilon$ for $0 \geq t \geq -T/2$ it follows that
$$
\dist(q_T, W^s_{loc, 4\varepsilon}(p)), \dist(r_T, W^u_{loc, 4\varepsilon}(p)) \to
0, \quad T \to+\infty.
$$
Since $q_T, r_T \in B(4\varepsilon, p) \setminus B(\varepsilon, p)$, we can choose
sequences $q_n = q_{T_n} \to q$ and $r_n = r_{T_n} \to r$ such that
$q, r\ne p$, $q \in W^s_{loc,4\varepsilon}(p)$, and $r \in
W^u_{loc,4\varepsilon}(p)$.
Denote by $O(q_n)$ the (closed) trajectory of the point $q_n$.
From Remark \ref{LPR} we know that $O(q_n)$ is a hyperbolic closed
trajectory.
Passing to a subsequence, if necessary, we may assume that the
values $\dim W^s(O(q_n))$ are the same for all $n$. Since
$$
\dim W^s(O(q_n)) + \dim W^u(O(q_n))= \dim M +1
$$
and
$$
\dim W^s(p) + \dim W^u(p) = \dim M,
$$
we see that at least one of the following inequalities holds:
\begin{equation}\notag
\dim W^s(O(q_n)) > \dim W^s(p)
\end{equation}
or
$$
\dim W^u(O(q_n)) > \dim W^u(p).
$$
Without loss of generality, we can assume that the first inequality
holds (in the other case we note that $O(q_n) = O(r_n)$ and consider
the vector field $-X$).
Denote $\sigma = \dim W^s(p)$. Consider the space $E^s_n = E^s(q_n)$
corresponding to inequalities \sref{8.5.1}, \sref{8.5.2}. Then the
following holds
$$
\dim E^s_n =\dim W^s(O(q_n)) - 1 \geq \sigma.
$$
Passing to a subsequence, if necessary, we may assume that $E^s_n
\to F^s \subset V_q$, where $V_q$ is the subspace in $T_q M$
orthogonal to $X(q)$ (here and below, we consider convergence of
linear spaces in the Grassman topology). Passing to the limit in
inequalities \sref{8.5.1}, we conclude that
\begin{equation}\notag
|P_{\phi(t, q)} D\phi(t, q) v^s| \leq C_2 \lambda_2^t
|v^s|,
\quad v^s\in F^s, \; t > 0.
\end{equation}
This inequality implies the inclusion $F^s \subset T W_q^s(p)$.
Hence,
$$
F^s \oplus \langle X(q) \rangle \subset T_q W^s(q),
$$
and $\dim W^s(q) \geq \sigma+1$. We get a contradiction which proves
Lemma \ref{LP8}.
\end{proof}
\subsection{Hyperbolicity of the chain-recurrent set}
Consider a point $y \in {\cal C}{\cal R}(X)$ that is not a rest point. Lemma
\ref{LP7} implies that there exists a sequence $x_n \in \Per(X)$
such that $x_n \to y$.
Consider the decomposition $V_{x_n} = E^s(x_n) +E^u(x_n)$
corresponding to inequalities \sref{8.5.1}, \sref{8.5.2}. Denote
$E^{s, u}_n = E^{s, u}(x_n)$. Passing if necessary to a subsequence,
we may assume that the dimensions $\dim E^s_n$ and $\dim E^u_n$ are
the same for all $n$. Since $y$ is not a rest point, $V_{x_n} \to
V_y$.
Since inequalities \sref{8.5.1} and \sref{8.5.2} hold for all closed
trajectories with the same constants $C_2$ and $\lambda_2$, standard
reasoning implies that the ``angles'' between $E^s_n$ and $E^u_n$
are uniformly separated from 0 (see, for instance,
\cite{PilSSBook}). So passing if necessary to a subsequence, we may
assume that $E^s_n \to E^s$ and $E^u_n \to E^u$.
Hence, $E^s \cap E^u = \{0\}$, $\dim (E^s + E^u) = \dim E^s + \dim
E^u = \dim V_y$, and $E^s + E^u = V_y$. Estimates \sref{8.5.1} and
\sref{8.5.2} for the points $x_n$ imply similar estimates for $y$.
Hence, the skew product flow \sref{Text11.1} is hyperbolic, and
Theorem 3 in Sacker and Sell \cite{SS2} implies that ${\cal C}{\cal R}(X)$ is
hyperbolic.
\subsection{No-cycle condition}
\label{secLipPer5}
In the previous two subsections we have proved that the vector field
$X$ (and its flow $\phi$) satisfies Axiom A$'$. It is known that in
this case, the nonwandering set of $X$ can be represented as a
disjoint union of a finite number of compact invariant sets (called
basic sets):
\begin{equation}
\label{spe} \Omega(X)=\Omega_1\cup\dots\cup\Omega_m,
\end{equation}
where each of the sets $\Omega_i$ is either a hyperbolic rest point
of $X$ or a hyperbolic set on which $X$ does not vanish and which
contains a dense positive semi-trajectory.
The basic sets $\Omega_i$ have stable and unstable ``manifolds'':
$$
W^s(\Omega_i)=\{x\in M:\;\mbox{dist}(\phi(t, x),\Omega_i)\to 0,
\quad t\to\infty\}
$$
and
$$W^u(\Omega_i)=\{x\in M:\;\mbox{dist}(\phi(t, x),\Omega_i)\to 0,
\quad t\to-\infty\}.
$$
If $\Omega_i$ and $\Omega_j$ are basic sets, we write
$\Omega_i\to\Omega_j$ if the intersection
$$
W^u(\Omega_i)\cap W^s(\Omega_j)
$$
contains a wandering point.
We say that $X$ has a 1-cycle if there is a basic set $\Omega_i$
such that $\Omega_i\to\Omega_i$.
We say that $X$ has a $k$-cycle if there are $k>1$ basic sets
$$
\Omega_{i_1},\dots,\Omega_{i_k}
$$
such that
$$
\Omega_{i_1}\to\dots\to\Omega_{i_k}\to\Omega_{i_1}.
$$
\begin{lem}
If $X\in\LipPerSh$, then $X$ has no cycles.
\end{lem}
\begin{proof} To simplify the presentation, we prove that $X$ has no 1-cycles
(in the general case, the idea is essentially the same, but the
notation is heavy).
To get a contradiction, assume that
$$
p\in(W^u(\Omega_i)\cap W^s(\Omega_i))\setminus \Omega(X).
$$
Then there are sequences of times $j_m,k_m\to\infty$ as $m\to\infty$
such that
$$
\phi(-j_m, p), \phi(k_m, p)\to\Omega_i,\quad m\to\infty.
$$
Since the set $\Omega_i$ is compact, we may assume that
$$
\phi(-j_m, p)\to q\in\Omega_i\;\mbox{ and }\; \phi(k_m, p)\to
r\in\Omega_i.
$$
Since $\Omega_i$ contains a dense positive semi-trajectory, there
exist points $s_m\to r$ and times $l_m>0$ such that
$\phi(l_m,s_m)\to q$ as $m\to\infty$.
Clearly, if we continue the mapping
$$
g(t) = \begin{cases} \phi(t, p), & t \in [0, k_m],\\
\phi(t - k_m, s_m), & t \in [k_m, k_m + l_m],\\
\phi(t-j_m-k_m-l_m, p), & t \in [k_m+l_m, k_m + l_m +j_m],
\end{cases}
$$
periodically with period $k_m+l_m+j_m$, we get a periodic
$d_m$-pseudotrajectory of $X$ with $d_m\to 0$ as $m\to\infty$.
Since $X\in\LipPerSh$, there exist points $p_m \in \Per(X)$ (for $m$
large enough) such that $p_m\to p$ as $m\to\infty$, and we get the
desired contradiction with the assumption that $p\notin\Omega(X)$.
The lemma is proved.
\end{proof}
\subsection{$\Omega$-stability implies Lipschitz periodic
shadowing}\label{secLipPer6}
The proof of Lemma \ref{lemLipPer2} is similar to the corresponding
proof in \cite{OPT}, where the case of diffeomorphisms is
considered. In the present article we give the most important steps
and leave the details to the reader.
\begin{proof}[Proof of Lemma \ref{lemLipPer2}]
Let us formulate several auxiliary definitions and statements.
Let us say that a vector field $X$ has the Lipschitz shadowing
property on a set $U$ if there exist positive constants $\mbox{$\cal{L}$},d_0$
such that if $g(t)$ with $\{g(t):\;t\in\mathds{R}\}\subset U$ is a
$d$-pseudotrajectory (in our standard sense:
$$
\mbox{dist}(g(\tau+t),\phi(t,g(\tau)))<d,\quad \tau\in\mathds{R},t\in[0,1])
$$
with $d\leq d_0$, then there exists a point $p\in U$ and a
reparametrization $\alpha$ satisfying inequality \sref{eqdef2.1} such
that
\begin{equation}
\label{os1} \mbox{dist}(g(t),\phi(\alpha(t),p))<\mbox{$\cal{L}$} d,\quad t\in\mathds{R}.
\end{equation}
We say that a vector field $X$ is expansive on a set $U$ if there
exist positive numbers $a$ (expansivity constant) and $\delta$ such
that if two trajectories $\{\phi(t,p):\;t\in\mathds{R}\}$ and
$\{\phi(t,q):\;t\in\mathds{R}\}$ belong to $U$ and there exists a
continuous real-valued function $\alpha(t)$ such that
$$
\mbox{dist}(\phi(\alpha(t),q),\phi(t,p))\leq a,\quad t\in\mathds{R},
$$
then $p=\phi(\tau,q)$ for some real $\tau \in (-\delta, \delta)$.
Let $X$ be an $\Omega$-stable vector field. Consider the
decomposition \sref{spe} of $\Omega(X)$. We will refer to the
following well-known statement \cite{Palmer1}.
\begin{thm}\label{thm4} If $\Omega_i$ is a basic set, then there exists a
neighborhood $U$ of $\Omega_i$ such that $X$ has the Lipschitz
shadowing property on $U$ and is expansive on $U$.
\end{thm}
We also need the following two lemmas. Analogs of these lemmas were
proved for diffeomorphisms in \cite{PilTarSak}; the proofs for flows
are the same.
\begin{lem}\label{lemOm3} For any
neighborhood $U$ of the nonwandering set $\Omega(X)$ there exist
positive numbers $B,d_1$ such that if $g(t)$ is a
$d$-pseudotrajectory of $\phi$ with $d\leq d_1$ and
$$
g(t)\notin U,\quad t\in[\tau,\tau+l],
$$
for some $l>0$ and $\tau\in\mathds{R}$, then $l\leq B$.
\end{lem}
\begin{lem}\label{lemOm4} Assume that the vector field $X$ is
$\Omega$-stable. Let $U_1,\dots,U_m$ be disjoint neighborhoods of
the basic sets $\Omega_1,\dots,\Omega_m$. There exist neighborhoods
$V_j\subset U_j$ of the sets $\Omega_j$ and a number $d_2>0$ such
that if $g(t)$ is a $d$-pseudotrajectory of $X$ with $d\leq d_2$,
$g(\tau)\in V_j$ and $g(\tau+t_0)\notin U_j$ for some
$j\in\{1,\dots,m\}$, some $\tau \in \mathds{R}$ and some $t_0>0$, then
$g(\tau+t)\notin V_j$ for $t\geq t_0$.
\end{lem}
Now we pass to the proof itself.
Apply Theorem \ref{thm4} and Lemmas \ref{lemOm3}, \ref{lemOm4} and
find disjoint neighborhoods $W_1,\dots,W_m$ of the basic sets
$\Omega_1,\dots,\Omega_m$ such that
\begin{itemize}
\item[(i)] $X$ has the Lipschitz shadowing property on each $W_j$ with
the same constants $\mbox{$\cal{L}$},d^*_0$;
\item[(ii)] $X$ is expansive on each $W_j$ with the same expansivity
constants $a$, $\delta$.
\end{itemize}
Find neighborhoods $V_j,U_j$ of $\Omega_j$ (and reduce $d^*_0$, if
necessary) so that the following properties are fulfilled:
\begin{itemize}
\item $V_j\subset U_j\subset W_j,\quad j=1,\dots,m$;
\item the statement of Lemma \ref{lemOm4} holds for $V_j$ and $U_j$ with some
$d_2>0$;
\item the $\mbox{$\cal{L}$} d^*_0$-neighborhoods of $U_j$ belong to $W_j$.
\end{itemize}
Apply Lemma \ref{lemOm3} to find the corresponding constants $B,d_1$
for the neighborhood $V_1\cup\dots\cup V_m$ of $\Omega(X)$.
We claim that $X$ has the Lipschitz periodic shadowing property with
constants $\mbox{$\cal{L}$},d_0$, where
$$
d_0=\min\left(d^*_0,d_1,d_2,\frac{a}{2\mbox{$\cal{L}$}}\right).
$$
Take a $\mu$-periodic $d$-pseudotrajectory $g(t)$ of $X$ with $d\leq
d_0$. Without loss of generality we can assume that $\mu > \delta$
(since $\mu$ is not necessarily the minimal period). Lemma
\ref{lemOm3} implies that there exists a neighborhood $V_j$ such
that the pseudotrajectory $g(t)$ intersects $V_j$; shifting time, we
may assume that $g(0)\in V_j$.
In this case, $\{g(t):\;t\in\mathds{R}\}\subset U_j$. Indeed, if
$g(t_0)\notin U_j$ for some $t_0$, then $g(t_0+k\mu)\notin U_j$ for
all $k$. It follows from Lemma \ref{lemOm4} that if $t_0+k\mu>0$,
then $g(t)\notin V_j$ for $t\geq t_0+k\mu$, and we get a
contradiction with the periodicity of $g(t)$ and the inclusion
$g(0)\in V_j$.
Thus, there exists a point $p$ such that inequalities (\ref{os1})
hold for some reparametrization $\alpha$ satisfying inequality
\sref{eqdef2.1}. Let us show that either $p$ is a rest point or the
trajectory of $p$ is closed. By the choice of $U_j$ and $W_j$,
$\phi(t,p)\in W_j$ for all $t\in\mathds{R}$. Let $q=\phi(\mu,p)$.
Inequalities (\ref{os1}) and the periodicity of $g(t)$ imply that
$$
\mbox{dist}(g(t),\phi(\alpha(t+\mu)-\mu,q))=
$$
$$
\mbox{dist}(g(t+\mu),\phi(\alpha(t+\mu),p))\leq \mbox{$\cal{L}$} d,\quad t\in\mathds{R}.
$$
Thus,
$$
\mbox{dist}(\phi(\alpha(t),p),\phi(\alpha(t+\mu)-\mu,q))\leq 2\mbox{$\cal{L}$} d\leq
a,\quad t\in\mathds{R},
$$
which implies that
$$
\mbox{dist}(\phi(\theta,p),\phi(\beta(\theta),q))\leq 2\mbox{$\cal{L}$} d\leq
a,\quad \theta\in\mathds{R},
$$
where $\beta(\theta)=\alpha(\alpha^{-1}(\theta)+\mu)-\mu$.
Since $\phi(t,p)\in W_j$ for all $t\in\mathds{R}$, our expansivity
condition on $W_j$ implies that $q=\phi(\tau,p)$ for some $\tau \in
(-\delta, \delta)$.
This completes the proof.
\end{proof}
|
\section{Introduction}
\label{intro}
The relationship between the radio and X-ray emission in black-hole binaries has been extensively studied, see, e.g., \citet*{gfp03}, \citet*{mhd03}, \citet{corbel00,corbel03,corbel04,corbel08}, \citet{yc05}, \citet{soleri10}, \citet{coriat11}. The qualitative form of the radio/X-ray correlation in Cyg X-1 is rather typical of black-hole binaries. In the hard spectral state, there is a positive correlation, as in other black-hole binaries \citep{corbel00,corbel04,gfp03}. At transitions to the soft state of Cyg X-1, the radio flux strongly decreases. This is also similar to other black-hole binaries. In particular, the radio flux strongly declines at the transition from the hard state to a soft one in GX 339--4 \citep{corbel00}, XTE J1650--500 \citep{corbel04}, and XTE J1859+226 \citep*{fbg04}. After the soft-state decline, the radio flux from black-hole transients may strongly increase in an intermittent manner, sometimes up to the overall maximum for a source, e.g., \citet{fbg04}, \citet{gallo04}, which is probably due to the ejection of radio-emitting blobs (e.g., \citealt{y09}). A similar overall radio-soft X-ray correlation with all of the above states is also seen in Cyg X-3 \citep*{szm08}, which, based on its X-ray properties \citep{sz08,hj09}, is probably a black-hole binary. The soft-state radio flaring is not seen in Cyg X-1, apparently due to the limited range of its X-ray flux.
Determining the relationship between the radio and X-ray emission is of major importance to our understanding of the physics of accretion and outflow in black-hole binaries. The bulk of the radio emission is clearly from the jet. There are two possible origins of the correlation of the radio emission with X-rays in black-hole binaries. One is that the level of X-ray emission in the hard spectral state, related to the accretion rate, is in turn related to the rate of the outflow forming the jet (e.g., \citealt*{mirabel98,corbel00,hs03,ycn05}). Another is that the X-ray emission of black-hole binaries is dominated by emission of the jet (e.g., \citealt*{vadawale01,georganopoulos02,markoff03,falcke04}).
The usual approach to studying the radio/X-ray correlation is to consider the hard spectral state only and to study the X-ray emission in a relatively narrow band, typically 2--10 keV or 3--9 keV. The relation with the radio flux is represented by a power law (e.g., \citealt{gfp03,mhd03,coriat11}) and theoretical consequences of the value of the power-law index are then are considered (e.g., \citealt{mhd03,hs03,coriat11}). Sometimes infrared or optical emission is also considered in addition to radio (e.g., \citealt{coriat09}). A common implicit assumption is that the correlation is the same in other X-ray bands as in the band considered. Attention is rarely paid to the correlation of the radio flux with the bolometric flux (dominated by the X-rays), a quantity which is crucial for the first class of models listed above. Here, we consider the correlation of the radio emission with X-rays in several energy bands in the case of the archetypical and well studied black-hole binary Cyg X-1. We consider all its spectral states, from hard to soft.
We first study the X-ray spectral variability patterns, in particular the dependence of the bolometric flux on the X-ray spectral slope, as well as on the fluxes in various X-ray bands. We calculate the spectra and bolometric fluxes averaged over a 1-day time scale using the monitoring by the All-Sky Monitor (ASM) on board {\it Rossi X-ray Timing Explorer\/} ({\textit{RXTE}}; \citealt*{brs93,levine96}) simultaneous with the monitoring either by the {\textit{Swift}}\/ Burst Alert Telescope (BAT; \citealt{barthelmy05,m05}) or by the Burst and Transient Source Experiment (BATSE; \citealt{harmon02}) on board of {\it Compton Gamma Ray Observatory\/} ({\textit{CGRO}}).
The discovery of substantial orbital modulation of the radio emission \citep*{pfb99} shows that free-free absorption in the stellar wind \citep{gies03} of the donor affects the observed radio emission and so complicates the interpretation of the observations. The level of attenuation of the intrinsic radio flux depends on the height along the jet where the bulk of the emission at a given frequency emerges \citep{sz07}, which is difficult to determine. In spite of this complication, given its importance and brightness and the wealth of available monitoring data, the study of Cyg X-1 source provides valuable insights into the relationship between the radio and X-ray emission in black-hole binaries.
\begin{figure}
\centerline{\includegraphics[height=\columnwidth,angle=-90]{coverage.eps}}
\caption{The time and photon energy/wavelength coverage by the instruments used in this work. The periods dominated by the hard state are indicated by the horizontal black lines.
}
\label{f:coverage}
\end{figure}
\section{The data}
\label{data}
We use monitoring data from the {\textit{RXTE}}/ASM, which has three channels at energies of 1.5--3 keV, 3--5 keV and 5--12 keV. The {\textit{CGRO}}/BATSE monitoring data used here are the same as those in \citet{z02}, hereafter Z02, which were obtained using the Earth occultation analysis technique \citep{harmon02}. They are given as energy fluxes in the 20--100 keV and 100--300 keV channels. For the {\textit{Swift}}/BAT, we use 14--195 keV 8-channel light curves created for this work. The channels are between energies of 14, 20, 24, 35, 50, 75, 100, 150 and 195 keV. The data used here from the ASM, BATSE and BAT are for MJD 50087--55700, 50087--51686 and 53355--55469, respectively. Fig.\ \ref{f:coverage} shows the time and energy coverage of the instruments used here.
The ASM count rates are converted into energy fluxes using the matrix of Z02. We convert the BAT 8-channel data into energy fluxes by scaling to the Crab spectrum using the results of \citet{rv01} up to 100 keV, and at higher energies, those of \citet{jr09}, see \citet*{z11}, hereafter ZPS11, for details.
\begin{figure*}
\centerline{\includegraphics[width=15cm]{cygx1_lc_1_renorm.eps}}
\centerline{\includegraphics[width=15cm]{cygx1_lc_2_renorm.eps}}
\caption{The light curves showing daily averages from the ASM (black error bars, 1.5--12 keV) and Ryle/AMI (red points, 15 GHz), normalized to their respective (un-weighted) average values within the hard state of MJD 53880--55375, an interval delineated in panel (b) by the vertical dashed lines. The averages are $\langle F\rangle= 20.7$ s$^{-1}$ and 12.7 mJy, respectively. For clarity, the uncertainty of the radio points is not shown. During some soft states, the radio fluxes were below the range of $F/\langle F\rangle$ shown. The 2010/11 soft state can be seen in panel (b).
}
\label{f:lc}
\end{figure*}
\begin{figure}
\centerline{\includegraphics[width=7cm]{flux_gamma.eps}}
\centerline{\includegraphics[width=7cm]{asm_g_radio.eps}}
\caption{Relationship between the 3--12 keV photon index calculated from the ASM data and (a) the bolometric flux calculated from the ASM and BATSE data (red points) and from the ASM and BAT data (black points), and (b) the daily-averaged radio flux at 15 GHz. The dotted and dashed vertical lines show our assumed boundaries between the hard and intermediate states, and intermediate and soft states, respectively. For clarity of display, we do not plot the error bars and include only the data with low index uncertainty, $\Delta \Gamma(3$--$12\, {\rm keV})<0.3$. The decreasing value of $\Gamma$ on the horizontal axis corresponds to the increasing spectral hardness.
}
\label{f:gamma}
\end{figure}
We divide the daily data from the ASM and those from other instruments simultaneous with the ASM monitoring into three groups based on their 3--12 keV photon index, $\Gamma$, which we calculate using the method of Z02 and ZPS11. The 3--12 keV energy range in Cyg X-1 is only weakly affected by absorption and is thus a good measure of the actual spectral slope in that range. We consider separately data with $\Gamma<1.9$ (hard state), $1.9<\Gamma<2.3$ (intermediate state) and $\Gamma>2.3$ (soft state). These boundaries differ slightly from those used in Z02; here, in particular, our definition of the hard state is more stringent and more suitable for studies of the correlation with the radio emission. We take $\Gamma_{\rm av}=1.7$, 2.1 and 2.5 as the average indices of the three respective groups. We stress that the soft state in Cyg X-1 usually contains a high energy tail much stronger than that typical of low-mass black-hole binaries in the soft state. In those binaries, the tail often virtually disappears (e.g., \citealt{gd04}) as well as the radio emission being strongly quenched (unlike the case in Cyg X-1). Thus the soft state of Cyg X-1 might also be called `soft intermediate' in the context of low-mass black-hole binaries.
We use 15-GHz data from the Ryle Telescope and the AMI Large Array. The AMI Large Array is the re-built and reconfigured Ryle Telescope. The monitoring by the Ryle Telescope was carried out during MJD 50226--53902, and the AMI monitoring of Cyg X-1 started on MJD 54573, the data used here being until MJD 55540. The new correlator has a useful bandwidth of about 4 GHz (compared with 0.35 GHz); but the effective centre frequency is similar, and in any case the radio spectrum of Cyg X-1, at least in the hard state, is known to be very flat \citep{fender00}. The data are subject to variations in the flux calibration of about 10 per cent from one day to another.
We also use the 2.25 GHz and 8.3 GHz monitoring data\footnote{ftp://ftp.gb.nrao.edu/pub/fghigo/gbidata/gdata/} from the Green Bank Interferometer (GBI). According to the web page description, the data are error-dominated below 15 mJy. However, we have found that using only the data $>F_{\rm R,min}=15$ mJy with the X-ray fluxes introduces a bias, making the resulting radio/X-ray distributions artificially flat. In fact, the values of the Spearman's correlation coefficients increase and the associated null hypothesis probabilities decrease when we decrease $F_{\rm R,min}$ (i.e., include more low-flux data). After studying different choices, we have opted for using the data $>F_{\rm R,min}=10$ mJy for fits. However, we show all the radio points in the figures. The GBI monitoring of Cyg X-1 lasted MJD 50409--51823.
\section{Results}
\label{results}
\subsection{The light curves}
\label{lc}
Fig.\ \ref{f:lc} presents the daily-average light curves from the ASM (1.5--12 keV) and Ryle/AMI (15 GHz). Both have been renormalized to the average count rate/flux over the long hard state of MJD 53880--55375 (discussed in detail by ZPS11). We see that the radio and the 1.5--12 keV X-ray fluxes closely follow each other during hard/low X-ray states but the radio emission is suppressed during the soft/high X-ray states. The radio emission during the latter states can become extremely low, but also can flare during the soft state back to the hard-state level, e.g., during the long soft state of MJD 52160--52570. Fig.\ 1 in ZPS11 shows also the 15--50 keV BAT light curve, normalized in the same way. We quantify the radio/X-ray correlation in Section \ref{corr} below.
During the 15 GHz monitoring of Cyg X-1, there were four events when the flux rose above 50 mJy. We show their light curves in detail in Appendix \ref{flares}.
\subsection{X-ray spectra and variability patterns}
\label{xray}
Appendix \ref{bol} gives details of our method of calculating the bolometric flux (i.e., that integrated over all energies) based on the ASM, BATSE and BAT data. Fig.\ \ref{f:gamma}(a) shows the relationship between the resulting bolometric flux and $\Gamma(3$--12 keV). In order to allow comparison with commonly plotted hardness-flux diagrams, the horizontal axis shows $\Gamma$ decreasing to the right. We see a positive correlation of $F_{\rm bol}$ with $\Gamma$ in the intermediate state, with the Spearman's rank-order correlation coefficient, $r_{\rm s}\simeq 0.16$, and the corresponding probability that this $r_{\rm s}$ is due to a chance, $P_{\rm s}\simeq 1\times 10^{-14}$. The correlation continues to the hard state for $F_{\rm bol}>30$ keV cm$^{-2}$ s$^{-1}$, within which $r_{\rm s}\simeq 0.52$, $P_{\rm s}\simeq 2\times 10^{-17}$. In the hard state at lower $F_{\rm bol}$, it changes by a factor of a few at a given $\Gamma$ without any apparent trend.
The soft state data, at $\Gamma(3$--12 keV$)\mathrel{\copy\simgreatbox} 2.3$, show no clear overall correlation. In this state, $\Gamma(3$--12 keV) is not uniquely correlated with the bolometric flux. We note that in a number of occurrences of the soft state $F_{\rm bol}$ was lower than that in the high-flux hard state. This happened mostly during the 2010/11 soft state, when the ASM fluxes reached relatively low values, see Figs.\ \ref{f:lc}(b), \ref{f:lc_bol}(b).
Fig.\ \ref{f:lc_bol} presents the light curve of $F_{\rm bol}$. We see changes up to a factor of $\sim$10. We clearly see the superorbital periodicities of Cyg X-1, $\sim$150 d in Fig.\ \ref{f:lc_bol}(a) and $\sim$300 d in Fig.\ \ref{f:lc_bol}(b), see, e.g., \citet*{pzi08}, ZPS11.
\begin{figure*}
\centerline{\includegraphics[height=6.3cm]{cygx1_lc_bol_1.eps}
\includegraphics[height=6.3cm]{cygx1_lc_bol_2.eps}}
\caption{The light curves showing daily averages of the bolometric flux, estimated based on (a) the ASM+BATSE data, and (b) the ASM+BAT data. The red, green and blue symbols correspond to $\Gamma(3$--$12\,{\rm keV}) <1.9$ (hard state), $1.9<\Gamma<2.3$ (intermediate state) and $\Gamma>2.3$ (soft state), respectively.
}
\label{f:lc_bol}
\end{figure*}
\begin{figure*}
\centerline{
\includegraphics[height=3.5cm]{1.5_3_fbol.eps}
\includegraphics[height=3.5cm]{3_5_fbol.eps}
\includegraphics[height=3.5cm]{5_12_fbol.eps}
\includegraphics[height=3.5cm]{14_20_fbol.eps}
\includegraphics[height=3.5cm]{75_100_fbol.eps}}
\centerline{
\includegraphics[height=7.81cm]{1.5_3_radio15.eps}
\includegraphics[height=7.81cm]{3_5_radio15.eps}
\includegraphics[height=7.81cm]{5_12_radio15.eps}
\includegraphics[height=7.81cm]{14_20_radio15.eps}
\includegraphics[height=7.81cm]{75_100_radio15.eps}}
\caption{Top panels (a--e): the X-ray variability patterns of Cyg X-1 as illustrated by the relationship between the bolometric flux (calculated using the ASM+BAT data) and the flux in a given band. Bottom panels (f--j): the relationship between the 15 GHz radio flux and the X-ray fluxes measured by either the ASM ($\leq 12$ keV) or BAT ($\geq 14$ keV) (daily averages).
The red, green and blue symbols correspond to data for which the 3--12 keV photon index is $\Gamma<1.9$ (hard state), $1.9<\Gamma<2.3$ (intermediate state) and $\Gamma>2.3$ (soft state), respectively. The lines show the best-fit power-law fits [equation (\ref{fit})] to the hard state data. The units of the X-ray flux are keV cm$^{-2}$ s$^{-1}$ and the error bars on the radio data are not shown for clarity. Only data with low index uncertainty, $\Delta \Gamma(3$--$12\, {\rm keV})<0.3$ are plotted.
}
\label{f:X_radio}
\end{figure*}
Figs.\ \ref{f:X_radio}(a--e) illustrate the X-ray spectral variability patterns of Cyg X-1 by showing the relationship between the bolometric flux and the flux in a given band, $F_{\rm X}$. We have fitted these relationships in the hard state (red in Figs.\ \ref{f:lc_bol}--\ref{f:hot_bol}) by a power law, $F_{\rm bol}\propto F_{\rm X}^{p'}$ (see Appendix \ref{fits}). The indices found are given in Table \ref{t:fit}. We see that $p'\simeq 1$ in the 14--150 keV range. This corresponds to a spectrum of constant shape moving up and down in normalization only. On the other hand, the dependencies are substantially weaker than direct proportionality in the 1.5--3 keV and 150--195 keV bands.
The intermediate state (green in Figs.\ \ref{f:lc_bol}--\ref{f:hot_bol}) shows a different variability pattern. Up to 12 keV, the intermediate state points are quite similar to those in the hard state, though they extend to higher fluxes. However, the dependence changes sign at $E\mathrel{\copy\simgreatbox} 20$ keV and it becomes increasingly negative (i.e, the local X-ray flux is anti-correlated with $F_{\rm bol}$), as shown in Fig.\ \ref{f:X_radio}(e) for 75--100 keV.
The soft state (blue in Figs.\ \ref{f:lc_bol}--\ref{f:hot_bol}) shows an approximate dependence of $F_{\rm bol}\mathrel{\copy\simpropbox} F_{\rm X}$ over all the observed energy range. However, the scatter increases with the increasing energy.
For coronal models (e.g., \citealt{mf02,merloni03}), the flux, $F_{\rm hot}$, due to emission of coronal hot electrons is an important quantity. It is predicted to be constant for coronae above gas-pressure dominated discs, and to decrease as $\dot M^{-1/2}$ for radiation-pressure dominated discs, see a discussion in \citet{mhd03}. It is equal to the bolometric X-ray flux minus the disc blackbody contribution. Ideally, to calculate $F_{\rm hot}$ accurately one should fit each spectrum with a model containing a disc blackbody, and then subtract its total flux from the resulting model bolometric flux. Given our method of determining $F_{\rm bol}$, this is not possible. Instead, we use the flux above an energy approximately dividing the spectral region dominated by the blackbody and that dominated by Comptonization. Specifically, we assume $F_{\rm hot}= F(>\!\!3\,{\rm keV})$ in the soft state, and $=F(>\!\!1.5\,{\rm keV})$ in the intermediate and hard states. These choices are based on the Cyg X-1 spectra shown in \citet{zg04}. Fig.\ \ref{f:hot_bol} shows the resulting estimate of the coronal emission as a fraction of $F_{\rm bol}$. We see it is approximately constant in the soft state.
\subsection{Radio/X-ray correlation}
\label{corr}
\begin{table*}
\centering
\caption{The values of the fit coefficients (the radio flux normalization, $F_{\rm R,0}$, calculated at $\langle F_{\rm X}\rangle$, and the index, $p$), equation (\ref{fit}), the Spearman's correlation coefficient, $r_{\rm s}$, and its null hypothesis probability, $P_{\rm s}$, for the radio/X-ray correlation. We also give the value of the index, $p'$, for the $F_{\rm bol}$--$F_{\rm X}$ correlation in the hard state, and the $p/p'$ ratio. Only the hard state data have been fitted except for the case of $F({\rm 15\, GHz})$--$F_{\rm hot}$, where the data in all three states have been included. The uncertainties are given at $1\sigma$.
}
\begin{tabular}{cccccccccc}
\hline
Correlation & $F_{\rm R,0}\,$[mJy] &$\langle F_{\rm X}\rangle\,$[keV cm$^{-2}$ s$^{-1}$] & $p$ & $p'$ & $p/p'$ & $r_{\rm s}$ & $P_{\rm s}$ & $F_{\rm R,min}\,$[mJy] \\
\hline
ASM\\
$F({\rm 2.25\, GHz})$--$F(1.5$--$3\,{\rm keV})$ & $15.9\pm 0.2$ & 2.09 & $0.75\pm 0.06$ & $0.59\pm 0.01$ & $1.27\pm 0.08$ & 0.26 & $3\times 10^{-13}$ & 10\\
$F({\rm 8.3\, GHz})$--$F(1.5$--$3\,{\rm keV})$ & $17.0\pm 0.2$ & 2.11 & $0.79\pm 0.06$ & {\tt "} & $1.34\pm 0.07$ & 0.30 & $3\times 10^{-17}$ & 10\\
$F({\rm 15\, GHz})$--$F(1.5$--$3\,{\rm keV})$ & $12.2\pm 0.1$ & 2.10 & $0.91\pm 0.03$ & {\tt "} & $1.54\pm 0.04$ & 0.49 & $<10^{-20}$ & 5\\
$F({\rm 2.25\, GHz})$--$F(3$--$5\,{\rm keV})$ & $15.9\pm 0.2$ & 1.52 & $1.07\pm 0.09$ & $0.84\pm 0.01$ & $1.27\pm 0.09$ & 0.25 & $2\times 10^{-12}$ & 10\\
$F({\rm 8.3\, GHz})$--$F(3$--$5\,{\rm keV})$ & $17.0\pm 0.2$ & 1.53 & $1.13\pm 0.09$ & {\tt "} & $1.35\pm 0.08$ & 0.27 & $1\times 10^{-14}$ & 10\\
$F({\rm 15\, GHz})$--$F(3$--$5\,{\rm keV})$ & $12.2\pm 0.1$ & 1.54 & $1.34\pm 0.05$ & {\tt "} & $1.59\pm 0.04$ & 0.51 & $<10^{-20}$ & 5\\
$F({\rm 2.25\, GHz})$--$F(5$--$12\,{\rm keV})$ & $15.9\pm 0.2$ & 3.10 & $1.28\pm 0.12$ & $0.90\pm 0.01$ & $1.42\pm 0.09$ & 0.23 & $6\times 10^{-11}$ & 10\\
$F({\rm 8.3\, GHz})$--$F(5$--$12\,{\rm keV})$ & $17.0\pm 0.2$ & 3.10 & $1.35\pm 0.12$ & {\tt "} & $1.50\pm 0.09$ & 0.24 & $1\times 10^{-11}$ & 10\\
$F({\rm 15\, GHz})$--$F(5$--$12\,{\rm keV})$ & $12.2\pm 0.1$ & 3.09 & $1.59\pm 0.06$ & {\tt "} & $1.77\pm 0.04$ & 0.50 & $<10^{-20}$ & 5\\
\hline
BAT\\
$F({\rm 15\, GHz})$--$F(14$--$20\,{\rm keV})$ & $12.6\pm 0.2$ & 1.75 & $1.68\pm 0.12$ & $1.03\pm 0.01$ & $1.63\pm 0.07$ & 0.56 & $<10^{-20}$ & 5\\
$F({\rm 15\, GHz})$--$F(20$--$24\,{\rm keV})$ & $12.6\pm 0.2$ & 1.00 & $1.75\pm 0.13$ & $1.04\pm 0.01$ & $1.68\pm 0.07$ & 0.56 & $<10^{-20}$ & 5\\
$F({\rm 15\, GHz})$--$F(24$--$35\,{\rm keV})$ & $12.6\pm 0.2$ & 2.30 & $1.78\pm 0.13$ & $1.05\pm 0.01$ & $1.69\pm 0.07$ & 0.53 & $<10^{-20}$ & 5\\
$F({\rm 15\, GHz})$--$F(35$--$50\,{\rm keV})$ & $12.6\pm 0.2$ & 2.47 & $1.78\pm 0.15$ & $1.05\pm 0.01$ & $1.69\pm 0.08$ & 0.47 & $2\times 10^{-20}$ & 5\\
$F({\rm 15\, GHz})$--$F(50$--$75\,{\rm keV})$ & $12.6\pm 0.2$ & 3.04 & $1.77\pm 0.16$ & $1.05\pm 0.01$ & $1.69\pm 0.09$ & 0.42 & $4\times 10^{-16}$ & 5\\
$F({\rm 15\, GHz})$--$F(75$--$100\,{\rm keV})$ & $12.6\pm 0.3$ & 2.22 &$1.74\pm 0.19$ & $1.04\pm 0.02$ & $1.67\pm 0.11$ & 0.33 & $2\times 10^{-10}$ & 5\\
$F({\rm 15\, GHz})$--$F(100$--$150\,{\rm keV})$ & $12.6\pm 0.3$ & 2.92 &$1.64\pm 0.23$ & $0.99\pm 0.02$ & $1.66\pm 0.14$ & 0.24 & $4\times 10^{-6}$ & 5\\
$F({\rm 15\, GHz})$--$F(150$--$195\,{\rm keV})$ & $12.6\pm 0.3$ & 1.63 &$0.87\pm 0.18$ & $0.48\pm 0.02$ & $1.81\pm 0.21$ & 0.12 & $0.03$ & 5\\
\hline
ASM+BAT\\
$F({\rm 15\, GHz})$--$F_{\rm bol}$ & $12.6\pm 0.2$ & 32.6 & $1.68\pm 0.11$ & 1 & -- & 0.58 & $<10^{-20}$ & 5\\
$F({\rm 15\, GHz})$--$F_{\rm hot}$ & $11.4\pm 0.2$ & 26.1 & $1.73\pm 0.07$ & -- & -- & 0.55 & $<10^{-20}$ & 1\\
\hline
ASM+BATSE\\
$F({\rm 2.25\, GHz})$--$F_{\rm bol}$ & $15.8\pm 0.2$ & 30.7 & $1.49\pm 0.15$ & 1 & -- & 0.24 & $2\times 10^{-9}$& 10\\
$F({\rm 8.3\, GHz})$--$F_{\rm bol}$ & $17.0\pm 0.2$ & 30.6 & $1.58\pm 0.18$ & 1 & -- & 0.19 & $9\times 10^{-7}$ & 10 \\
$F({\rm 15\, GHz})$--$F_{\rm bol}$ & $12.2\pm 0.2$ & 30.2 & $1.81\pm 0.11$ & 1 & -- & 0.46 & $<10^{-20}$ & 5\\
\hline
\end{tabular}
\label{t:fit}
\end{table*}
We now consider correlations between the X-ray and radio fluxes. Figs.\ \ref{f:X_radio}(f--j) shows the relationships between the X-ray flux in selected ASM and BAT bands and the 15 GHz radio flux. The colours identify the spectral state. The reason that there are fewer points in the BAT range (i--j) is because each requires simultaneous BAT, 15 GHz radio and ASM coverage (the last to determine the spectral state).
It can be seen that the hard-state fluxes (shown in red) approximately follow a power-law relationship, $F({\rm 15\,GHz})\propto F_{\rm X}^{p}$, but the power-law index is a strong function of the X-ray energy band. The fitted parameters (see Appendix \ref{fits} for the fitting method) are given in Table \ref{t:fit}. The high significance of the correlations is confirmed by the values of the Spearman's correlation coefficient, $r_{\rm s}$, and the null hypothesis probability, $P_{\rm s}$, also given in Table \ref{t:fit}. The index, $p$, changes from $\simeq 0.9$ at 1.5--3 keV to $\simeq 1.7$--1.8 in the 14--150 keV range.
The variation of $p$ with energy is summarised in Fig.\ \ref{f:index}. The low value of the index in the 150--195 keV band index is apparently due the fact that this energy range corresponds to where the high-energy cut-off of the spectra starts to have a major effect. Thus much of the variability is associated with a different physical process -- changes in the cut-off energy, apparently due to a variable electron temperature.
\begin{figure}
\centerline{\includegraphics[height=6.3cm]{hot_bol_ratio.eps}}
\caption{The flux in the Comptonization part of the spectrum (approximated as described in Section \ref{xray}) divided by the bolometric flux, based on the ASM+BAT data. The red, green and blue symbols correspond to $\Gamma(3$--$12\, {\rm keV}) <1.9$ (hard state), $1.9<\Gamma<2.3$ (intermediate state) and $\Gamma>2.3$ (soft state), respectively.
}
\label{f:hot_bol}
\end{figure}
\begin{figure}
\centerline{\includegraphics[width=7cm]{rx_index.eps}}
\caption{The power-law index of the hard-state correlation between $F(15\,{\rm GHz})$ and $F_{\rm X}$ as a function of the photon energy, see Table \ref{t:fit}. The dotted horizontal lines show the uncertainty range of $p$ for the Comptonization flux, $F_{\rm hot}$.
}
\label{f:index}
\end{figure}
The relationship between the radio flux, $F(\rm 15\, GHz)$, and the X-ray hardness is illustrated in Fig.\ \ref{f:gamma}(b). There is a strong correlation in the hard state, $r_{\rm s} = 0.22$, $P_{\rm s} = 2\times 10^{-19}$, which continues up to $\Gamma\sim 2.1$. When the X-ray spectra are soft, there is a strong decrease in the radio emission; for $\Gamma>2.3$ the (anti-)correlation has $r_{\rm s} = -0.57$, $P_{\rm s} = 2\times 10^{-37}$.
Fig.\ \ref{f:bol_radio15}(a--b) shows the relationship between the 15-GHz flux and the bolometric flux. The approximate power-law relationship in the hard state now extends to most of the intermediate state, though it breaks down in the soft state. The fit results for the hard state are given in the third part of Table \ref{t:fit}. Although the low energy data are included in $F_{\rm bol}$, the correlation is much steeper ($p_{\rm bol}=1.68\pm 0.11$) than at $E\leq 5$ keV, where the power law index was $p\simeq 0.9$--1.3.
As $p_{\rm bol} = {\rm d} \ln F({\rm 15\,GHz}) / {\rm d} \ln F_{\rm bol}$, while $p = {\rm d} \ln F({\rm 15\, GHz}) / {\rm d} \ln F_{\rm X}$ and $p' = {\rm d} \ln F_{\rm bol} / {\rm d} \ln F_{\rm X}$ (where the differential coefficients are those of the fitted functional dependencies), in the absence of hidden systematic variability patterns one might expect $p_{\rm bol}\simeq p/p'$, and indeed it can be seen from Table \ref{t:fit} that this is very close to being the case, especially at $E\mathrel{\copy\simgreatbox} 5$ keV.
We note that Figs.\ \ref{f:X_radio}(f--j) show that the radio flux in the states with a strong X-ray blackbody contributions, i.e., intermediate and soft, depends as a power law on the emission at high energies, $\mathrel{\copy\simgreatbox} 5$ keV. This indicates that the level of the radio emission is associated with the high-energy electrons in the X-ray emitting region, presumably a corona in the intermediate and soft states (e.g., \citealt{gierlinski99}). To test this hypothesis, we have considered the dependence of the 15 GHz flux on the integrated X-ray flux integrated excluding the disc blackbody contribution, i.e., $F_{\rm hot}$ calculated in Section \ref{xray}. Fig.\ \ref{f:bol_radio15}(c) shows the resulting relationship. In contrast to Fig.\ \ref{f:bol_radio15}(a--b), we see that now there is an approximate single power-law dependence encompassing {\it all\/} the states of Cyg X-1, with $p=1.73\pm 0.07$.
The GBI data provide information at frequencies below 15 GHz and were simultaneous with part of the ASM and BATSE data. We find correlations at 2.25 and 8.3 GHz in the hard state similar to that found at 15 GHz -- see Fig.\ \ref{f:radio28} and the first and last sections of Table \ref{t:fit}. Despite the relatively low quality of those data, the correlations are highly significant, as seen from the low values of $P_{\rm s}$. The fitted indices are, however, subject to some systematic errors associated with the choice of the minimum radio flux, $F_{\rm R,min}$, considered. Using $F_{\rm R,min}= 15$ mJy at 2.25 GHz and 8.3 GHz would lead to somewhat lower values of $p$ than with the limit of 10 mJy adopted here. In any case, $p$ decreases with the decreasing frequency.
Fig.\ \ref{f:sprx} shows the broad-band spectra of Cyg X-1 from radio to X-rays. The black circles and vertical error bars show the average hard-state radio fluxes from \citet{fender00}. The dotted line shows an $\alpha=0$ ($F_E\propto E^{-\alpha}$, $\alpha=\Gamma-1$) dependence, which approximately fits the hard-state radio data. We show the line extending to 0.1 eV, which approximately correspond to the turnover frequency (at which the jet becomes optically thin to synchrotron self-absorption) of $2.9\times 10^{13}$ Hz found by \citet{rahoui11}. The red, green and blue crosses show the (geometric) average fluxes in the hard, intermediate and soft state, respectively, in the radio and X-ray ranges. Only the ASM and BAT data simultaneous with the 15 GHz monitoring were included. The 2.25 GHz and 8.3 GHz data are the averages over all the data for which there were simultaneous ASM data for determining the spectral state. There are no blue crosses at those frequencies because there was virtually no soft state during the GBI monitoring. The cyan crosses and error bars show the measurements of \citet{persi80} and \citet{mirabel96} of infrared emission, which is from the companion and its wind. The dashed lines are extrapolations of the 20--35 keV average spectra (which emission contains virtually no contribution from the disc blackbody) down to 0.1 eV.
\begin{figure*}
\centerline{\includegraphics[height=8cm]{bat_radio15.eps}
\includegraphics[height=8cm]{batse_radio15.eps}
\includegraphics[height=8cm]{hot_bat_radio15.eps}}
\caption{Relationship between the daily-averaged radio flux at 15 GHz and the corresponding (a) $F_{\rm bol}$ from the ASM+BAT data, (b) $F_{\rm bol}$ from the ASM+BATSE data, and (c) the Comptonization flux, $F_{\rm hot}$ from the ASM+BAT data. $F_{\rm hot}$ is approximated as $F(>3\,{\rm keV})$ in the soft state, and $F(>1.5\,{\rm keV})$ in the intermediate and hard states. The red triangles, green squares and blue circles correspond to $\Gamma(3$--12 keV) of $\Gamma<1.9$, $1.9<\Gamma<2.3$ and $\Gamma>2.3$, respectively. The error bars on the radio measurements are not shown for clarity. The solid lines show the fits to the hard-state data for (a--b) and to the data in all states for (c) using equation (\ref{fit}), see Table \ref{t:fit}.
}
\label{f:bol_radio15}
\end{figure*}
\section{Discussion}
\label{discussion}
\subsection{The nature of the X-ray variability}
\label{xvar}
We find that the hard state, with $\Gamma(3$--$12\,{\rm keV})<1.9$, has the dominant variability pattern of the $\sim$14--150 keV spectrum with a constant shape moving up and down, see Figs.\ \ref{f:X_radio}(c--e), \ref{f:gamma}(a). $F_{\rm bol}$ changes by a factor of $\simeq$5 within the hard state. This pattern can be caused by a changing accretion rate, $\dot M$, in a stable geometry (Z02).
The superorbital variability has been found by ZPS11 to show the same variability pattern. However, its amplitude is only by about $\pm 0.2$ above 15 keV (ZPS11), and thus it can explain only a small part of the observed factor of $\sim$5 variability.
On the other hand, the softest ($<$3 keV) and hardest ($>$150 keV) X-rays have faster, roughly quadratic, relationships with $F_{\rm bol}$. Below 3 keV, absorption is strong and there is a soft-excess component (e.g., \citealt{gierlinski97}). The observed relationship is likely to be connected to the absorbing column increasing at decreasing fluxes (whereas $F_{\rm X}$ have been calculated under the assumption of a constant absorbing column, see Appendix \ref{bol}). Also, it is possible that the soft excess is weaker at lower fluxes. We cannot evaluate the relative contributions of these effects based on our data alone.
We note here that ZPS11 argued for precession of the accretion disc as the preferred model for the superorbital variability based on the constancy of the photon index above 20 keV during the superorbital cycle. In the light of the finding of this work, precession is still possible but not proven. Our results in Figs.\ \ref{f:X_radio}(a--e) show $F_{\rm bol}\propto F_{\rm X}$ in the 14--150 keV range, which is compatible with the spectrum with the same photon index moving up and down. In the present case, the amplitude of this variability is by a factor of $\simeq$5, and it has to be driven by changing $\dot M$. Also, Figs.\ \ref{f:X_radio}(a--b) are compatible with increasing absorption at low fluxes, similar to the behaviour found during the superorbital cycle (ZPS11). Thus, it is possible that the superorbital variability of Cyg X-1 is driven by a modulation of $\dot M$ on the time scales of about 150 or 300 d, which cause, however, would be unknown.
Above 150 keV, the hard-state spectra show a cut-off, which in the framework of the thermal Comptonization model is expected at $E\mathrel{\copy\simgreatbox} (2$--$3) kT_{\rm e}$, where $T_{\rm e}$ is the temperature of the Comptonizing electrons. The low value of the correlation index, $p'$, at these energies indicates that $kT_{\rm e}$ increases with increasing $F_{\rm bol}$. This trend is opposite to that modeled by Z02, where the increased plasma cooling at a higher luminosity leads to a small decrease of $kT_{\rm e}$ (see fig.\ 14c of Z02). We note, however, that given the strong noise in the 150--195 keV data, the exact variability pattern in this band is relatively uncertain. Also, the trend of fig.\ 14(c) Z02 may be overcome by the increase of $kT_{\rm e}$ for harder spectra, see figs.\ 14(a--b) in Z02, and see fig.\ 13 in Z02 for example spectra from pointed observations showing this trend.
For the intermediate state, up to $\Gamma\simeq 2.3$--2.5, the bolometric flux increases with the increasing $\Gamma$, see Fig.\ \ref{f:gamma}(a). The broad-band variability shows also a pivot at $\sim$20 keV, above which the intermediate-state X-ray fluxes in Figs.\ \ref{f:X_radio}(a--e) change from being correlated with $F_{\rm bol}$ to an anti-correlation. This has been modelled by Z02 and \citet*{gzd11} as due to a changing ratio of the flux irradiating a Comptonizing plasma to that dissipated in that plasma (see figs.\ 14a--b in Z02, 10c in \citealt{gzd11}). This may be caused by a varying the inner radius of an accretion disc surrounding the hot plasma \citep*{dgk07}.
\begin{figure}
\centerline{\includegraphics[height=7cm]{batse_radio2.eps}
\includegraphics[height=7cm]{batse_radio8.eps}}
\caption{Correlations of the daily-averaged radio flux at (a) 2.25 GHz and (b) 8.3 GHz with the corresponding bolometric flux estimated based on the ASM and BATSE data. The red triangles, green squares and blue circles correspond $\Gamma(3$--12 keV) of $<1.9$, $1.9<\Gamma<2.3$ and $>2.3$, respectively. The error bars on the radio measurements are not shown for clarity. The solid lines show the best fits of the power-law correlation to the hard-state data, see Table \ref{t:fit}. Only the data above 10 mJy have been fitted.}
\label{f:radio28}
\end{figure}
The soft-state X-ray bolometric flux is dominated by photons below $E\sim$3 keV. Below this energy the soft-state spectra are dominated by the disc component and above it, by a high-energy tail. The soft state variability has been modelled by \citet*{cgr01}, and later by Z02 and \citet{gzd11} as a variable non-thermal corona above a stable disc, see, e.g., fig.\ 16 in Z02 and fig.\ 10c in \citet{gzd11}. However, the disc blackbody emission is stable only on short time scales, whereas it does change significantly on long time scales, see, e.g., Fig.\ \ref{f:X_radio}(a). The resulting variability pattern is of the overall spectrum changing its normalization but with much more scatter if $F_{\rm X}$ vs.\ $F_{\rm bol}$ at high energies, dominated by the emission of the fast-varying corona, than in the blackbody component.
\begin{figure*}
\centerline{\includegraphics[height=9cm]{radio_X_spectra.eps}}
\caption{The broad-band spectra of Cyg X-1 from radio to X-rays. The black circles and vertical error bars show the average hard-state radio fluxes from \citet{fender00}, and the dotted line shows the $F_\nu\propto \nu^0$ dependence fitting those points extended up to the turnover energy of 0.1 eV. The red, green and blue crosses show our results in the hard, intermediate and soft state, respectively. The X-ray data are absorption-corrected. The cyan circles and error bars show the infrared fluxes, which are mostly from emission of the companion and its wind. The dashed lines are extrapolations of the 20--35 keV average spectra down to 0.1 eV.
}
\label{f:sprx}
\end{figure*}
Figs.\ \ref{f:gamma}(a), \ref{f:lc_bol} show that the daily-averaged $F_{\rm bol}$ changes, over the three states, by a factor of $\simeq$10. As expected, the highest and lowest fluxes correspond to the soft and hard state, respectively, consistent with the view that state transitions in Cyg X-1 are driven by changes of the accretion rate. The soft-state $F_{\rm bol}$ is generally about a factor of $\simeq$3 higher that in the hard state. We note, however, that $F_{\rm bol}$ in the intermediate and soft states can become less than that in the hard state, see Figs.\ \ref{f:gamma}(a), \ref{f:lc_bol}, \ref{f:X_radio}(a--e). This appears to be the first such finding in Cyg X-1. It may be indicative of hysteresis, usually present in transient sources (e.g., \citealt{dunn08}). Its presence there is seen as a characteristic q track in the flux/hardness diagram, see, e.g., fig.\ 1 in \citet{russell10}. An analogue to it is visible in Fig.\ \ref{f:gamma}(a), where some points belonging to the soft and intermediate states have $F_{\rm bol}$ below that of the hard state. However, the tracks covered are much more chaotic than those in transients, where the hard-to-soft transitions occur at clearly higher flux than the reverse ones. Although we consider this result for Cyg X-1 secure, it relies on the absorption-corrected ASM observations, and should be tested by pointed X-ray observations.
\subsection{The nature of the radio/X-ray correlation}
\label{rx}
Our findings concerning radio/X-ray correlations fall into two categories. Those regarding the dependence of the correlation on the X-ray energy range and on the spectral state are basically independent of the uncertain strength of free-free absorption in the stellar wind from the donor of Cyg X-1. On the other hand, the values of the radio/X-ray correlation indices are, most likely, affected by the free-free absorption.
\subsubsection{The correlation dependencies on the X-ray energy range and on the spectral state}
\label{range}
One of our findings is the strong dependence of the radio/X-ray correlation index on the energy band, see Table \ref{t:fit} and Fig.\ \ \ref{f:index}. The energy-dependence of the correlation can be understood as an intrinsic dependence of the radio flux on the bolometric flux combined with flux-dependent changes in the spectral shape (seen in our data at $E<12$ keV and $E>150$ keV). The latter leads to the bolometric flux and that in a narrow X-ray band being non-proportional in a way that depends on the band chosen.
This dependence has mostly been ignored in previous work on this subject except, e.g., \citet{mhd03}, where its effect has been accounted for by a theoretical model, \citet{z04}, who fitted the radio correlation with $F_{\rm bol}$ in GX 339--4, obtaining $p_{\rm bol}=0.79\pm 0.07$, and \citet{corbel03}, who calculated the correlation in GX 339--4 in four separate bands between 3 keV and 200 keV, but have not calculated the correlation index for $F_{\rm bol}$.
In models in which the accretion rate, $\dot M$, governs both the accretion flow and the jet (see Section \ref{accretion} below), it is crucial to determine its value. It is related to the bolometric flux by,
\begin{equation}
\dot M= {4\upi d^2 F_{\rm bol}\over \epsilon c^2},
\label{mdot}
\end{equation}
where $d$ is the source distance, and $\epsilon$ is the accretion radiative efficiency. Our results indicate that the jet power is a fraction not of the total accretion power but only of that responsible for emission associated with hot electrons in the accretion flow, $F_{\rm hot}$. In this case we should write $\dot M \propto F_{\rm hot}/\epsilon_{\rm hot}$, where $\epsilon_{\rm hot}$ is the efficiency of producing the hot flow. However, the difference between the two is negligible in the hard state, where the hot flow is energetically dominant.
The common procedure has been to use a narrow band, such as 3--9 keV (e.g., \citealt{coriat11}), as the proxy for $F_{\rm bol}$, i.e. implicitly assuming $F_{\rm X}\propto F_{\rm bol}$, and then proceed to theoretical interpretation. However, we have shown that this can lead to significant errors. For our data, we can estimate the 3--9 keV flux as the sum of the 3--5 keV flux and a fraction of the 5--12 keV flux, the fraction being determined assuming a power law with the 3--12 keV photon index. Using the flux in this band, we find $p=1.17\pm 0.10$, $1.23\pm 0.10$, and $1.48\pm 0.05$, for 2.25 GHz, 8.3 GHz, and 15 GHz, respectively. Comparing with Table \ref{t:fit}, we see that the 3--9 keV values of $p$ are smaller than those for $p_{\rm bol}$ by $\simeq 0.2$--0.3. This problem occurs for any system with spectral variability, which is generally the case in either black-hole or neutron-star X-ray binaries.
Our next main finding is that we are able to extend the correlation to the intermediate and soft states. The radio emission has been known to correlate in the soft states of some objects with the level of the high-energy tail, e.g., in Cyg X-3 \citep{szm08}; however, that dependence was different from that in the hard state. Here, we find that if we consider, instead of the bolometric flux, that in the emission in the hot plasma only (i.e., $F_{\rm bol}$ minus the disc blackbody component), we obtain a single radio/X-ray power law correlation going through all the states. Its index, $1.73\pm 0.07$, is consistent within uncertainties with the correlation index for $F_{\rm bol}$ (for the ASM+BAT data) in the hard state only, $1.68\pm 0.11$. This appears to indicate that the jet is launched from the hot electrons responsible for the observed X-ray emission rather than by the optically-thick disc, and that the launching mechanism may be similar in the hard spectral state and in the soft one. We stress that Cyg X-1 appears to always have some level of the high energy tail in the soft state, and consequently has often relatively strong radio emission in that state, as also found by \citet{rushton11}. This is unlike low-mass black-hole systems, which often go to an extended soft state in which the tail virtually disappears (e.g., \citealt{gd04}), and where consequently there is virtually no radio emission. Also, Cyg X-1 does not show the very high state, where ejection of radio-emitting blobs occurs instead of a jet, and where the launching mechanism is likely to be different. No such blob ejections have indeed been found during the 2010 state transition by \citet{rushton11}.
We have also found that the maximum of the radio emission occurs for relatively soft X-ray spectra, with $\Gamma(3$--$12\,{\rm keV})\simeq 2$--2.1. Thus, the transition to the soft state is first associated with an increase of the radio emission and only then with a decrease.
\subsubsection{Free-free absorption and the intrinsic correlation index}
\label{free-free}
\begin{figure*}
\centerline{\includegraphics[width=11cm]{correlation15.eps}}
\centerline{\hbox to 0.5cm{\hfill} \includegraphics[width=5.5cm]{correlation8.eps} \includegraphics[width=5.5cm]{correlation2.eps}}
\caption{Comparison of the radio/X-ray correlation for Cyg X-1 in the hard state (assuming $d=2$ kpc; black crosses) with those for GX 339--4 (\citealt{corbel03}; green crosses), V404 Cyg (\citealt{corbel08}; blue squares) and H1743--322 (\citet{coriat11}; red circles). The dotted and dashed lines have $p=0.6$ and $p=1.4$, respectively, approximately fitting the two apparent branches of the correlation for the three sources. The Cyg X-1 radio data are for (a) 15 GHz, (b) 8.3 GHz, and (c) 2.25 GHz.
}
\label{f:correlation}
\end{figure*}
Radio emission of Cyg X-1 from 2.25 GHz to 15 GHz is known to exhibit orbital modulation due to free-free absorption in the stellar wind of the companion \citep{pfb99}. The full modulation depth changes from $\sim$4 per cent at 2.25 GHz to 30 per cent at 15 GHz \citep{l06}. Thus, free-free absorption certainly affects the observed radio fluxes. However, it is not trivial to determine the degree of the attenuation, $\exp(-\bar{\tau}_{\rm ff})$, where $\bar{\tau}_{\rm ff}$ is the orbit-averaged optical depth. Using an estimate of the mass loss rate from the companion towards the polar regions (where the jet radio emission is absorbed) and of the inclination, and calculating the temperature of the wind irradiated both by the star and the X-ray source, we can use the strength of the observed orbital modulation to calculate the height along the jet, $z$, from which the bulk of the radio emission at a given frequency is received \citep{sz07}. This height corresponds to the place above which the jet becomes optically thin to synchrotron self-absorption \citep{bk79}. Based on the value of $z$, we can calculate the optical depth, $\bar{\tau}_{\rm ff}$, through the wind, provided its density and temperature profile are known.
Given that a radio flux, $F_{\rm R}=F_{\rm R,intr} \exp(-\bar{\tau}_{\rm ff})$ and $p={\rm d}\ln F_{\rm R}/{\rm d}\ln F_{\rm X}$, where $F_{\rm X}$ is now either the flux in an X-ray band, the bolometric flux or the flux emitted by hot electrons, an index, $p$, derived from the observations is related to the corresponding intrinsic index, $p_{\rm intr}={\rm d}\ln F_{\rm R,intr}/{\rm d}\ln F_{\rm X}$, by
\begin{equation}
p_{\rm intr}= r p,\quad r= {{\rm d} \ln F_{\rm R,intr}\over {\rm d}\ln F_{\rm R}}
=1+{{\rm d}\bar{\tau}_{\rm ff} \over {\rm d}\ln F_{\rm R}}.
\label{index}
\end{equation}
Generally, we expect the jet power to positively correlate with the height along the jet where the emission becomes optically thin to synchrotron self-absorption. Thus $F_{\rm R}$ should be anti-correlated with the depth within the stellar wind, given by $\bar\tau_{\rm ff}$, which then implies $r<1$. Consequently, the effect of free-free absorption of the radio fluxes is to increase their observed variability range with respect to the intrinsic range, and to make the radio/X-ray correlation steeper. This effect has been suggested to qualitatively explain the relative steepness of the correlation in Cyg X-1 by \citet{gfp03}.
A quantitative investigation of this effect has been made by \citet{zdz11}, where it was found that the depth of the orbital radio modulation at 15 GHz (related to $\bar\tau_{\rm ff}$) indeed significantly anti-correlates with $F_{\rm R}$. This can yield the factor $r$ of equation (\ref{index}), provided the wind parameters are specified. Here, an important factor is the weakness of the polar component of the wind of Cyg X-1 recently found by \citet{gies08}. Consequently, \citet{zdz11} found values of $\bar{\tau}_{\rm ff}$ significantly lower than those obtained by \citet{sz07}, who assumed the wind is isotropic. The best resulting estimate of $r$ is $\sim$0.8 at 15 GHz, which yields $p_{\rm bol,intr}\simeq 1.3\pm 0.1$ for the correlation with the bolometric flux estimated based on the ASM+BAT data. That estimate also implies that the wind attenuation of the 15 GHz flux is rather small, which is consistent with the observed 2--220 GHz spectrum (shown also in Fig.\ \ref{f:sprx}) being relatively straight with $\alpha\simeq 0$, and not showing deviations from that power law by more a factor of $\sim$2. Also, the relative mildness of the wind attenuation is consistent with the orbital modulation at 89 GHz, 146 GHz and 221 GHz not being reported by \citet{fender00}. We caution, however, that the theoretical model of \citet{zdz11} is a subject to uncertainties due to the uncertain wind density in polar regions.
In order to show the Cyg X-1 hard-state correlation in the context of other sources, we have compared it to those of the black-hole binaries H1743--322, GX 339--4 and V404 Cyg, see fig.\ 5 in \citet{coriat11}. We have calculated the 3--9 keV fluxes as described in Section \ref{range} above. To obtain the luminosities, we have assumed isotropy and a distance to Cyg X-1 of $d=2$ kpc, a value is in the middle of the range of about 1.8--2.2 kpc of \citet{zi05}, and consistent with the 1.5--2 kpc range found by \citet{cn09}. The results are shown in Fig.\ \ref{f:correlation}. (Changing $d$ would move the Cyg X-1 points diagonally.)
We see that the Cyg X-1 dependence at 15 GHz is only slightly steeper than $p\sim 1.4$ found for H1743--322. The Cyg X-1 correlation slope for 3--9 keV is $p=1.48\pm 0.05$ (Section \ref{range}). Given the above, relatively minor, correction to the correlation slope for free-free absorption of $r\sim 0.8$, we deduce $p_{\rm intr}\sim 1.2$ for 3--9 keV. The 2.25 GHz and 8.3 GHz data are more uncertain but suggest a similar slope. Thus, Cyg X-1 appears to be intrinsically similar to H1743--322, rather than to the standard binaries showing the radio/X-ray correlation, GX 339--4 and V404 Cyg, with $p\sim 0.6$. We also note that the normalizations of the 2.25 GHz, 8.3 GHz and 15 GHz flux distributions are very similar, which is not compatible with the 15 GHz data being strongly affected by attenuation and again suggests that the correction due to free-free absorption is relatively low.
\subsubsection{Accretion flow models and radio emission}
\label{accretion}
\citet{coriat11} gives an extensive discussion of theoretical models that could explain $p\simeq 1.4$, and we refer the reader there for details. Assuming a scale-free jet, \citet{hs03} have obtained the following scaling between the jet power, $Q_{\rm j}$, and the flux at a given radio frequency (below the turnover $\nu$),
\begin{equation}
F_{\rm R} \propto Q_{\rm j}^\xi, \qquad \xi={2p_{\rm e} +(p_{\rm e}+6)\alpha+13\over 2(p_{\rm e}+4)}.
\label{fr_Q}
\end{equation}
Here, $p_{\rm e}$ is the index of the power-law relativistic electron distribution, $n_{\rm e}(\gamma)\propto \gamma^{-p_{\rm e}}$, and $\gamma$ is the Lorentz factor, with typical values of $p_{\rm e}\simeq 2$--3 (from observations of optically-thin parts of synchrotron spectra). For Cyg X-1 the radio photon index is $\alpha\simeq 0$, in which case $\xi\simeq 1.4$, with only a weak dependence on $p_{\rm e}$. We assume the jet power is a constant fraction, $f_{\rm j}<1$, of the accretion power, i.e., $Q_{\rm j}=f_{\rm j}\dot M c^2$. Then, $\dot M\propto F_{\rm bol}/\epsilon$, equation (\ref{mdot}). Advection models predict $\epsilon$ decreasing with the decreasing accretion rate, $\epsilon\propto \dot M^{q-1}$, where $q\sim 2$ is the index relating $F_{\rm bol}$ to $\dot M$. On the other hand, $\epsilon$ is a constant, $\sim 0.1$, in fully efficient accretion flows, and thus $q=1$. This scaling connects the radio flux to the bolometric one,
\begin{equation}
F_{\rm R}\propto F_{\rm bol}^{\xi/q}.
\label{fr_fbol}
\end{equation}
Thus, $p_{\rm intr}=\xi/q$, implying $p_{\rm intr}\simeq 1.4$ for sources with radiatively efficient accretion flows, and $p_{\rm intr}\simeq 0.7$ for sources with advective accretion flows.
Our results of Section \ref{free-free} indicate that $p_{\rm intr}\sim 1.3$ or so through all the states of Cyg X-1 (although with with caveats due to the effect of wind absorption being not fully known), and thus its accretion flow appears to be efficient. In the hard state, this conclusion supports the same finding of \citet*{mbf09} but based on the difference in the efficiency between the average hard state and the soft state in Cyg X-1 being small. In this state, a likely model is a luminous hot accretion flow \citep{yuan01}.
In the soft state, the hot electrons most likely form a corona above a disc, which corona is radiatively efficient, being cooled by the disc blackbody photons. Then, $q=1$ and we expect the same correlation index in the soft state for the hot electron emission ($F_{\rm hot}$) as in the hard state provided the power supplied to the corona is a constant fraction of the accretion power. Fig.\ \ref{f:hot_bol} shows that this is indeed the case, with at most a weak dependence, and the fraction equals $\sim$1/4. Both the constancy and the fractional value are consistent with theoretical expectations for the gas-pressure dominated disc \citep{merloni03,mhd03}. This is also in agreement with the finding that the disc in Cyg X-1 is in the gas-pressure dominance regime \citep{gierlinski99}. Then, the fact that the radio/X-ray correlation extends, with about the same $p$, from the hard state to the soft one (for $F_{\rm hot}$), represents then another argument for the high accretion efficiency in the hard state Cyg X-1.
\subsubsection{X-ray jet models}
\label{xray_jet}
An alternative to the accretion flow model as an explanation of the radio/X-ray correlation is that both radio and X-rays are emitted by the jet. In particular, both could be emitted by the synchrotron emission of the same population of non-thermal electrons, with the radio and X-rays from the optically thick and thin parts, respectively. For this case, \citet{hs03} have derived,
\begin{equation}
F_{\rm R} \propto F_{\rm X}^\zeta, \qquad \zeta={2+\alpha +(3+\alpha)/(p_{\rm e}+5) \over 2+p_{\rm e}/2},
\label{fr_fx}
\end{equation}
which yields $\zeta\simeq 0.8$--0.7 for $p_{\rm e}=2$--3. Thus, this model would have required $p_{\rm intr}=\zeta\simeq 0.7$--0.8, and a large correction to $p$ due to free-free absorption, which is in conflict with our results in Section \ref{free-free}.
Moreover, Fig.\ \ref{f:sprx} shows that the normalization of the X-ray emission in the hard state is too high by a factor of $\sim$30 for the X-rays to be due to optically thin non-thermal synchrotron emission above the turnover frequency, which is $\simeq$0.1 eV in Cyg X-1 \citep{rahoui11}. The situation becomes even worse for the intermediate state (in which the radio emission peaks), where the disagreement in the normalization at 0.1 eV is by three orders of magnitude. As discussed in Section \ref{free-free}, the effect of free-free absorption on the average 2--220 GHz spectrum can be at most a factor of $\sim$2, which cannot explain this discrepancy. This rules out non-thermal synchrotron as the origin of the X-rays. This has also been independently found by \citet{rahoui11}. We note that another ground for ruling out this model for luminous hard state of black-hole binaries is that it does not reproduce the shape of the high-energy cut-off, and it has problems reproducing the narrow distribution of the observed cut-off energies \citep{z03}.
We note that \citet{russell10} found the X-ray spectra of the black-hole binary XTE J1550--564 to be dominated by thermal Comptonization in its luminous hard states, at $L\mathrel{\copy\simgreatbox} 2\times 10^{-3} L_{\rm E}$. Cyg X-1 in the hard state is relatively luminous, typically with $L\mathrel{\copy\simgreatbox} 0.01 L_{\rm E}$ (Z02), where $L_{\rm E}$ is the Eddington luminosity. Thus, our findings above are consistent with those of \citet{russell10}. On the other hand, the spectra of XTE J1550--564 were found to be dominated by jet non-thermal synchrotron emission at lower luminosities. A weak jet contribution, not detectable in our X-ray data, remains also possible in Cyg X-1, given the relative normalization of the radio power law and that of the X-rays, see Fig.\ \ref{f:sprx}.
X-ray jet models for luminous states of black-hole binaries, and for Cyg X-1 in particular, have been also ruled out on several other grounds by \citet{mbf09}. In particular, they discuss the model of \citet*{mnw05}, in which the X-rays in the hard state of Cyg X-1 (and other black-hole binaries) are due to synchrotron-self-Compton by thermal electrons with $kT_{\rm e}\sim 3$--5 MeV of a very low Thomson optical depth at the jet base. \citet{mbf09} find that the parameters fitted by \citet{mnw05} strongly violate the e$^\pm$ pair equilibrium, with the self-consistent optical depth being two orders of magnitude higher than the fitted one due to the produced pairs. Furthermore, that model relies on first-order Compton scattering to fit the X-rays, and thus it requires strong fine-tuning (as noted by \citealt{y07}) to reproduce the observed spectral cut-off, which is at $\sim$100 keV in both Cyg X-1 and in most of the black-hole binaries in the hard state. Other problems with the X-ray jet models are discussed by \citet{heinz04} (effect of electron energy loss on optically-thin synchrotron spectra) and by \citet{maccarone05} (comparison of black-hole and neutron-star sources).
A number of papers have also attributed the high-energy tail observed in the soft state of X-ray binaries to optically thin non-thermal synchrotron or self-Compton emission from the jet, e.g., \citet{vadawale01}, \citet{fiocchi06}. We note that for the soft state in Cyg X-1, the extrapolation of the high-energy tail down to 1 eV is a few orders of magnitude above any possible extrapolation of the radio emission in that state. This rules out this model. In addition, it requires that the actual bolometric luminosity of the modelled sources is a few orders of magnitude above the ones determined based on the X-rays, as well as that the emission is unbeamed, which appear highly unlikely. Instead, these high-energy tails appear compatible with Compton scattering of disc blackbody photons by a hybrid electron distribution, e.g., \citet{gierlinski99}.
\section{Summary and conclusions}
\label{conclusions}
\subsection{X-rays}
We have obtained broad-band X-ray spectra of Cyg X-1 as a function of time using the broad-band monitoring by {\textit{RXTE}}/ASM together with either {\textit{CGRO}}/BATSE or {\textit{Swift}}/BAT. Based on these spectra, we have calculated both the bolometric fluxes ($F_{\rm bol}$), and approximate values of the Comptonization flux ($F_{\rm hot}$; not including the blackbody component). We have classified the spectra into three spectral states based on their 3--12 keV photon index. We present light curves of $F_{\rm bol}$ for the available periods. The range of variability of the bolometric flux using 1-day averages is by a factor of $\simeq 10$. We find the fluxes in different states overlap, e.g., some fluxes in the hard state are higher than some in the soft state. This indicates the presence of some X-ray hysteresis in Cyg X-1, though weaker than that in low-mass X-ray binaries.
We have studied X-ray variability patterns of Cyg X-1. In the hard state, the dominant pattern in the $\sim$10--150 keV range is of the intrinsic spectrum changing its normalization only, but with apparently more absorption at soft X-rays, causing their flux to respond to changes of the bolometric luminosity more strongly than the hard X-ray flux. In the intermediate state, there is strong spectral variability with the overall spectra changing their slope with a pivot around 20 keV. In the soft state, there is an approximate proportionality of the flux in a given energy band to $F_{\rm bol}$, but with the scatter strongly increasing with the photon energy. Still, $F_{\rm hot}/F_{\rm bol}$ was found approximately constant on average in this state, $\simeq 1/4$, see Fig.\ \ref{f:hot_bol}.
\subsection{Radio vs.\ X-rays}
We have shown that the character of the 15 GHz radio/X-ray correlation in Cyg X-1 strongly depends on the chosen X-ray band. Results based on soft X-rays significantly differ from those based on hard X-rays. The correlation indices in the hard state vary from $p\simeq 0.9$ to $\simeq 1.8$ across the X-ray energy band. This dependence can be understood as a combination of the relationship between radio flux and bolometric flux with that between the latter and the narrow-band X-ray flux. Instead of using such narrow X-ray bands, the bolometric and Comptonization fluxes provide a much better direct measure of the underlying physics. In the hard state, the correlation index for either of them is $p\simeq 1.7$. The values of $p$ obtained based on the broad-band fluxes are larger by $\simeq$0.2--0.3 than those obtained using the commonly used X-ray band of 3--9 keV.
We have found the radio/X-ray correlation is present in all spectral states of Cyg X-1. The radio flux peaks for relatively soft spectra, with $\Gamma\simeq 2$. For softer spectra, the radio flux drops rapidly if compared to soft X-rays, but it retains a power law dependence with hard X-rays. In particular, the radio correlation with the Comptonization flux forms a single dependence, with $p\simeq 1.7$, across all of the spectral states of Cyg X-1. This indicates that the radio jet is formed by the hot Comptonizing electrons in the accretion flow, and not by the blackbody disc.
The radio flux in Cyg X-1 is attenuated by free-free absorption in the stellar wind from the companion, which increases the value of the observed correlation index, $p$, above the intrinsic value, $p_{\rm intr}=r p$, $r<1$. The results of \citet{zdz11} supported by our results at 2.25 GHz, 8.3 GHz and 15 GHz and their comparison with the correlation for low-mass X-ray binaries containing black holes suggest $p_{\rm intr}\sim 1.3$ or so (for the bolometric luminosity). This is similar to $p_{\rm bol}\simeq 1.4$ characteristic of efficient hot accretion flow in the framework of accretion X-ray models, implying the presence of such a flow in Cyg X-1. Based on the relative normalization of the radio and X-ray fluxes in Cyg X-1, we rule out X-ray jet models in which the X-rays are due to an extrapolation of the radio spectrum.
\section*{ACKNOWLEDGMENTS}
We thank T. Belloni, S. Corbel, R. Narayan, M. Ostrowski, A. R. Rao and F. Yuan for valuable discussions. We also thank M. Coriat for providing us with the data used in Fig.\ \ref{f:correlation}, and the referee for valuable suggestions. This research has been supported in part by the Polish MNiSW grants N N203 581240, N N203 404939 and 362/1/N-INTEGRAL/2008/09/0. The AMI Arrays are operated by the University of Cambridge and supported by the STFC. The Green Bank Interferometer is a facility of the National Science Foundation operated by the NRAO in support of NASA High Energy Astrophysics programs. We acknowledge the use of data provided by the {\textit{RXTE}}\/ ASM and {\textit{Swift}}\/ BAT teams.
|
\section{Introduction}
The most massive early-type galaxies reside within dark matter halos with masses of $10^{13}~{\rm M_\odot}$ or
more \citep[e.g.,][]{bro08}. One may expect the gas within these dark matter halos to radiatively cool,
gravitationally collapse and form stars \citep[e.g.,][]{whi78}. In contrast to this expectation,
stellar population synthesis modeling reveals that the bulk of the stars in very massive galaxies
formed $\simeq 10~{\rm Gyr}$ ago, with relatively little star formation since \citep[e.g.,][]{tin68,tra00}.
While some early-type galaxies harbor cold gas, the most massive galaxies are typically bathed in
plasma with temperatures on the order of millions of Kelvin \citep[e.g.,][]{can87}. What is heating the plasma
and regulating star formation in very massive galaxies has yet to be robustly identified.
In the recent literature, a popular mechanism for heating the plasma is the injection of energy from
an active galactic nucleus \citep[AGN feedback; e.g.,][]{tab93,cro06}. While a quasar could heat the
gas within its host galaxy \citep[e.g.,][]{hop06}, powerful quasars are so rare that they
probably cannot be responsible for the lack of star formation in nearby massive galaxies. For nearby galaxies,
it is more plausible that low luminosity AGNs (with a high duty cycle) are heating the plasma.
Bubbles in the distribution of X-ray emitting gas surrounding nearby radio galaxies correspond to the locations of
radio lobes \citep[e.g.,][]{fab03a,mcn00}. The dissipation of shocks and sound waves resulting from the
production of these bubbles and jets may inject sufficient energy into the plasma to offset radiative
cooling \citep{fab03a}.
Radio observations provide insights into the plausibility and nature of AGN feedback.
We may expect to observe radio emission from all massive galaxies, resulting from star formation
and/or AGNs. Radio emission is associated with the cavities in the X-ray emitting plasma produced by
AGNs and radio emission results from recent star formation \citep{con92}. However, in isolation,
the presence of radio emission from massive galaxies is not proof of AGN regulation of star formation.
For example, AGNs with low power jets \citep[e.g.,][]{bal09} may have little impact on the surrounding plasma.
If there are massive galaxies without radio emission, this may indicate that brief bursts of AGN activity are sufficient
to truncate star formation \citep[e.g.,][]{hop06}. Alternatively, a mechanism other than AGN feedback may be
heating the plasma surrounding galaxies \citep[e.g.,][]{bir03}.
Previous studies show that $\sim 30\%$ of the most massive galaxies are radio continuum
sources \citep[e.g.,][]{fab89,sad89,wro91,bes05,sha08}. These studies matched optical and radio source catalogs,
which limits the sample to radio sources that meet conservative signal-to-noise criteria, so catalogs
are not swamped by noise. (Even when conservative criteria are applied, the tail of the noise distribution
may produce spurious sources.) Recent studies have generally utilized large redshift
surveys that exclude the nearest galaxies, and thus miss radio sources fainter than $10^{22}~{\rm W~{Hz}^{-1}}$.
Consequently, the faint radio emission from nearby early-type galaxies has not been completely characterized
by the prior literature.
In this letter we present a study of the 1.4~GHz radio emission from $K<9$ early-type galaxies. The choice of 1.4~GHz is
pragmatic, as it allows us to utilize existing NRAO VLA Sky Survey \citep[NVSS;][]{con98} and single-dish imagery.
We assume the radio emission is the consequence of either recent star formation or an AGN. If this assumption holds,
our conclusions do not depend on which emission process is dominant in these galaxies (i.e., synchrotron, free-free).
Rather than match our early-type galaxies to radio source catalogs alone, we also measure flux densities from radio images.
We can thus include significant (albeit noisy) information on the
radio flux densities of early-type galaxies that would have otherwise been excluded from our study. This allows us
to characterize the very faint radio emission from the most massive early-type galaxies.
\section{The Sample}
Our parent sample is the 2MASS Extended Source Catalog \citep{jar00}, from which we select
objects with apparent magnitude $K<9$ \citep[dust corrected;][]{sch98}, declination $\delta>-40^{\circ}$ and galactic latitude
of $|b|>15^{\circ}$. Of the 1107 objects selected with these criteria, 979 have morphologies
available from the Third Reference Catalog of Bright Galaxies \citep{rc3} while virtually
all of the remaining objects are Galactic.
Our principal sample is the 400 galaxies that are classified as elliptical or lenticular
galaxies in the RC3 (with T type classifications of -1 or less). Many of these galaxies
have redshifts in the RC3 catalog, while the remainder have redshifts provided by
\citet{fal99}, \citet{huc99}, \citet{weg03}, \citet{jon09} and the NASA Extragalactic Database.
For 170 galaxies we have redshift independent distances from Extragalactic Distance
Database \citep{tul09}, while for the remaining galaxies we use a Hubble
constant of $73~{\rm km~s^{-1}~Mpc^{-1}}$ \citep{spe07}. The luminosity distances
of the galaxies span from $0.7$ to $10^2~{\rm Mpc}$
Our principal radio imaging is the $1.4~{\rm GHz}$ NRAO VLA Sky Survey \citep{con98},
which has an angular resolution of $45^{\prime\prime}$ and an RMS noise of $\simeq 0.45~{\rm mJy}$.
As the NVSS can underestimate the flux densities of powerful and extended radio sources, we also measure
flux densities using lower resolution ($\sim 12^\prime$) imaging from the 300-ft Green Bank and the
64-m Parkes radio telescopes \citep[][Calabretta in prep.]{con85,con86,bar01}.
We employed the following method to measure flux densities for each galaxy. At the 2MASS position
of each galaxy we measured a flux density per beam directly from the NVSS images. For those
galaxies with a flux density per beam greater than $2~{\rm mJy}$, we searched for the nearest
counterpart in the NVSS catalog within $2^{\prime}$ and used the NVSS catalog deconvolved flux density.
To account for powerful extended sources, we measured the flux density per beam using Green Bank 300-ft and Parkes
single dish data, and utilized this flux density measurement if it was greater than $0.6~{\rm Jy}$.
If the flux density per beam measured with single dish imagery was greater than $5~{\rm Jy}$, we utilized
the integrated flux density rather than the flux density per beam.
Our flux density measurements are imperfect, but our principal conclusions remain unchanged
unless gross systematic errors are present.
To verify that our data was free of such errors, we visually inspected the radio imagery and compared
our flux densities with those from the literature. Visual inspection also revealed that few galaxies
in our sample are extended \citet{far74} class II radio sources. The flux densities of 4 galaxies
neighboring bright radio sources could not be reliably measured, and excluding these galaxies reduced
our final sample to 396 early-type galaxies. The properties of the sample
galaxies are summarized in Table~\ref{table:summary}.
\section{The Radio Flux Density and Luminosity Distributions}
The $1.4~{\rm GHz}$ flux densities of the 396 $K<9$ early-type galaxies are
plotted as a function of $K$-band absolute magnitude in Figure~\ref{fig:fluxes}.
For the lowest mass galaxies, with $M_K\sim -22.5$, only a small
fraction have counterparts in the NVSS catalogs and most galaxies have flux densities
consistent with zero plus a random measurement error.
In contrast, all 59 of the $M_K<-25.5$ early-type galaxies (corresponding to stellar masses
of $\gtrsim 2.5\times 10^{11}~M_\odot$) have measured flux densities greater than zero
and just two of the 61 $-25.0<M_K<-25.5$ early-type galaxies have measured flux densities below zero.
As random errors will result in measured flux densities scattered above and below
the true flux densities, it is highly likely that all $M_K<-25.5$ early-type galaxies in our sample are
radio continuum sources, with flux densities of $\gtrsim 1~{\rm mJy}$.
\begin{figure}[hbt]
\begin{center}
\resizebox{3.25in}{!}{\includegraphics{fluxes.ps}}
\end{center}
\caption{The flux densities of early-type galaxies as a function of $K$-band absolute magnitude.
For the lowest mass galaxies, relatively few galaxies have counterparts in the NVSS catalogs
and most galaxies have flux densities consistent with zero plus a random measurement error.
Galaxies with $M_K<-25.5$ (corresponding to stellar masses of $>2\times 10^{11}~M_\odot$)
have counterparts in the NVSS catalogs or flux densities greater than zero. As random errors
will broaden the flux density distribution, it is likely that all $M_K<-25.5$ early-type
galaxies in our sample are radio continuum sources with flux densities of $\gtrsim 1~{\rm mJy}$.}
\vspace*{0.5cm}
\label{fig:fluxes}
\end{figure}
In Figure~\ref{fig:lums} we present the radio power of early-type galaxies as a function
of $K$-band absolute magnitude. As noted by others \citep[e.g.,][]{fab89,sad89},
the radio powers of early-type galaxies are a strong function of absolute magnitude.
Although our $K=9$ magnitude limit excludes some extremely powerful radio sources (e.g., Cygnus A),
the radio power of $M_K<-24$ galaxies at fixed $K$-band magnitude (or stellar mass) still spans 4 orders of magnitude.
The most massive galaxies can have $1.4~{\rm GHz}$ powers as large as that of M~87 ($\simeq 7\times 10^{24}~{\rm W~Hz^{-1}}$)
or less than the Milky Way ($\simeq 4\times 10^{21}~{\rm W~Hz^{-1}}$).
\begin{figure}[hbt]
\begin{center}
\resizebox{3.25in}{!}{\includegraphics{lums.ps}}
\end{center}
\caption{The $1.4~{\rm GHz}$ radio power of early-type galaxies as a function of $K$-band absolute
magnitude. The data are color coded by RC3 T type and for sources with measured flux densities less
than $2\sigma$ above zero, we plot $2\sigma$ upper limits. At a fixed absolute magnitude, the distribution
of radio powers spans four orders of magnitude for early-type galaxies brighter than $M_K = -24$.
The median radio power of early-type galaxies (derived from fitting a log-normal probability density function to
the data) is proportional to $K$-band luminosity to the power of $2.78 \pm 0.16$.}
\vspace*{0.5cm}
\label{fig:lums}
\end{figure}
To model the distribution of radio powers, $P_{1.4}$, we have utilized a log-normal probability density function where
\begin{equation}
\rho (P_{1.4};\mu,\sigma) = \frac{1}{P_{1.4}\sigma\sqrt{2\pi}}{\rm exp}\left[-\frac{({\rm ln} P_{1.4} - \mu)^2}{2\sigma^2}\right]
\end{equation}
\begin{equation}
\mu= {\rm ln} \left[\alpha \left( \frac{L_K}{10^{11}L_\odot}\right)^\beta \right]
\end{equation}
\begin{equation}
\sigma = \gamma \mu
\end{equation}
where $L_K$ is the $K$-band luminosity and $\alpha$, $\beta$ and $\gamma$ are free parameters.
This parameterization is empirical (i.e., lacks physical motivation) but, as discussed below, provides
a good description of the observed distribution of radio powers as a function of $L_K$.
If the radio emission from early-type galaxies was a function of $L_K$ and time only, then
this parameterization would directly measure the radio duty cycle of early-type galaxies.
However, the radio emission from early-type galaxies is a function of environment
\citep[host halo mass and location within a halo; e.g.,][]{bes07,wak08,man09}, so this
parameterization constrains rather than measures the duty cycle.
The integral of the probability density function provides an estimate of the fraction of galaxies with
luminosity $L_K$ that have radio powers above (or below) a particular threshold, and this can be compared
with estimates from the prior literature \citep[e.g.,][]{bes05}. The radio luminosity function can be derived by
convolving the probability density function with the $K$-band luminosity function of early-type galaxies.
We used our model of the distribution of radio powers and an empirical model of the NVSS flux density
measurement errors to determine the likelihood of a particular galaxy (with known luminosity distance and $M_K$)
having a particular measured flux density. Our model of the NVSS flux density measurement errors is the distribution of
flux densities measured in each pixel of the NVSS, so we account for the non-Gaussian error
distribution and source confusion. We then used the maximum likelihood method to determine the
best-fit values and uncertainties for our model of the radio powers of early-type galaxies.
All galaxies were used when fitting the model, irrespective of their measured flux densities.
However, $M_K>-23$ galaxies provide limited constraints on the model parameters so our conclusions
principally apply to $M_K<-23$ early-type galaxies.
Our best-fit values for the log-normal distribution parameters $\alpha$, $\beta$ and $\gamma$ are
$1.16 \pm 0.20 \times 10^{20}~{\rm W~Hz^{-1}}$, $2.78 \pm 0.17$ and $6.62\pm 0.29 \times 10^{-2}$ respectively.
To verify our model, we have used it to generate 500 realizations of the anticipated observed properties of our sample,
and four of these ``mock catalogs'' are plotted in Figure~\ref{fig:mock}. 2D Kolmogorov-Smirnov (KS) tests \citep{pea83} run on
the mocks and the original sample show that the our model is consistent
with the observed distribution of radio powers. We also used the mocks to test the assumption that the
measured minimum flux density is an underestimate of the true minimum flux density for $M_K<-25.5$
early-type galaxies (due to random errors broadening the measured flux density distribution).
For 96\% of the mock catalogs, the measured minimum $1.4~{\rm GHz}$ flux density is less
than the true minimum $1.4~{\rm GHz}$ flux density for $M_K<-25.5$ early-type galaxies. It is thus
is highly likely that all $M_K<-25.5$ early-type galaxies are radio continuum sources.
\begin{figure}[hbt]
\begin{center}
\resizebox{3.25in}{!}{\includegraphics{mock.ps}}
\end{center}
\caption{The radio powers of early-type galaxies as a function of $K$-band
absolute magnitude for four mock catalogs. The distribution of radio powers approximates
what is seen in Figure~\ref{fig:lums}. Apparent features are seen in
the real and mock catalogs, but these are not significant departures from
our model of the distribution of radio powers.}
\vspace*{0.5cm}
\label{fig:mock}
\end{figure}
The median $1.4~{\rm GHz}$ radio power of early-type galaxies is proportional to $K$-band luminosity to the power of $2.78 \pm 0.17$.
For comparison, \cite{fab89} and \cite{sad89} find $5~{\rm GHz}$ radio power is proportional to $B$-band luminosity to the
power of $2.36\pm^{0.37}_{0.27}$ and $2.2\pm 0.3$ respectively. As these studies use $B$-band and $5~{\rm GHz}$ imagery, the modest
differences between these results and our own may arise from the different wavelengths used.
The fraction of early-type galaxies hosting a radio source
more powerful than $10^{23}~{\rm W~Hz^{-1}}$ is roughly proportional to $L_K^{2.2}$, increasing from
$0.2\%$ at $M_K=-23.5$ to $26\%$ at $M_K=-26$, in good agreement with \citet{mau07}. The observed trend
also agrees well with the radio-loud fraction as a function of stellar mass derived by \citet{bes05},
if $M_K=-26$ corresponds to a stellar mass of $\simeq 4\times 10^{11}~M_\odot$ \citep[e.g.,][]{bel01}.
As the relationship between radio power and $K$-band absolute magnitude is extremely steep and the mass function of galaxies has an
exponential cut-off at high masses, the most luminous radio sources will be hosted by galaxies spanning a
relatively small range of $K$-band absolute magnitude.
The distribution of radio powers at fixed $K$-band absolute magnitude is extremely broad,
spanning approximately 4 orders of magnitude. A consequence of this is that stacking of radio sources
will converge slower towards the mean or median value than one would expect for a narrower Gaussian
distribution of radio powers. A stack of 100 early-type galaxies will result in an estimate
of the mean radio power that is accurate to only $\simeq 20\%$, even if individual galaxies are detected
with high signal-to-noise.
While it is beyond the scope of this letter to explore all the possibilities that could influence
radio power, we can briefly explore some of the obvious options. At a given $K$-band absolute
magnitude, there is no clear correlation between radio power and morphological type in Figure~\ref{fig:lums},
and this conclusion is consistent with previous studies of early-type galaxies \citep[e.g.,][]{wro91}.
For $M_K<-25$ early-type galaxies, the average RC3 T types of galaxies that are above and below
the median radio power differ by just $0.6$. More subtle correlations between morphology
(including transitory features) and radio power may only be revealed by high resolution
imaging of early-type galaxies \citep[e.g.,][]{der05}.
While others have explored the correlations between radio power and
environment (e.g., dark matter halo mass) in detail, two galaxies in our sample illustrate
that any correlation between environment and radio power is likely to have considerable
scatter. Coma cluster galaxies NGC~4874 and NGC~4889 (the brightest cluster galaxy) both
have $M_K\simeq -26.2$ and RC3 T Types of -4, but their $1.4~{\rm GHz}$ flux densities
are $724~{\rm mJy}$ and $1~{\rm mJy}$ respectively. This large difference in radio flux density
for two otherwise similar galaxies suggests that the radio powers of early-type galaxies may
vary by roughly than 3 orders of magnitude over long periods of time.
\section{Weak Radio Sources in Early-Type Galaxies}
We have previously concluded that all $M_K<-25.5$ early-type galaxies are probably
radio sources, as their measured flux densities are greater than zero.
To verify the presence of weak radio sources in massive galaxies, we have stacked NVSS images of the
16 $M_K<-25.5$ early-type galaxies without counterparts in the NVSS catalogs. For comparison, we have
also stacked images of the 119 $M_K>-24$ early-type galaxies without counterparts in the NVSS catalogs.
As we show in Figure~\ref{fig:stack}, a $0.9~{\rm mJy}$ source is present in the stack of massive
galaxies while radio emission is only marginally detected in the stack of lower mass galaxies. This
is not unexpected, as the median radio power of early-type galaxies is an extremely strong function
$K$-band absolute magnitude.
The radio continuum emission from $M_K<-25.5$ early-type galaxies indicates that they are rarely (perhaps never)
truly passive, as the vast majority harbor an AGN or have recently ($\lesssim 100~{\rm Myr}$) undergone star formation.
\begin{figure}[hbt]
\begin{center}
\resizebox{3.25in}{!}{\includegraphics{ds9.ps}}
\end{center}
\caption{Median stacks of early-type galaxies with $M_K>-24$ (left) and $M_K<-25.5$ (right)
without counterparts in the NVSS catalogs. The images are $10^{\prime}$ on a side, the pixel scale
is $15^{\prime\prime}$, white corresponds to $-0.5~{\rm mJy}$ and black corresponds to $1~{\rm mJy}$.
While little radio emission is detected from the low mass galaxies, the stack of high mass galaxies
has a flux density of $0.9~{\rm mJy}$.}
\label{fig:stack}
\end{figure}
The most powerful radio sources in our sample are clearly AGNs, but the weakest radio sources could result
from star formation. Star formation is known to produce $10^{21}~{\rm W~Hz^{-1}}$ of radio emission
in some early-type galaxies \citep{wro88}, which would result from just $\sim 1~M_\odot/{\rm yr}$ of star
formation \citep{bel03}. However, there are weak arguments for AGNs being responsible for the broad distribution
of observed radio powers for $M_K<-25.5$ early-type galaxies. The radio emission in the stacked images is compact,
as one may expect from low luminosity AGNs, although the NVSS beam is $45^{\prime\prime}$ wide. \citet{sle94}
found parsec-sized radio cores in 70\% of early-type galaxies, even when the total radio power was only $10^{22}~{\rm W~Hz^{-1}}$.
If radio emission is being produced by multiple mechanisms, one may expect an obvious superposition of multiple distributions.
This should be the case for X-ray emission from early-type galaxies \citep[e.g.,][]{fab89}, but we do not see this in
the radio in Figure~\ref{fig:lums}. However, these are not definitive arguments for
the presence of radio AGNs in all massive galaxies and further observations are required.
Inconsistencies between star formation rates derived from observations at radio and other wavelengths (e.g., mid-IR)
may identify the weakest AGNs in early-type galaxies. Alternatively, AGNs may be identified using
source sizes and brightness temperatures determined with high frequency and high resolution radio imaging.
\section{Summary}
We have measured the $1.4~{\rm GHz}$ radio continuum emission from $K<9$ early-type galaxies,
utilizing archival imagery from the NVSS, Green Bank 300-ft Telescope and 64-m Parkes Radio Telescope.
The distribution of radio powers at fixed absolute magnitude spans 4 orders of magnitude
for $M_K<-24$ early-type galaxies, and the median radio power is proportional to $K$-band luminosity
to the power $2.78\pm 0.16$.
Our analysis is not restricted to galaxies with high signal-to-noise radio detections,
and the distribution of noisy flux density measurements provides important constraints
on the radio emission from early-type galaxies. Relatively few low mass early-type
galaxies are detected with high signal-to-noise, and the flux density measurements for the remaining
objects are scattered around zero. In contrast, most $M_K<-25.5$ early-type galaxies have robust
detections, and the remaining galaxies consistently have flux density measurements greater than zero.
As random measurement errors will broaden the measured distribution of flux densities, it
is highly likely that all massive galaxies are radio continuum sources. If this is the case,
all $M_K<25.5$ early-type galaxies harbor an active galactic nucleus or have recently undergone star formation.
\begin{acknowledgments}
The National Radio Astronomy Observatory is a facility of the National Science Foundation
operated under cooperative agreement by Associated Universities, Inc. The Parkes telescope is
part of the Australia Telescope National Facility which is funded by the Commonwealth of
Australia for operation as a National Facility managed by CSIRO. We thank the HIPASS
collaboration and Mark Calabretta for making their radio continuum maps available to us prior to publication.
This publication makes use of data products from the Two Micron All Sky Survey, which is a
joint project of the University of Massachusetts and the Infrared Processing and Analysis
Center/California Institute of Technology, funded by the National Aeronautics and Space
Administration and the National Science Foundation. The development of this letter was
assisted by discussions with many astronomers, including Geoffrey Bower, Scott Croom,
Steve Croft, Arjun Dey and Elaine Sadler.
\end{acknowledgments}
\bibliographystyle{astroads}
|
\section{Introduction}
\label{section1}
M\,43 (NGC1982) is an apparently spherical \hii\
region several arcmin to the northeast of the well-known Orion nebula.
HD\,37061 (NU Ori, Par 2074, Brun 747), an early B-type star with broad
lines located at the center of the nebula, is the main ionizing source
of M\,43.
The spectral classifications found in the literature for HD\,37061 range
between B0.5\,V and B1\,V, probably because of the
difficulty in detecting the faint \ion{He}{ii}\,4541 line when using low
resolution spectra and photographic plates. Up to three components have
been identified within this stellar system \citep{Abt91, Pre99}.
The primary star is known to be a spectroscopic binary (SB1) with an
estimated stellar mass ratio $M_2$/$M_1$\,$\sim$\,0.19 \citep{Abt91}.
Using bispectrum speckle
interferometry, \cite{Pre99} found a third companion at $\sim$ 470 mas
with a flux ratio in the K-band of \solu{0.03}{0.02}, implying a stellar
mass ratio $M_3$/$M_{1,2}$\,$\sim$\,0.07. Although the less massive
components may affect the spectrum of the primary, they can be neglected
in terms of ionization of the nebula.
M\,43 belongs to the same molecular complex as the Orion nebula \citep{Gou82}.
In particular, M\,43 is located at the northeast border of the extended Orion nebula (EON),
a large elliptical structure surrounding the Huygens region (corresponding to the central,
brighter part of the Orion nebula). Although the precise structure of the
EON is still unknown, recent studies \citep{Ode09, Ode10} have associated
the diffuse\footnote{The H$\alpha$ surface brightness in the EON is
$\sim$~ two orders of magnitude lower than the brightest regions in M\,42.} and
extended emission arising from this region with scattered light, produced
by dust reflecting the stellar continuum emission from the
Trapezium cluster stars and the nebular line emission from the Huygens region.
A dust lane defines the outer northern boundary of the EON and separates
M\,43 from the Huygens region. As inferred from the optical polarization
map of the M\,43 region obtained by \citet{Kha80}, this dust lane is not a
foreground obscuration but a wall of dense material between M\,42 and M\,43.
This wall of material ensures that it is quite unlikely that ionizing
light from the Trapezium cluster stars reaches the nebular material ionized
by HD\,37061;
however, owing to the proximity of M\,43 to the {\em assumed} northern
boundary of the EON, a scattered light contribution (similar
to the one described above) may still affect affect the nebular emission received from M\,43.
Most studies of M\,43 have been performed at radio and far-infrared
wavelengths. These studies were centered on the investigation of the global
properties of the nebula \citep[i.e. geometry, $T_{\rm e}$\ and $n_{\rm e}$,][]{Mil68,
Thu78, Thr86, Sub92}, and the characteristics of the dust re-radiation
inside the nebula \citep{Thu78, Smi87}. As for optical images, radio
continuum observations indicate that M\,43 has a circular shape \citep{Sub92}.
In addition, images at 60 $\mu$m published by \cite{Thr86}, and the kinematical
study by \cite{Han87} allow one to associate the nebula with a shell-shaped
cavity on the side of a dense molecular ridge.
\begin{figure}[!t]
\centering
\includegraphics[width=8.5cm,angle=0]
{16608_fig1.eps}
\caption{Color composite (RGB) image of the M\,42+M\,43 region obtained from a
combination of narrow band images taken with the WFC-INT\ (only CCDs \#1, \#2,
and \#4 are shown). The following color code was used: H$\alpha$ (red),
[\ion{S}{ii}]\,$\lambda\lambda$6716+30 (green), [\ion{O}{iii}]\,$\lambda$5007 (blue). The field-of-view of the image is
$\sim$\,22\,arcmin\,x\,33\,arcmin. M43 is the roundish nebulosity at the north-east
of the center.}
\label{f1}
\end{figure}
Its proximity\footnote{The distance to HD\,37061 and M\,43 can be assumed to be similar to that of its
companion nebula M\,42. Two different studies using
Very Long Baseline Array observations \citep{Men07, San07} determined the distance to
some stellar objects in M\,42 to be \solu{414}{7}\,pc and 389$^{+24}_{-21}$\,pc, respectively.
These new measurements are $\sim$\,10\% lower than the 450 pc distance
often assumed for the Orion nebular cloud (ONC). We refer the reader to \citet{Men07}
for a discussion of the different distances that have been determined
using different methods (viz. \citet{Jef07}: 392 pc,
based on the statistical properties of rotation in pre-MS-stars;
\citet{Gou82}: 480 pc, for a long time considered the canonical distance).},
relatively high surface brightness, simple geometry, and
isolated ionizing source make the Mairan's nebula an ideal object
to investigate several topics of interest in the field of \hii\
regions and massive stars.
To this aim, we collected a set of high quality observations comprising
the optical spectrum of HD\,37061, along with optical imaging and long-slit
spatially resolved spectroscopy of M\,43. These observations were used to
perform a comprehensive study of M\,43 and its ionizing star, which is presented
in two parts. In this first paper, we (1) obtain the stellar parameters of HD\,37061,
(2) investigate the presence of the scattered light in the region of M\,43 and
its effect on the derived nebular properties, and (3) determine the physical conditions
and nebular abundances in M\,43 obtained at various distances to the center of the nebula.
All this information is then used in a second paper (Sim\'on-D\'iaz et al., Paper II, in prep.)
to construct a specialized photoionization model of the nebula using as many
observational constraints as possible. The main drivers of this second part of our study
are to (1) test the reliability of the ionizing spectral energy distributions (SEDs)
provided by the modern stellar atmosphere codes using nebular constraints, (2)
investigate the effect of stellar pumping \citep{Fer99, Lur09} on the nebular
emission arising from the inner part of this \hii\ region, and (3) test our
understanding of the nebular energy budget temperature distribution.\\
\begin{figure*}[!t]
\centering
\includegraphics[width=17.cm,angle=0]
{16608_fig2.eps}
\caption{\ion{H}{$\alpha$}, [\ion{S}{ii}]\,$\lambda\lambda$6716+30, and [\ion{O}{iii}]\,$\lambda$5007 WFC-INT\ images of
M\,43. The size and position of the apertures extracted from the ISIS-WHT\ long-slit spectroscopic
observations are also indicated.
}
\label{f2}
\end{figure*}
This paper is structured as follows. The observational data\,set is presented in Sect.
\ref{section2}. A quantitative spectroscopic analysis of the optical
spectrum of HD\,37061 is performed in Sect. \ref{section5}. The morphological
characteristics of the nebula,
along with its physical conditions and
nebular abundances are determined in Sects. \ref{section3} and \ref{section4} by analyzing
of the {H$\alpha$}, {H$\beta$}, [\ion{O}{iii}], and [\ion{S}{ii}] narrow-band images, and the nebular
spectra, respectively. Both nebular analyses also allow us to find several indications
of an extended nebular emission, not directly related to M\,43,
which must be adequately subtracted for a correct interpretation of the nebular images
and spectroscopy. Finally, we outline the main results of this study in Sect. \ref{section7}.
\section{The observational data\,set}\label{section2}
\subsection{Nebular imaging}
\label{section21}
\begin{table}[t!]
\begin{center}
\caption{\footnotesize Summary of the observations collected for the study.
\label{t1}
}
\scriptsize{
\begin{tabular}{ccccc}
\hline
\hline
\multicolumn{5}{c}{Nebular imaging} \\
\hline
\noalign{\vskip2pt}
Instrument & Filter & \# & $\lambda_0$ (\AA) & FWHM (\AA) \\
(Telescope) & & & & \\
\noalign{\vskip1pt}
\hline
WFC & {H$\alpha$} & 197 & 6568 & 95 \\
(INT) & {H$\alpha$}\ redsh. & 228 & 6805 & 93 \\
(0.33 arcsec/pix) & {H$\beta$}\ narrow & 225 & 4861 & 30 \\
& {H$\beta$}\ broad & 224 & 4861 & 170 \\
& [\ion{O}{iii}] & 196 & 5008 & 100 \\
& [\ion{S}{ii}] & 212 & 6725 & 80 \\
\hline
\multicolumn{5}{c}{Nebular spectroscopy} \\
\hline
\noalign{\vskip2pt}
& Grid & $\lambda_0$ (\AA) & Spect. range (\AA) & $\Delta\lambda$ (\AA/pix)\\
\noalign{\vskip1pt}
\hline
ISIS & R600B & 4368 & 3386--5102 & 0.45 \\
(WHT) & R600R & 6718 & 6064--7585 & 0.49 \\
\hline
\multicolumn{5}{c}{Stellar spectroscopy} \\
\hline
\noalign{\vskip2pt}
& Grid & $\lambda_0$ (\AA) & Spect. range (\AA) & $\Delta\lambda$ (\AA/pix)\\
\noalign{\vskip1pt}
\hline
IDS & H\,2400B & 4320 & 4060 - 4590 & 0.24 \\
(INT) & H\,2400B & 4800 & 4550 - 5070 & 0.24 \\
& H\,1800V & 6400 & 6090 - 6760 & 0.31 \\
\hline
\hline
\end{tabular}
\\
}
\end{center}
\end{table}
The Wide Field Camera\footnote{\tt{http://www.ing.iac.es/Astronomy/instruments/wfc/}} (WFC)
attached to the Isaac Newton Telescope (INT)
at the Roque de los Muchachos Observatory (La Palma, Spain) was
used to obtain several narrow-band filter images of M\,43 in 2004 February 16, and
2006 October 21. The
WFC-INT\ consists
of four thinned EEV 2kx4k CCDs located at the primary focus of the telescope.
Each CCD yields a field of view of 11.2' $\times$ 22.4', which results
in a edge to edge limit of the mosaic (neglecting the $\sim$1 arcmin inter-chip
spacing) of 34.2 arcmin. The pixel size is 13.5\,$\mu$m corresponding to
0.33 arcsec/pixel.
We centered M\,43 on chip\#4 of the WFC, and obtained images of the
region using the narrow-band filters H$\beta$, [\ion{O}{iii}]\,$\lambda$5007,
H$\alpha$, and [\ion{S}{ii}]\,$\lambda\lambda$6716+30.
The log of the observations, along with the
characteristics of the various filters can be found in Table \ref{t1}.
While the [\ion{O}{iii}] and [\ion{S}{ii}] images were only obtained to
provide some qualitative information about the nebula, we wished to produce
pure emission-line, flux-calibrated images for H$\alpha$ and H$\beta$.
We hence obtained these images during a photometric night, and included
the H$\alpha$ redshifted, and H$\beta$ broad filters in our set
of observations, to correct the H$\alpha$ and H$\beta$ images
for the adjacent continuum. The exposure times considered for the images in the
various filters were 4$\times$60\,s for {H$\alpha$},
4$\times$120\,s for {H$\alpha$}\ continuum, 3$\times$120\,s for {H$\beta$}, 3$\times$180\,s for
{H$\beta$}\ continuum, 5$\times$12\,s for [\ion{O}{iii}], and 5$\times$12\,s for [\ion{S}{ii}].
The reduction of the images was performed following the standard procedures
(trimming, bias subtraction, flat-fielding, and alignment) with
\iraf\footnote{\iraf\ is distributed by NOAO which is operated by AURA Inc., under
cooperative agreement with NSF}. H$\alpha$ and H$\beta$ images were flux
calibrated using observations of the spectrophotometric standard star BD\,+28$^{\circ}$4211
\citep{Lan07} at different airmasses. We obtained the
continuum-subtracted {H$\alpha$}\ and {H$\beta$}\ images following the procedure described in
\cite{Lop08}.
The large field of view of the WFC\ also allowed us to include M\,42 in the images.
A color-composite image of the M\,43+M\,42 region obtained from a combination
of the [\ion{O}{iii}] (blue), H$\alpha$ (green), and [\ion{S}{ii}] (red) images is
presented in Fig.~\ref{f1}. Individual images in each one of these three filters, showing a square
region of $\sim$\,6 arcmin centered in M\,43, are presented in Fig.~\ref{f2}.
\subsection{Stellar spectroscopy}
\label{section23}
\begin{figure}[t]
\centering
\includegraphics[height=8.5cm,angle=0]
{16608_fig3.eps}
\caption{IDS-INT\ optical spectrum of HD\,37061. The \ion{H}{i} and \ion{He}{i-ii} lines
used for the spectroscopic analysis of the star are indicated.}
\label{f4}
\end{figure}
The spectroscopic observations of HD\,37061 were carried
out with the INT\ on 2006 August 30 and September 5. The Intermediate
Dispersion Spectrograph (IDS) was used with the 235\,mm camera and two
different gratings (see Table \ref{t1}). We observed the spectral region between
$\lambda\lambda$4000 and 5050\,\AA\
using the H2400B grating, which resulted in an effective resolving
power R\,$\sim$\,7500 (equivalent to a 0.24\,\AA/pixel resolution
and $\sim$\,2.6\,pixel \fwhm\ arc lines). For the {H$\alpha$}\ region
($\sim$\,$\lambda\lambda$6090\,--\,6760\,\AA), the H1800V
grating was used, resulting in a similar spectral resolution (0.31 \AA/pixel,
R\,$\sim$\,8000). With these configurations, three exposures were
needed to cover the whole range. We obtained spectra with exposure times of 180\,s and 70\,s for the blue (2)
and red (1) configurations, respectively, and obtained two spectra per spectral
range. A large number of flat fields and
arcs for the data reduction process were obtained.
The reduction and normalization of the spectra was made following
standard techniques, with \iraf\ and our own software developed in \idl.
The signal-to-noise ratio of the reduced spectra is about 300\,--\,400, depending on the
spectral range. Fig.~\ref{f4} shows the global stellar spectrum, along with the \ion{H}{i}
and \ion{He}{i-ii} lines used for the quantitative spectroscopic analysis.
\subsection{Nebular spectroscopy}
\label{section22}
\begin{figure}[t]
\centering
\includegraphics[width=8.5cm,angle=0]
{16608_fig4.eps}
\caption{ISIS-WHT\ optical spectrum of M\,43 (aperture \#5). The \ion{H}{i} Balmer
lines, along with \ion{He}{i} and other metal lines used in our study, are indicated.}
\label{f3}
\end{figure}
A long-slit, intermediate-resolution spectrum of M\,43 was obtained on
2006 December 23 with the Intermediate dispersion Spectrograph and Imaging
System (ISIS) spectrograph attached to the 4.2m William Herschel
Telescope (WHT) at Roque de los Muchachos Observatory (La Palma, Spain).
The 3.7$'$\,$\times$\,1.02$''$ slit
was located to the west of HD\,37061, along the E-W direction (PA=90$^{\circ}$,
see Fig.~\ref{f2}). Two different CCDs were used at the blue and red arms of
the spectrograph: an EEV CCD with a configuration 4096 $\times$ 2048 pixels at
13.5 $\mu$m, and a RedPlus CCD with 4096 $\times$ 2048 pixels
at 15.0 $\mu$m, respectively. The dichroic used to separate the blue and red
beams was centered on 5400 \AA. The gratings R600B and R600R were used for the blue
and red observations, respectively (see log of the observations in Table \ref{t1}).
These gratings give a reciprocal dispersion of 33 \AA\ mm$^{-1}$ in both
cases, and effective spectral resolutions of 2.2 and 2.0 \AA, respectively.
The blue spectra cover from $\lambda\lambda$3386 to 5102 \AA\ and the red ones
from $\lambda\lambda$6064 to 7585 \AA. The spatial scales are 0.20$''$ pixel$^{-1}$
and 0.22$''$ pixel$^{-1}$, respectively. The seeing during the
observations was between $\sim$0.5$''$ and $\sim$0.8$''$. The exposure times were
3$\times$300\,s, in both the blue and red observations (obtained at the same time).
The spectra were wavelength calibrated with a CuNe+CuAr lamp. The correction
for atmospheric extinction was performed using the average curve for continuous
atmospheric extinction at Roque de los Muchachos Observatory. The absolute
flux calibration was achieved by observations of the standard stars
HD\,19445, H\,600, and HD\,84937. We used the standard \iraf\ TWODSPEC reduction
package to perform bias correction, flat-fielding, cosmic-ray rejection, wavelength,
and flux calibration. Fig.~\ref{f3} shows an illustrative example of the
global appearance of our nebular spectroscopic observations, also indicating
the main nebular lines used in this study.
\section{Quantitative spectroscopic analysis of HD\,37061}
\label{section5}
The stellar parameters of the star were derived by visually
comparing the observed optical \ion{H}{i}, \ion{He}{i}, and \ion{He}{ii}
line profiles with the synthetic ones resulting from the stellar atmosphere
code \fastwind\ \citep{San97, Pul05}, which is an established technique.
Similar analyses, along with some notes on
the methodology, can be found in \cite{Her02}, \cite{Rep04}, \cite{Sim06},
and references therein.
To this aim, we constructed a grid of \fastwind\ models with \Teff, and \grav,
ranging from 28000 to 33000 K (500 K steps), and from 3.9 to 4.3 dex (0.1 dex steps), respectively. The microturbulence,
the helium abundance, and the wind-strength parameter\footnote{$Q\,=\,\dot{M}/(R\,v_{\infty})^{1.5}$.}
were fixed to characteristic values for an early B dwarf star (\micro\,=\,5 \kms,
$\epsilon$\,=\,0.09, and log\,Q\,=\,-15, respectively) and metal abundances were assumed to be solar
\citep[following the set of abundances by][]{Asp09}.
This technique requires the projected rotational velocity (\vsini) of the star to be
previously established. All metal lines in the spectrum of HD\,37061 are blended or
very shallow because of the large \vsini\ of the star; therefore, we decided to apply the
Fourier method (c.f. \citeauthor{Gra76} \citeyear{Gra76}; see also \citeauthor{Sim07}
\citeyear{Sim07} for a recent application to OB-type stars) to the \ion{He}{i}
lines, and obtained a \vsini\,$\sim$\,200 \kms.
\begin{figure}[t!]
\centering
\includegraphics[width=8.5cm,angle=0]{16608_fig5.eps}
\caption{Fit of FASTWIND \ion{H}{i} and \ion{He}{i-ii} synthetic line profiles to
the observed ones (red solid line, \Teff=31000 K, \grav=4.2 dex). Blue
dashed lines illustrate the effect of a variation of $\pm$1000 K in the
FASTWIND models. Note that the red wings of the \ion{H}{i} and \ion{He}{i}
lines are affected by the cool companion.}
\label{f_fit}
\end{figure}
\begin{table}[t!]
\begin{center}
\caption{\footnotesize Stellar parameters derived through the
\fastwind\ analysis of the optical spectrum of HD\,37061.}
\label{t5}
\begin{tabular}{l c | l c}
\hline
\hline
\noalign{\smallskip}
\Teff (K) & \solu{31000}{500} & R (R$_{\odot}$) & \solu{5.7}{0.8} \\
\grav (dex) & \solu{4.2}{0.1} & logL/L$_{\odot}$ & \solu{4.42}{0.12} \\
$\epsilon$(He) & 0.09 (assumed) & M (M$_{\odot}$) & \solu{19}{7} \\
log\,Q & -15 (assumed) & log\,Q(H$^0$) & \solu{47.2}{0.2} \\
\noalign{\smallskip}
\hline
\hline
\\
\end{tabular}
\end{center}
\end{table}
The synthetic \ion{H}{i} and \ion{He}{i-ii} line profiles resulting from the models were then
convolved with the corresponding instrumental and rotational profiles and
compared with the observations.
The best fit was found for \Teff\,=\,31000 K and \grav\,=\,4.2 dex (see
Fig. \ref{f_fit}). Given the quality of the spectrum and the sensitivity of
the used lines to the variation in \Teff\ and \grav\ for this range of
stellar parameters, an accuracy of 500 K and 0.1 dex, respectively,
could be achieved. As an example, in Fig. \ref{f_fit} we show the
effect of a variation of $\pm$1000 K in the FASTWIND models. For this
range of stellar parameters, the \ion{He}{ii} lines are decisive in constraining
the \Teff.
We note the poor quality of the fits in the red wings of the \ion{H}{i} and \ion{He}{i}
lines (but not in the \ion{He}{ii} lines). These lines are likely affected by the spectrum
of the cooler secondary spectroscopic component (see Sect. \ref{section1}).
Once the stellar parameters have been determined, the stellar
radius can be derived from the absolute visual magnitude ($M_{\rm V}$) and the
synthetic spectral energy distribution provided by the stellar atmosphere model.
By using $m_{\rm V}$=6.84 and $A_{\rm V}$=2.09
\citep{Hil97}, and adopting a distance to the star of 400$\pm$50 pc, we obtained
an $M_{\rm V}$\,=\,-3.3$\pm$0.3, quite consistent with the spectral type of HD\,37061.
The resulting stellar parameters, along with the total number of ionizing photons
are summarized in Table \ref{t5}.
\section{Qualitative and quantitative analysis of the nebular images}
\label{section3}
Fig.~\ref{f1} shows a color composite image of the M\,42+M\,43 region.
The object under study, M\,43, appears as a roundish \hii\ region
centered around HD\,37061, and well-separated from the Orion nebula (M\,42)
by a dust lane (known as the northeast dark lane). The nebula has a
diameter of $\sim$4.5\,arcmin ($\sim$\,0.65\,pc at a distance of 400\,pc), and is
crossed by a dark lane oriented N-S in the eastern side (known as the M43 dark lane).
This dark cloud is located in front of the nebula, blocking the nebular
light coming from behind.
Fig.~\ref{f2} shows a closer view of the nebula, where three images of M\,43, taken
through filters isolating H$\alpha$ (+[\ion{N}{ii}]\,$\lambda\lambda$6548+84),
[\ion{S}{ii}]\,$\lambda\lambda$6716+30, and [\ion{O}{iii}]\,$\lambda$5007, are presented
separately. The images show diffuse and extended emission
beyond the limits of the nebula (mainly to the south and west). This
emission can be more clearly seen in the [\ion{O}{iii}] image. If the entire nebula were spherical
with a density of 500 cm$^{-3}$,
the size of the O$^{2+}$ region would be $\sim$\,25\,\% of the total
size of the nebula\footnote{Computed from a simple spherical \cloudy\ model using the
ionizing spectral energy distribution from a FASTWIND model with \Teff\,=\,31000\,K, and \grav\,=\,4.2.};
however, Fig.~\ref{f2} shows that the [\ion{O}{iii}] emission originates in a more extended
region (i.e. is unlikely to be related\footnote{This extended emission is more likely to be associated
with the scattered nebular light arising from the EON, described by \citet{Ode09} and \citet{Ode10}.
See a more detailed discussion about this possibility in Sect. \ref{section49}.}
to nebular material ionized by HD\,37061).
To more clearly illustrate this, we plot in Fig.~\ref{f5} (three upper panels) the spatial distribution of
{H$\alpha$}, [\ion{S}{ii}], and [\ion{O}{iii}] emission in a line passing by HD\,37061 along
the east-west direction, obtained from the corresponding WFC-INT\ images. Since
we are only interested in the relative spatial changes, the three distributions
are normalized to their maximum value. In all cases, a faint emission,
not expected to be related to M\,43, is clearly detected beyond the limits
of the nebula. In the [\ion{O}{iii}] panel, we also superimpose the spatial distribution
derived from the {H$\alpha$}-continuum image to show that the extended emission observed in the [\ion{O}{iii}]
image is not associated with the nebular continuum.
\begin{figure}[!t]
\centering
\includegraphics[width=8cm,angle=0]
{16608_fig6.eps}
\caption{Three upper panels: Spatial distribution of {H$\alpha$}, [\ion{S}{ii}]\,$\lambda\lambda$6716+30,
and [\ion{O}{iii}]\,$\lambda$5007 emission in a line passing by HD\,37061 in the west-east
direction (see Fig.~\ref{f2}), obtained from the corresponding WFC-INT images.
The dashed red line indicates the position of HD\,37061. The vertical dotted
line indicate the western limit of M\,43. Bottom panel:
{H$\alpha$}/{H$\beta$}\ spatial distribution in the same line direction. Some dust features
outlined by \citet{Smi87} and the position and size of the ISIS-WHT slit are
indicated. Horizontal dotted lines show the {H$\alpha$}/{H$\beta$}\ value in the M43 west
region.}
\label{f5}
\end{figure}
\subsection{Extinction}
\label{section31}
The flux-calibrated, continuum-subtracted H$\alpha$ and H$\beta$
images (i.e. pure emission) were used to obtain the H$\alpha$/H$\beta$ flux ratio
distribution along the diameter in the east-west direction passing
by the ionizing star of M\,43 (see bottom panel in Fig.~\ref{f5}).
Some interesting information about the extinction in the studied region can be
extracted by inspecting of this figure:
\begin{itemize}
\item[a)]{There is a jump in the H$\alpha$/H$\beta$ flux ratio
140 arcsec west of HD37061 (i.e. western boundary of M\,43), indicating
a clear difference in the type of emission arising from outside the limits of
the nebula. We note that the lower H$\alpha$/H$\beta$ ratio
found in the ``extended emission'' region agrees with the hypothesis
that this emission is associated with scattered light (see also Sect. \ref{section49}),
and is not necessarily an indication of a lower extinction in this region.}
\item[b)]{The H$\alpha$/H$\beta$ ratio in the western region of M\,43 is fairly
constant, though there is a small decrease close to the star. On the other hand,
this ratio is larger in the eastern part.}
\item[c)]{The M\,43 dark lane location (labeled as a foreground dust lane in
Fig. \ref{f5}) is clearly represented in the H$\alpha$/H$\beta$ distribution.}
\end{itemize}
\citet{Smi87} presented a general view of the spatial distribution of dust in the
M\,43 region. They found
three regions where dust reradiation in 60 $\mu$m is concentrated (see a schematic
map of dust features in their Fig 1.c). The first one is associated with the M\,43 dark lane
(they called it the foreground dust lane), the second one (M\,43 east) is located
to the east of HD\,37061, between the ionizing star and the M\,43 dark lane,
and a third one (M\,43 front\,+\,M\,43 west) follows the border of the
nebula from the north to the west. This third feature of dust emission is
separated $\sim$\,60-70 arcsec from the star, leaving a ``dust-empty'' space
in-between. We indicate in the bottom panel of Fig. \ref{f5} the location
of these dust features using the same nomenclature as \citeauthor{Smi87}
The observed behavior of the H$\alpha$/H$\beta$ flux ratio within M\,43 along the
west-east direction is perfectly correlated with the dust distribution presented
by \citet{Smi87}. It is remarkable that, in contrast to expectations, the amount of
extinction indicated by the H$\alpha$/H$\beta$ flux ratio in the foreground
dust lane is very low. The explanation is simple. This dark feature obscures
the emission from M\,43, hence the observed light originates in the material
located in front. This emission follows the trend indicated by the region marked as
``extended emission''.
\subsection{Total nebular H$\alpha$ luminosity and surface brightness}
\label{section32}
\begin{figure}[t!]
\centering
\includegraphics[height=8.cm,angle=90]{16608_fig7.eps}
\caption{Top and middle panels: integrated {H$\alpha$}\ luminosity profile and extinction
corrected {H$\alpha$}\ surface brightness profile, respectively, obtained from the nebular image.
Bottom panel: reddening coefficient obtained from the empirical analysis of the nebular
spectra (see Sect.~\ref{section4}).
In the upper panels we compare the L({H$\alpha$}) and S({H$\alpha$}) profiles of the original
images (dotted lines) with those of images corrected for the contribution of the
extended emission (solid lines). Some data obtained from
the ISIS-WHT spectroscopic observations are also included as triangles and
diamonds (see text for explanation).}
\label{f6}
\end{figure}
To derive the total nebular H$\alpha$ luminosity we proceeded as
follows. First, the H$\alpha$ image was corrected for the effects of distance,
extinction, and [\ion{N}{ii}] contamination\footnote{Note that what we
refer to as a ''H$\alpha$ image'' is actually a ''H$\alpha$+[\ion{N}{ii}]\,$\lambda\lambda$6548+84''
image.}. Denoting with $F_{\rm H\alpha}$ the nebular flux received at Earth,
the corresponding extinction-corrected H$\alpha$ luminosity can be derived from:
\begin{equation}
L_{\rm H\alpha}=4\pi d^2 F_{\rm H\alpha} 10^{c_{\rm H\beta}\,{f(\rm H\alpha)}},
\end{equation}
where $d$ is the distance to the nebula, $c$({{H$\beta$}}) is the reddening coefficient, and $f(\rm H\alpha)$ is
the value of the extinction function in H$\alpha$ relative to H$\beta$.
The analysis of the H$\alpha$/H$\beta$ image (Section \ref{section31})
showed that the extinction is not constant across the nebula. The
optimal strategy to follow would be to
obtain $c$({{H$\beta$}}) for every pixel from the H$\alpha$/H$\beta$ images,
correct the H$\alpha$ image, and then integrate the whole nebula to obtain the
total L(H$\alpha$).
However, when we attempted to follow this strategy we found that
it introduces many sources of uncertainty. In addition, the nebular region
hidden by the dark lane cannot be corrected in this way. We therefore decided to
follow a different approach, taking into account what we learned from the inspection
of the nebular images, and making use of the information extracted from the
spectroscopic observations (see Sect.~\ref{section4}). We decided to integrate only the
west quadrant of the nebula, and multiply the resulting value by four (i.e.
we assume the nebula is symmetric). Extinction in this quadrant
is lower and more constant than in the other three. Furthermore, we have information
about $c$({{H$\beta$}}) ($\sim$\,0.76) and the nitrogen contribution to the {H$\alpha$}\ image
([\ion{N}{ii}]\,$\lambda\lambda$6548+84/H$\alpha$\,$\sim$\,0.59) for this region from
the ISIS-WHT spectroscopy (see Sect.~\ref{section4}).
Once the distance, extinction, and nitrogen corrections had been applied to the continuum-subtracted,
flux-calibrated {H$\alpha$}\ image,
we obtained the L(H$\alpha$) inside circles of increasing sizes\footnote{Actually, we
integrated the mentioned quadrant and multiply the obtained value by four.} (centered
on the star), and the surface brightness distribution within concentric rings of
increasing radii. Fig.~\ref{f6} shows the measured quantities as a function
of distance to the star. Dotted lines show the initial determinations where the images
were not corrected for the extended emission contribution. The {H$\alpha$}\ surface
brightness distribution again shows
that outside the limits of the nebula ($\theta\,\ge$\,160 arcsec) there remains a non-negligible,
more or less constant emission. This implies that L(H$\alpha$) increases continuously, even
outside the nebula (top panel in Fig.~\ref{f6}). We therefore assumed a
constant\footnote{We note
that the H$\alpha$ surface brightness of the extended emission is not necessarily constant
across the nebula; however, its variation is negligible compared to the H$\alpha$
emission from M\,43, hence can be considered to be constant for the sake of simplicity.}
contribution of the extended emission to the surface brightness ($\sim$\,0.06\,erg\,s$^{-1}$\,cm$^{-2}$, see
middle panel of Fig. \ref{f6}) and subtracted it from the {H$\alpha$}\
image. The resulting L(H$\alpha$) and S(H$\alpha$) distributions, calculated from the
corrected images in the same way as described above, are indicated as solid lines in the top and middle panels
of Fig.~\ref{f6}. These quantities behave as expected, i.e., the surface brightness
becomes zero, and L(H$\alpha$) remains constant for R$>$R$_{\rm neb}$. The resulting
total {H$\alpha$}\ luminosity is (3.3$\pm$1.1)\,x\,10$^{\rm 35}$ erg\,s$^{\rm -1}$, where the
associated uncertainty was calculated by taking into account uncertainties of
12.5\%, 10\%, and 20\% in distance, $F$({H$\alpha$}), and $c$({{H$\beta$}}), respectively.
The bottom panel of Fig.~\ref{f6} shows the $c$({{H$\beta$}}) values obtained from the
spectroscopic analysis of the 9 apertures extracted from the long-slit ISIS-WHT
observations (see Sect.~\ref{section42}). As for the {H$\alpha$}/{H$\beta$}\
images (bottom panel in Fig.~\ref{f5}), we found that $c$({{H$\beta$}})
is fairly constant for $\theta\,>$\,50\,arcsec, but is somewhat smaller in the
inner region (close to the star). We also include some information obtained from
the spectroscopy in the middle panel of Fig.~\ref{f6} (surface brightness profile).
Red triangles correspond to the calculated S({H$\alpha$}) in each aperture assuming a constant
$c$({{H$\beta$}}) of 0.76. The derived values follow the
S({H$\alpha$}) profile obtained from the {H$\alpha$}\ image. The middle panel in Fig.~\ref{f6} also
shows the corresponding S({H$\alpha$}) values computed from the actual $c$({{H$\beta$}})
values derived from the spectroscopic analysis (black diamonds).
These values deviate from the constant extinction-corrected values in the inner part
of the nebula. We can conclude that the increase
in S(H$\alpha$) in the inner part of the nebula is not necessarily caused by an
increase in the emission measure or a deviation from the spherical geometry, but is more
likely to be an effect of a non-constant extinction or stellar pumping effects on H recombination
lines \citep[see][]{Fer99, Lur09}.
The total L({H$\alpha$}) we derived above must hence be corrected for this
effect, resulting in (3.0$\pm$1.1) x 10$^{\rm 35}$ erg\,s$^{\rm -1}$ (about 10\%
below the previously derived value).
In Sect. \ref{section5}, we derived the total number of ionizing photons emitted by
the central star (Q(H$^0$)\,=\,10$^{47.2\,\pm\,0.2}$ photons\,s$^{-1}$). This
would imply a maximum nebular H$\alpha$ luminosity of (2.5$\pm$1.0)\,x\,10$^{35}$
erg\,s$^{-1}$, a value that is in quite good agreement with the total H$\alpha$
luminosity obtained from the analysis of the nebular images. This result is
compatible to first order with M\,43 being a (mostly) ionization-bounded
nebula in which dust does not absorb any significant fraction of the Lyman
continuum photons.
We note that the stellar and nebular results would be incompatible if we
considered the total nebular H$\alpha$ luminosity obtained from the images
that are not corrected for the extended emission contribution. The total number of ionizing
photons necessary to explain the derived nebular H$\alpha$ luminosity would
be at least double the value obtained from the analysis of the star.
\section{Empirical analysis of the nebular spectra}
\label{section4}
\subsection{More indications of extended nebular emission}
\label{section41}
Fig.~\ref{f7} shows the spatial distribution of the nebular emission in four selected
lines ({H$\alpha$}, [O~{\sc ii}]\,$\lambda\lambda$3720+30, \ion{He}{i}\,$\lambda$6678, and
[O~{\sc iii}]\,$\lambda$5007) and their respective adjacent nebular continua. These
distributions were obtained
from the ISIS@WHT spectroscopic observations. The colored lines represent the
measured line fluxes (blue), the adjacent continuum emission (red), and the continuum-subtracted
line fluxes (black). These figures display non-negligible emission beyond the
limits of the nebula ($\theta\,>\,$150 arcsec), that is not
associated with the continuum emission. In
most cases, this external emission remains fairly constant in the studied
region (e.g. {H$\alpha$}, [O~{\sc ii}]\,$\lambda\lambda$3720+30, \ion{He}{i}\,$\lambda$6678, as
well as [N~{\sc ii}], [S~{\sc ii}], and [S~{\sc iii}]\ lines). However, in some cases (e.g. [O~{\sc iii}]),
this emission increases with increasing distance from the star.
\subsection{Aperture selection and line flux measurements}
\label{section42}
\begin{figure}[!t]
\centering
\includegraphics[width=8.5cm,angle=0]
{16608_fig8.eps}
\caption{Spatial distribution of the nebular emission in four selected
nebular lines ({H$\alpha$}, [O~{\sc ii}]\,$\lambda\lambda$3720+30, \ion{He}{i}\,$\lambda$6678, and
[O~{\sc iii}]\,$\lambda$5007) and their respective adjacent nebular continua, obtained from the ISIS-WHT
spectroscopic observations. The colored lines represent the measured line fluxes (blue),
the adjacent continuum emission (red), and the continuum-subtracted line fluxes (black).
The apertures used to extract nebular spectra at different distances from the central star
are also indicated (see Sect.~\ref{section42}). Apertures labeled as v1, v2, v3, and v4, are
those used to correct the other nine apertures for the contamination by the extended emission component.}
\label{f7}
\end{figure}
We obtained one-dimensional (1D) spectra of regions of the nebula at different distances
from the central star by dividing the long-slit used for the ISIS-WHT\ nebular
observations into nine small apertures within the limits of the nebula ($\theta<$150 arcsec)
with the aim of obtaining: a) the radial distribution of the physical
properties of the nebula, b) information about nebular ionic and total abundances,
and c) the radial distribution of critical nebular line ratios used as observational
constraints on photoionization models of the nebula.
The size and position of the apertures are summarized in Table \ref{t2}
and shown in Figs. \ref{f2} and \ref{f7}. It is crucial to correct these spectra
for the contribution of the extended emission component to obtain
information about M\,43 itself. To achieve this aim, spectra from two extra-apertures outside
the nebula were extracted to correct the line flux measurements
of the apertures for the extended emission component. The size of the extra-apertures
was selected to be the same as the corresponding apertures inside the nebula (11 and 16.5 arcsec
for the three innermost and the other outer apertures, respectively). Two different
locations of these extra-apertures were considered depending on whether the considered
nebular emission line had been emitted in the whole nebula (i.e. {H\,{\sc i}}, {[O~{\sc ii}]}, {[N~{\sc ii}]}, {[S~{\sc ii}]}
and {[S~{\sc iii}]} lines; apertures v1 and v2 in Fig,~\ref{f7}), or only in the inner region
(i.e. {He~{\sc i}}, {[O~{\sc iii}]}, and {[Ar~{\sc iii}]} lines; apertures v3 and v4 in Fig.~\ref{f7}).
For the latter set of lines, it is impossible to use the same
apertures (v1 and v2) as for the other lines, since the extended emission in these
cases is not constant, but increases far away from the star. We hence considered
two apertures as being close to the region where we were confident that the contribution of
the nebular emission associated with M\,43 had vanished. We expect the spectrum from
these apertures to represent the properties of the extended emission close to the
region where M\,43 is emitting in the given nebular line.
The extraction of 1D spectra from each aperture was performed using the \iraf\
task $apall$. The same zone and spatial coverage was considered in the blue
and red spectroscopic ranges.
We detected {H\,{\sc i}} and {He~{\sc i}} optical recombination lines, along with
collisionally excited lines (CELs) of several ions, such as {[O~{\sc i}]}, {[O~{\sc ii}]},
{[O~{\sc iii}]}, {[N~{\sc ii}]}, {[S~{\sc ii}]}, {[S~{\sc iii}]}, {[Ne~{\sc iii}]}, and {[Ar~{\sc iii}]}. Line fluxes were
measured using the SPLOT routine of the \iraf\ package by integrating all the flux
included in the line profile between two given limits and over a local continuum
estimated by eye. Although we detected two {[Ne~{\sc iii}]} lines, these were unsuitable
for analysis because they were extremely faint (in the case of $\lambda$3869) or
were severely blended with {H\,{\sc i}}$\lambda$3967.
In the case of {[O~{\sc ii}]} $\lambda$7330 \AA , we corrected
the flux for the overlapping telluric emission line
OH 8--3 P2 $\lambda$7329.148 \AA\ by estimating
its intensity from the observed OH 8--3 P2 $\lambda$7329.148 \AA /OH 8--3
P1 $\lambda$7340.885 \AA\ line ratio in standard star exposures.
Each emission line in the spectra from the nine apertures was corrected for the
extended emission component by subtracting the corresponding emission measured in the appropriate
extra-aperture (v1 to v4). Line intensities were then normalized to a particular {H\,{\sc i}}
recombination line present in each wavelength interval ({{H$\beta$}} and {{H$\alpha$}} for
the blue and red spectra, respectively).
To produce a final homogeneous set of line intensity ratios, the red spectra were then
rescaled to {H$\beta$}\ using the theoretical {H$\alpha$}/{H$\beta$}\ ratio I({H$\alpha$})/I({H$\beta$})\,=\,2.92 obtained
for the physical conditions of $T_e$\,=\,7500 K and $n_e$\,=\,500 cm$^{-3}$ (see Sect. \ref{section45}).
\begin{table*}
\centering \caption{Line intensities corrected from foreground emission and extinction ({H$\beta$}\,=100), and
results from the empirical analysis of the nebular spectra corresponding to the nine extracted apertures$^{(1)}$. }
\label{t2}
\scriptsize
\begin{tabular}{c@{\hspace{2.8mm}}c@{\hspace{2.8mm}}c@{\hspace{2.8mm}}c@{\hspace{2.8mm}}c@{\hspace{2.8mm}}c@{\hspace{2.8mm}}c@{\hspace{2.8mm}}c@{\hspace{2.8mm}}c@{\hspace{2.8mm}}c@{\hspace{2.8mm}}c@{\hspace{2.8mm}}c@{\hspace{2.8mm}}}
\noalign{\hrule}
\noalign{\vskip3pt}
& & & \multicolumn{9}{c}{Aperture} \\
\cline{4-12}
\noalign{\vskip3pt}
& & & A1 & A2 & A3 & A4 & A5 & A6 & A7 & A8 & A9 \\
\cline{4-12}
\noalign{\vskip3pt}
\multicolumn{3}{r}{Center position (arcsec)} & 15.80 & 26.80 & 37.80 & 51.55 & 68.05 & 84.55 & 101.05 & 117.55 & 134.05 \\
\multicolumn{3}{r}{Size (arcsec)}& 11.0 & 11.0 & 11.0 & 16.5 & 16.5 & 16.5 & 16.5 & 16.5 & 16.5 \\
\noalign{\vskip3pt}
\noalign{\hrule}
\noalign{\vskip3pt}
$\lambda$ (\AA) & Ion & Mult. & \multicolumn{7}{c}{$I$($\lambda$)/$I$(H$\beta$)} \\
\noalign{\vskip3pt} \noalign{\hrule} \noalign{\vskip3pt}
3726.03 & {[O~{\sc ii}]} & 1F & 93.5$\pm$1.8& 99.7$\pm$2.0 & 106$\pm$2 & 104$\pm$2 & 104$\pm$2 & 108$\pm$2 & 117$\pm$2 & 129$\pm$3 & 150$\pm$3 \\
3728.82 & {[O~{\sc ii}]} & 1F & 84.2$\pm$1.7& 91.5$\pm$1.8 & 98.1$\pm$2.0 & 99.5$\pm$2.0 & 98.9$\pm$2.0 & 103$\pm$2 & 114$\pm$2 & 120$\pm$2 & 143$\pm$3 \\
3835.39 & {H\,{\sc i}} & H9$^{(2)}$ & 7.50$\pm$0.75& 7.94$\pm$0.79 & 8.28$\pm$0.83 & 7.91$\pm$0.79 & 7.83$\pm$0.78 & 7.57$\pm$0.76 & 7.98$\pm$0.80 & 7.37$\pm$0.74 & 7.42$\pm$0.74 \\
4068.60 & {[S~{\sc ii}]} & 1F & 1.58$\pm$0.32& 1.80$\pm$0.36 & 1.86$\pm$0.37 & 2.43$\pm$0.49 & 2.70$\pm$0.54 & 3.30$\pm$0.66 & 3.84$\pm$0.77 & 4.77$\pm$0.95 & 5.63$\pm$0.56 \\
4076.35 & {[S~{\sc ii}]} & 1F & 0.64$\pm$0.19& 0.67$\pm$0.20 & 0.76$\pm$0.23 & 0.74$\pm$0.22 & 0.71$\pm$0.21 & 1.00$\pm$0.30 & 1.16$\pm$0.23 & 1.29$\pm$0.26 & 1.24$\pm$0.25 \\
4101.74 & {H\,{\sc i}} & H$\delta$ & 26.2$\pm$1.3 & 26.2$\pm$1.3 & 26.0$\pm$1.3 & 25.7$\pm$1.3 & 25.6$\pm$1.3 & 25.6$\pm$1.3 & 25.7$\pm$1.3 & 25.6$\pm$1.3 & 25.8$\pm$1.3 \\
4340.47 & {H\,{\sc i}} & H$\gamma$ & 46.2$\pm$2.3 & 46.2$\pm$2.3 & 46.5$\pm$2.3 & 46.9$\pm$2.3 & 46.9$\pm$2.4 & 47.0$\pm$2.4 & 46.8$\pm$2.3 & 46.9$\pm$2.4 & 46.8$\pm$2.3 \\
4471.09 & {He~{\sc i}} & 14 & 1.09$\pm$0.22& 1.35$\pm$0.27 & 1.02$\pm$0.20 & 0.19$\pm$0.08 & --- & --- & --- & --- & --- \\
4861.33 & {H\,{\sc i}} & H$\beta$ & 100$\pm$2 & 100$\pm$2 & 100$\pm$2 & 100$\pm$2 & 100$\pm$2 & 100$\pm$2 & 100$\pm$2 & 100$\pm$2 & 100$\pm$2 \\
4958.91 & {[O~{\sc iii}]} & 1F & 2.52$\pm$0.50& 0.98$\pm$0.29 & 0.15$\pm$0.06 & --- & --- & --- & 0.93$\pm$0.28 & 0.88$\pm$0.26 & 7.59$\pm$0.76 \\
5006.94 & {[O~{\sc iii}]} & 1F & 7.86$\pm$0.79& 2.78$\pm$0.56 & 0.25$\pm$0.10 & --- & 0.35$\pm$0.14 & 0.13$\pm$0.05 & 2.78$\pm$0.56 & 3.06$\pm$0.61 & 25.42$\pm$1.27 \\
6300.30 & {[O~{\sc i}]} & 1F & 0.64$\pm$0.03& 0.64$\pm$0.03 & 0.60$\pm$0.03 & 0.86$\pm$0.04 & 1.05$\pm$0.05 & 1.30$\pm$0.06 & 2.47$\pm$0.12 & 3.50$\pm$0.17 & 5.39$\pm$0.27 \\
6312.10 & {[S~{\sc iii}]} & 3F & 0.80$\pm$0.04& 0.70$\pm$0.03 & 0.69$\pm$0.03 & 0.58$\pm$0.03 & 0.47$\pm$0.02 & 0.42$\pm$0.02 & 0.45$\pm$0.02 & 0.38$\pm$0.02 & 0.36$\pm$0.02 \\
6548.03 & {[N~{\sc ii}]} & 1F & 36.31$\pm$1.82& 38.4$\pm$1.9 & 41.3$\pm$2.1 & 41.8$\pm$2.1 & 42.0$\pm$2.1 & 44.2$\pm$2.2 & 45.7$\pm$2.3 & 50.2$\pm$1.0 & 54.1$\pm$1.1 \\
6562.82 & {H\,{\sc i}} & H$\alpha$ & 292.00$\pm$5.84& 292$\pm$6 & 292$\pm$6 & 292$\pm$6 & 292$\pm$6 & 292$\pm$6 & 292$\pm$6 & 292$\pm$6 & 292$\pm$6 \\
6583.41 & {[N~{\sc ii}]} & 1F & 107.12$\pm$2.14& 113$\pm$2 & 122$\pm$2 & 123$\pm$2 & 124$\pm$2 & 130$\pm$3 & 135$\pm$3 & 148$\pm$3 & 159$\pm$3 \\
6678.15 & {He~{\sc i}} & 46 & 1.11$\pm$0.06& 1.13$\pm$0.06 & 0.89$\pm$0.04 & 0.20$\pm$0.01 & --- & --- & --- & 0.03$\pm$0.01 & 0.26$\pm$0.01 \\
6716.47 & {[S~{\sc ii}]} & 2F & 18.79$\pm$0.94& 20.2$\pm$1.0 & 22.2$\pm$1.1 & 25.8$\pm$1.3 & 29.4$\pm$1.5 & 36.9$\pm$1.8 & 42.3$\pm$2.1 & 49.1$\pm$2.5 & 58.8$\pm$1.2 \\
6730.85 & {[S~{\sc ii}]} & 2F & 20.02$\pm$1.00& 21.4$\pm$1.1 & 23.0$\pm$1.2 & 26.6$\pm$1.3 & 30.4$\pm$1.5 & 38.7$\pm$1.9 & 42.8$\pm$2.1 & 53.3$\pm$1.1 & 60.8$\pm$1.2 \\
7065.28 & {He~{\sc i}} & 10 & 0.92$\pm$0.05& 0.85$\pm$0.04 & 0.65$\pm$0.03 & 0.15$\pm$0.01 & --- & --- & 0.06$\pm$0.01 & 0.09$\pm$0.01 & 0.34$\pm$0.02 \\
7135.78 & {[Ar~{\sc iii}]} & 1F & 2.90$\pm$0.14& 2.66$\pm$0.13 & 1.70$\pm$0.08 & 0.31$\pm$0.02 & 0.03$\pm$0.01 & 0.03$\pm$0.01 & 0.19$\pm$0.01 & 0.24$\pm$0.01 & 1.16$\pm$0.06 \\
7319.19 & {[O~{\sc ii}]} & 2F & 1.74$\pm$0.09& 1.67$\pm$0.08 & 1.67$\pm$0.08 & 1.51$\pm$0.08 & 1.52$\pm$0.08 & 1.66$\pm$0.08 & 1.80$\pm$0.09 & 2.53$\pm$0.13 & 2.67$\pm$0.13 \\
7330.20 & {[O~{\sc ii}]} & 2F & 1.46$\pm$0.07& 1.41$\pm$0.07 & 1.34$\pm$0.07 & 1.24$\pm$0.06 & 1.26$\pm$0.06 & 1.39$\pm$0.07 & 1.50$\pm$0.08 & 2.07$\pm$0.10 & 2.20$\pm$0.11 \\
\noalign{\vskip3pt} \noalign{\hrule} \noalign{\vskip3pt}
\multicolumn{3}{r}{$c$(H$\beta$)} & 0.60$\pm$0.09 & 0.49$\pm$0.08 & 0.51$\pm$0.05 & 0.75$\pm$0.06 & 0.76$\pm$0.06 & 0.79$\pm$0.07 & 0.77$\pm$0.06 & 0.73$\pm$0.06 & 0.99$\pm$0.07 \\
\multicolumn{3}{r}{$F$(H$\beta$)} & 1.508$\times$10$^{-12}$ & 1.443$\times$10$^{-12}$ & 1.185$\times$10$^{-12}$ & 1.861$\times$10$^{-12}$ & 1.770$\times$10$^{-12}$ & 1.765$\times$10$^{-12}$ & 1.271$\times$10$^{-12}$ & 1.150$\times$10$^{-12}$ & 7.131$\times$10$^{-13}$ \\
\noalign{\vskip3pt}
\multicolumn{3}{r}{$n_{\rm e}$([O~{\sc ii}])}& 560$\pm$50 & 520$\pm$50 & 500$\pm$40 & 440$\pm$40 & 450$\pm$40 & 450$\pm$40 & 420$\pm$40 & 510$\pm$50 & 460$\pm$40 \\
\multicolumn{3}{r}{$n_{\rm e}$([S~{\sc ii}])}& 650$\pm$170 & 630$\pm$170 & 570$\pm$160 & 560$\pm$150 & 560$\pm$150 & 600$\pm$160 & 520$\pm$150 & 690$\pm$140 & 580$\pm$60 \\
\multicolumn{3}{r}{$T_{\rm e}$([O~{\sc ii}])}& 7850$\pm$160 & 7600$\pm$150 & 7360$\pm$140 & 7260$\pm$130 & 7270$\pm$140 & 7420$\pm$140 & 7440$\pm$140 & 8070$\pm$180 & 7810$\pm$180 \\
\multicolumn{3}{r}{$T_{\rm e}$([O~{\sc iii}])$^{(3)}$} & 7360$\pm$1700 & 7000$\pm$1650 & 6660$\pm$1600 & --- & --- & --- & --- & --- & --- \\
\noalign{\vskip3pt}
\multicolumn{3}{r}{He$^+$/H$^+$} &10.44$\pm$0.02 &10.45$\pm$0.03 &10.34$\pm$0.03 & --- & --- & --- & --- & --- & --- \\
\multicolumn{3}{r}{O$^+$/H$^+$} & 8.34$\pm$0.05 & 8.45$\pm$0.05 & 8.55$\pm$0.05 & 8.58$\pm$0.05 & 8.58$\pm$0.05 & 8.54$\pm$0.05 & 8.57$\pm$0.05 & 8.42$\pm$0.05 & 8.56$\pm$0.06 \\
\multicolumn{3}{r}{O$^{2+}$/H$^+$} & 6.97$\pm$0.20 & 6.64$\pm$0.22 & 5.83$\pm$0.26 & --- & --- & --- & --- & --- & --- \\
\multicolumn{3}{r}{N$^+$/H$^+$} & 7.63$\pm$0.04 & 7.70$\pm$0.04 & 7.78$\pm$0.04 & 7.80$\pm$0.04 & 7.80$\pm$0.04 & 7.80$\pm$0.04 & 7.81$\pm$0.04 & 7.73$\pm$0.04 & 7.81$\pm$0.04 \\
\multicolumn{3}{r}{S$^+$/H$^+$} & 6.28$\pm$0.04 & 6.35$\pm$0.04 & 6.43$\pm$0.04 & 6.51$\pm$0.04 & 6.56$\pm$0.04 & 6.64$\pm$0.04 & 6.68$\pm$0.04 & 6.66$\pm$0.04 & 6.77$\pm$0.04 \\
\multicolumn{3}{r}{S$^{2+}$/H$^+$} & 6.78$\pm$0.04 & 6.80$\pm$0.04 & 6.87$\pm$0.04 & 6.83$\pm$0.04 & 6.73$\pm$0.04 & 6.64$\pm$0.04 & 6.66$\pm$0.04 & 6.39$\pm$0.05 & 6.44$\pm$0.05 \\
\multicolumn{3}{r}{Ar$^{2+}$/H$^+$} & 5.79$\pm$0.16 & 5.82$\pm$0.17 & 5.69$\pm$0.17 & --- & --- & --- & --- & --- & --- \\
\noalign{\vskip3pt}
\multicolumn{3}{r}{O/H} & 8.36$\pm$0.05 & 8.45$\pm$0.05 & 8.55$\pm$0.05 & 8.58$\pm$0.05 & 8.58$\pm$0.05 & 8.54$\pm$0.05 & 8.57$\pm$0.05 & 8.42$\pm$0.05 & 8.56$\pm$0.06 \\
\multicolumn{3}{r}{N/H} & 7.65$\pm$0.08 & 7.71$\pm$0.08 & 7.78$\pm$0.08 & 7.80$\pm$0.04 & 7.80$\pm$0.04 & 7.80$\pm$0.04 & 7.81$\pm$0.04 & 7.73$\pm$0.04 & 7.81$\pm$0.04 \\
\multicolumn{3}{r}{S/H} & 6.90$\pm$0.04 & 6.93$\pm$0.03 & 7.01$\pm$0.03 & 7.00$\pm$0.03 & 6.96$\pm$0.03 & 6.94$\pm$0.03 & 6.97$\pm$0.03 & 6.85$\pm$0.03 & 6.94$\pm$0.03 \\
\noalign{\vskip3pt} \noalign{\hrule} \noalign{\vskip3pt}
\end{tabular}
\begin{description}
\scriptsize
\item[] $^{(1)}$ The errors in the line fluxes refer only to uncertainties in the line measurements (see text). $F$(H$\beta$) in erg\,cm$^{-2}$s$^{-1}$;
$n_{\rm e}$\ in cm$^{-3}$; $T_{\rm e}$\ in K; ionic abundances in log\,(X$^{+i}$/H$^+$)\,+\,12.
\item[] $^{(2)}$ Blended with {He~{\sc i}} $\lambda$3833.57 line.
\item[] $^{(3)}$ $T_{\rm e}$([O~{\sc iii}]) obtained using the empirical relation between $T_{\rm e}$([O~{\sc ii}]) and $T_{\rm e}$([O~{\sc iii}]) obtained from the data by \citet{Gar07}
(see text).
\end{description}
\end{table*}
\subsection{Extinction correction}
\label{section43}
We assume the extinction law derived by \cite{Bla07} for
the Orion nebula. Both \hii\ regions have a deficiency in small particles that produces the relatively
grey extinction observed in Orion \citep{Mag86, Bal91}. Furthermore,
\citet{Rod99, Rod02} did not find large extinction variations in her optical spectroscopic studies
of the M\,42 and M\,43 \hii\ regions.
The reddening coefficient, $c$({{H$\beta$}}) was obtained by fitting the observed
H$\delta$/{{H$\beta$}} and H$\gamma$/{{H$\beta$}} line intensity ratios --\,the three lines lie in
the same spectral range\,-- to the theoretical ones computed by \cite{Sto95} for
{$T_{\rm e}$}\,=\,7500 K and {$n_{\rm e}$}\,=\,500 cm$^{-3}$. As pointed out by
\cite{Mes09a, Mes09b}, the use of a \cite{Bla07} law instead of the classical one
by \cite{Cos70} produces slightly higher
$c$({{H$\beta$}}) values and also slightly different dereddened line fluxes depending on
the spectral range.
The final set of line intensities corrected for both the extended emission component
and the extinction are indicated in Table~\ref{t2}.
\subsection{Uncertainties}
\label{section44}
Several sources of uncertainties must be taken into account to obtain
the errors associated with the line intensity ratios. We estimated that
the uncertainty\footnote{Although uncertainties in the line flux measurements depend
on the line flux instead of on $F$($\lambda$)/$F$({{H$\beta$}}), the similarity between
the {{H$\beta$}} flux measured in each aperture (see Table~\ref{t2}) makes this approach
reasonable.} in the line intensity measurement due to the signal-to-noise ratio
of the spectra and the placement of the local continuum is typically
\begin{itemize}
\item [$\sim$]\,2\% for $F$($\lambda$)/$F$({{H$\beta$}})$\geq$0.5,
\item [$\sim$]\,5\% for 0.1$\leq$ $F$($\lambda$)/$F$({{H$\beta$}}) $\leq$0.5,
\item [$\sim$]\,10\% for 0.05$\leq$ $F$($\lambda$)/$F$({{H$\beta$}}) $\leq$0.1,
\item [$\sim$]\,20\% for 0.01$\leq$ $F$($\lambda$)/$F$({{H$\beta$}})$\leq$0.05,
\item [$\sim$]\,30\% for 0.005$\leq$ $F$($\lambda$)/$F$({{H$\beta$}})$\leq$0.01, and
\item [$\sim$]\,40\% for 0.001$\leq$ $F$($\lambda$)/$F$({{H$\beta$}})$\leq$0.005.
\end{itemize}
We did not consider any lines of weaker intensity than 0.001 $\times$ $F$({{H$\beta$}}).
We note that then uncertainties indicated in Table \ref{t2} refer inly to this type
of errors.
By comparing the resulting flux-calibrated spectra of one of our standard stars
with the corresponding tabulated flux, we could estimate that the line ratio uncertainties
associated with the flux calibration is
$\sim$\,3\% when the wavelengths are separated by 500\,--\,1500 \AA\ and $\sim$\,5\%
if they are separated by more than that. Where
the corresponding lines are separated by less than 500 \AA, the uncertainty
in the line ratio due to uncertainties in the flux calibration is negligible.
The uncertainty associated with extinction correction was computed by error propagation.
The contribution of this uncertainty in the total error is again negligible when
line ratios of close-by lines are considered
(e.g. [\ion{S}{ii}]\,$\lambda$6716/[\ion{S}{ii}]\,$\lambda$6730).
The final errors in the line intensity ratios used to derive the physical properties
of the nebula were computed by quadratically adding these three sources of uncertainty.
\subsection{Physical conditions}
\label{section45}
The electron temperature ($T_{\rm e}$) and density ($n_{\rm e}$) of the ionized gas were
derived from classical CEL ratios, using the \iraf\ task
\emph{temden} of the package \emph{nebular} \citep{Sha95} with updated
atomic data for several ions \cite[see][]{Gar09}. We computed {$n_{\rm e}$} from
the {[S~{\sc ii}]} $\lambda$6717/$\lambda$6731 and {[O~{\sc ii}]} $\lambda$3729/$\lambda$3726
line ratios and {$T_{\rm e}$} from the nebular to auroral
{[O~{\sc ii}]} $\lambda\lambda$(7320+30)/$\lambda\lambda$(3726+29) line ratio\footnote{In
the case of M\,43, $T_{\rm e}$({[O~{\sc ii}]}) is free of any contamination to
{[O~{\sc ii}]} $\lambda$7325 by recombination because of the low \ion{O}{$^{2+}$} abundance.}.
We could not directly determine {$T_{\rm e}$}({[O~{\sc iii}]}) because
the {[O~{\sc iii}]} $\lambda$4363 auroral line was not detectedbecause of the low ionization degree
of the nebula. As an alternative, we used a empirical relation between {$T_{\rm e}$}({[O~{\sc iii}]}) and
{$T_{\rm e}$}({[O~{\sc ii}]}) obtained from the nebular information published by \citet{Gar07} for a sample of
Galactic {\hii} regions\footnote{The data for NGC3603, a giant Galactic {\hii} region, was discarded
to obtain this relation.}
\begin{equation}
\label{te_rel}
T_e({\rm [O\,III]})=[T_e({\rm [O\,II]})-(2640\pm1270)]/(0.7105\pm0.1524).
\end{equation}
This fit is almost identical to that obtained by \citet{Pil06} and \citet{Gar92}
for integrated spectra of giant extragalactic {\hii} regions and
photoionization models of giant extragalactic {\hii} regions, respectively.
We only computed {$T_{\rm e}$}({[O~{\sc iii}]}) for the three innermost apertures. For more
distant apertures, the O$^{2+}$ emission (if any) is not associated to
M\,43 itself, but to the residual emission of the diffuse component (see Sects. \ref{section3}
and \ref{section41}).
The methodology used to determine the physical conditions was as follows:
we assumed a representative initial value of {$T_{\rm e}$} of 10000 K and computed the
electron densities. The value of {$n_{\rm e}$} was then used to compute
{$T_{\rm e}$}({[O~{\sc ii}]}) from the observed line ratios, and {$T_{\rm e}$}({[O~{\sc iii}]}) from equation 2.
We iterated until convergence to obtain the final values of {$n_{\rm e}$} and {$T_{\rm e}$}.
Uncertainties were computed by error propagation. The final {$n_{\rm e}$}({[O~{\sc ii}]}),
{$n_{\rm e}$}({[S~{\sc ii}]}), {$T_{\rm e}$}({[O~{\sc ii}]}), {$T_{\rm e}$}({[O~{\sc iii}]}) estimate, along with their uncertainties
are indicated in Table~\ref{t2}.
In general, densities derived from the {[O~{\sc ii}]} line ratio are about 100 cm$^{-3}$
lower than those derived from {[S~{\sc ii}]} lines, but consistent within the
errors.
\subsection{Chemical abundances}
\label{section46}
To derive He$^+$/H$^+$, we used two observed lines of {He~{\sc i}} at $\lambda\lambda$4471 and 6678.
Case B emissivities were taken from the collision less (low-density limit) calculations by
\citet{Bau05} using an on line available code\footnote{Available at \tt{http://www.pauKy.edu/$\sim$rporter/j-resolved}}.
The collisional-to-recombination contribution was estimated from \citet{Kin95}, using the interpolation
formula provided by \citet{Por07}. The effective recombination coefficients for H$^+$ were taken
from \citet{Sto95}
Ionic abundances of N$^+$, O$^+$, O$^{2+}$, S$^+$, S$^{2+}$ and Ar$^{2+}$ were
derived from CELs, using the \iraf\ task $ionic$ of the package $nebular$. We
assumed a two-zone scheme, in which we adopted {$T_{\rm e}$}({[O~{\sc ii}]}) for the low
ionization potential ions (N$^+$, O$^+$, S$^+$ and S$^{2+}$) and the value of {$T_{\rm e}$}({[O~{\sc iii}]}) derived
from Eq.~\ref{te_rel} for the high ionization potential ions (O$^{2+}$ and Ar$^{2+}$).
In both cases, we assumed {$n_{\rm e}$}({[O~{\sc ii}]}) instead of {$n_{\rm e}$}({[S~{\sc ii}]}) because the first indicator
is more representative of the whole nebula, while {$n_{\rm e}$}({[S~{\sc ii}]}) provides insight mostly
into the density of the material close to the ionization edge of the nebula.
For O$^{2+}$, Ar$^{2+}$, and He$^+$, we only have available lines in the three innermost apertures.
Owing to the low \Teff\ of the ionizing star, these relatively high ionization species
are only present in regions very close to the star.
We then derived the total abundances of O, N, and S for each aperture\footnote{Owing to the
low ionization degree of M\,43, the ionization correction factors for He$^+$ and Ar$^{2+}$ are
large and very uncertain. We did not
compute the total abundances for these two elements.}.
As a consequence of the low ionization degree of the nebula, M\,43 has a very small amount of O$^{2+}$
in comparison with the dominant species, O$^{+}$, except in the innermost zones where a non-negligible fraction of
O$^{2+}$ is present. We note that for the intermediate apertures (apertures 4 to 6),
the O abundance is given by O$^+$/H$^+$. In the case of S, the abundance of this element is given by the sum of
S$^+$ and S$^{2+}$, because S$^{+3}$ is not expected to be present in this nebula. Similarly to
O$^{2+}$, N$^{2+}$ is only expected to be present in the innermost parts of the nebula. For the corresponding
apertures (1\,--\,3), we applied the classical ionization correction factor (ICF) scheme of N (i.e. N$^+$/O$^+$=N/O) to correct for the
N$^{2+}$ abundance (note that this correction is very small). As for O, for the intermediate
apertures the N abundance is directly obtained from the N$^+$ abundance.
The ionic abundances of He$^{+}$ O$^{+}$, O$^{2+}$, N$^{+}$, S$^{+}$, S$^{2+}$, and Ar$^{2+}$, along
with the total abundances of O, N, and S are shown in Table~\ref{t2}. We stress again that
in this nebula the total abundances of O, N, and S can be obtained directly from observable ions
(no ICFs are needed) from the intermediate apertures.
\subsection{Comparison of O, N, and S abundances derived in M\,43 and the Orion nebula}
\label{section47}
Given that the ionized gas from both nebulae, M42 and M43, comes from the
same molecular cloud, it is reasonable to assume that both have very similar
total abundances. In this section, we compare the O, S, and N abundances obtained
from our study of M\,43 with those determined by \citet[][GRE07, an update of Esteban et al. 2004]{Gar07}
and \citet[][SDS10]{Sim10b}
from the analysis of the nebular spectrum of the Orion nebula. Table \ref{t3}
summarizes the final abundances indicated in the three studies. We note that
GRE07, and SDS10 present analyses of the same spectroscopic
dataset, but use different atomic data and ICFs. Our study of M\,43 uses the same
atomic data as that considered by GRE07.
\subsubsection{Oxygen}
\label{section471}
For this element, where the total abundance is obtained directly from observable
ions (\ion{O}{$^+$} and \ion{O}{$^+$}+\ion{O}{$^{2+}$} in the case of M\,43
and M\,42, respectively), good agreement is found within the errors.
\subsubsection{Nitrogen}
\label{section472}
\begin{table}[t!]
\begin{center}
\caption{\footnotesize Summary of O, N, and S abundances (from CEL) resulting from
the spectroscopic analysis of M\,43 (this work) and the Orion nebula (GRE07, SDS10).
For M\,43, we assumed the mean value of abundances derived from apertures 4\,--\,6 (the
uncertainty given by the values indicated in Table \ref{t2} for each aperture).
\label{t3}
}
\begin{tabular}{cccc}
\hline
\hline
\noalign{\vskip2pt}
Element & M\,43 & M\,42 (GRE07) & M\,42 (SDS10) \\
\noalign{\vskip1pt}
\hline
O & 8.57$\pm$0.05 & 8.54$\pm$0.03 & 8.52$\pm$0.01 \\
S & 6.97$\pm$0.03 & 7.04$\pm$0.04 & 6.87$\pm$0.04 \\
N & 7.80$\pm$0.04 & 7.73$\pm$0.09 & 7.90$\pm$0.09 \\
\hline
\hline
\end{tabular}
\\
\end{center}
\end{table}
For this element, GRE07 and SDS10 evaluate different values.
This discrepancy is caused by the assumed ICF(N$^+$). While GRE07
obtained this ICF from the widely used empirical relation N$^+$/O$^+$=N/O,
SDS10 determined the ICF from a model fitted to the data of M\,42
\citep[including a detailed description of the ionizing SED of $\Theta^1$\,Ori\,C
predicted by the stellar atmosphere code WM-basic, ][]{Pau01}.
The GRE07 and SDS10 abundances differ by 0.17 dex. Interestingly, our determination for
M\,43, in which no ICF is needed produces an intermediate value, and is in
perfect agreement with the B-type stars abundance derived by \citet{Nie11}.
This introduces a new important constraint on the nebular CEL/RL abundance
conundrum \citep{Sim10b}.
\subsubsection{Sulphur}
\label{section472}
GRE07 and SDS10 derive different
values (0.17 dex difference) of the total S abundance in M\,42. In
this case, the cause of the difference is not, however, the assumed ICF, but the use
of different sets of atomic data. Both computations considered the same
data for the collision strength \citep{Kee96, Tay99}, but different transition
probabilities: while GRE07 based their calculation in data from \cite{Men82a, Men82b}
for [\ion{S}{ii}] and [\ion{S}{iii}], respectively, SDS10 used \cite{Fro04a} and \cite{Fro04b}
values. For M\,43, we obtained 12\,+\,log(S/H)=6.97$\pm$0.04 using the
same set of atomic data as GRE07. This value is slightly lower, but in agreement within
the errors, than the determination by GRE07, and somewhat
higher than the value provided by SDS10. The influence
of the atomic data considered for this element is hence relevant.
\subsection{Comparison with previous studies of M\,43}
\label{section48}
The only nebular abundance studies found in the literature
for M\,43 are those of \citet{Rod99, Rod02}, who
analyzed long-slit optical spectroscopic data for five different slit
positions along the nebula, deriving the physical conditions ($T_{\rm e}$\
and $n_{\rm e}$), and O, S, Cl, N, Ar, He, C, and Fe abundances of the nebular
gas. \citet{Ode10} analyzed a set of optical spectrophotometric
observations of M\,43 as part of a more general study of the EON, and derived the
physical conditions of the nebula. We note that none of these observations were corrected
for the diffuse emission component as in our study.
For consistency, we reanalyzed the data of \citet{Rod99} using the same atomic data as
in our work. We obtained {$T_{\rm e}$}({[N~{\sc ii}]}) between 7800 K and 8000 K for slit positions similar
to ours. \citet{Ode10} obtained very similar results ({$T_{\rm e}$}({[N~{\sc ii}]}) between 7780 K and 7940 K)
from several aperture extractions of a slit that crosses the nebula in the east-west direction.
Both authors derived $n_{\rm e}$\ ({[S~{\sc ii}]}) $\sim$500-600 cm$^{-3}$, which is very similar to the
value derived in this work.
We compared our abundances with those resulting from the reanalysis\footnote{The observations by
\citet{Ode10} do not include {[O~{\sc ii}]}$\lambda\lambda$3727, 7325 lines,
hence the determination of the oxygen abundance is not possible.} of the observational
dataset by \citet{Rod99}. The derived O abundances are 0.14\,--\,0.23 dex lower
than those resulting from the analysis of our spectra. As we indicate in Sect. \ref{section49},
this effect is produced by the contamination of the M\,43 emission by the diffuse component.
Finally, we compared our estimated {$T_{\rm e}$}({[O~{\sc ii}]}) with previous determinations
from radio continuum observations. \cite{Sub92} derived
$T_{\rm e}$=\solu{9000}{1700} K by combining 330 MHz and 10.7 GHz continuum observations and a
model of an isothermal \hii\ region; \cite{Mil68} estimated a value of $T_{\rm e}$$\sim$8000 K from 408 MHz
continuum observations; finally, \cite{Thu78} derived a much lower $T_{\rm e}$$\sim$6700 K
from the HI 91$\alpha$ radio recombination line to continuum ratio. In general, the different estimates,
except for Thum et al., agree within the uncertainties.
\begin{figure}[!t]
\centering
\includegraphics[width=8.5cm,angle=0]
{16608_fig9.eps}
\caption{Upper panel: Difference between the non-extinction-corrected H$\gamma$/{{H$\beta$}} and H$\delta$/{{H$\beta$}} ratios from
the C and NC spectra (see text for explanation).
Lower panel: {He~{\sc i}} 7065/6678 line ratio from the NC spectra (9 apertures). The mean and standard
deviation values corresponding to the bright Huygens region \citep[obtained from data by][]{Ode10}
are also indicated as a dashed and dotted lines, respectively.}
\label{f8}
\end{figure}
\begin{figure*}[!t]
\centering
\includegraphics[angle=90, width=18cm]
{16608_fig10.eps}
\caption{Observed line ratios obtained from the corrected (black diamonds) and
non corrected (red squares) spectra from extended emission component.}
\label{f9}
\end{figure*}
\subsection{Scattered nebular light from the EON affecting the spectrum of M\,43?}
\label{section49}
\citet{Ode09} and \citet{Ode10} demonstrated that scattered light from
the bright Huygens Region of M\,42 can affect the physical conditions derived from emission
lines in the EON. \citet{Ode10} performed a detailed spectroscopic study of several zones
of the EON, including the M\,43 region. These authors could not find conclusive
evidence of signatures of scattered light from the Huygens region affecting the nebular
emission from M\,43 (except for the case of an aperture located in the M\,43
dark lane). Our study (based on more detailed observational dataset) allowed us
to observationally prove that there is an extended diffuse emission not produced by
ionization of the surrounding material by NU Ori affecting the whole nebula.
To investigate whether this extended emission relates to scattered light from the
Huygens region, we compare in Fig.~\ref{f8} (upper panel) the
non-extinction-corrected H$\gamma$/{{H$\beta$}} and H$\delta$/{{H$\beta$}} line ratios with (C) and without (NC) correction
for diffuse emission for apertures from A1 to A9. The H$\delta$/{{H$\beta$}}
line ratio is more affected than
H$\gamma$/{{H$\beta$}}. This is consistent with the diffuse extended emission being produced by
dust scattered light, a well-know property of this type of emission is that blue
lines are more affected than red lines. An additional argument supporting this hypothesis
is presented in the bottom panel of Fig.~\ref{f8}. The value of the
{He~{\sc i}} 7065/6678 ratio, obtained from the NC spectra of M\,43 outside the \ion{He}{$^+$}
Str\"omgen sphere associated with NU Ori
agrees with the associated average value in the Huygens region taken from the data of
\citet{Ode10}\footnote{We considered extractions corresponding to slits 2 to 15 that are
labeled as ''inner''.}.
We also investigated the effect that scattered light contaminating the spectrum of
M\,43 has on line ratios used to determine the physical conditions and chemical
abundances (as well as constraints of the photoionization models presented in
Paper~II). Fig.~\ref{f9} shows several line ratios obtained from the
C and NC spectra. Although the effect is negligible for the density constraints
and the \ion{S}{$^+$} and \ion{O}{$^+$} abundance indicators, the influence of
scattered light on the ionization structure constraints and the \ion{He}{$^+$},
\ion{N}{$^+$}, \ion{Ar}{$^2+$}, \ion{S}{$^2+$}, and \ion{O}{$^{2+}$} abundance indicators
is not negligible (especially in the outer parts of the nebula). In addition, as already pointed out by
\citet{Ode10}, our comparison identifies an important effect on $T_{\rm e}$([\ion{O}{ii}]).
If the scattered light contribution is not removed, the spectroscopic analysis
overestimates the $T_{\rm e}$\ and hence underestimates of the abundances.
To quantitatively illustrate this effect on the computed physical conditions and
abundances, we present in Table \ref{t4} the results of two empirical analyses of
the C and NC spectra from apertures 4 and 5. We also compare the result of the
analysis of the NC integrated spectrum (extracting the global long\,slit spectra).
As pointed out before, $n_{\rm e}$\ is basically similar but the NC spectra result in a
higher $T_{\rm e}$\ and, consequently lower abundances. In particular, ionic abundances can
be affected by $\sim$\,0.05\,--\,0.15 dex.
An adequate correction of the scattered light contribution in the spectra of M\,43 is hence
important for a reliable empirical analysis of the nebular spectra of
this \hii\ region.
\begin{table}[t!]
\begin{center}
\caption{\footnotesize Summary of physical conditions, ionic and total abundances derived
for apertures 4 and 5 with and without correcting for
extended emission component and for the longslit extraction in M\,43.$^{(1)}$}
\label{t4}
\begin{tabular}{l l c c c c c}
\hline
& & \multicolumn{2}{c}{Uncorrected} & \multicolumn{2}{c}{Corrected} & Uncorrected \\
\cline{3-6}
& & A4 & A5 & A4 & A5 & Total Slit \\
\cline{3-7}
$n_{\rm e}$ (cm$^{-3}$) & {[O~{\sc ii}]} & 475 & 485 & 440 & 450 & 510 \\
& {[S~{\sc ii}]} & 560 & 565 & 560 & 560 & 600 \\
\hline
$T_{\rm e}$ (K) & {[O~{\sc ii}]} & 7650 & 7700 & 7260 & 7270 & 7900 \\
& {[O~{\sc iii}]} & --- & --- & --- & --- & 7450 \\
\hline
$\epsilon$(X$^{+i}$) & O$^+$ & 8.45 & 8.44 & 8.58 & 8.58 & 8.38 \\
& O$^{++}$ & --- & --- & --- & --- & 7.45 \\
& N$^{+}$ & 7.71 & 7.70 & 7.80 & 7.80 & 7.66 \\
& S$^{+}$ & 6.42 & 6.46 & 6.51 & 6.56 & 6.42 \\
& S$^{++}$ & 6.77 & 6.70 & 6.83 & 6.73 & 6.69 \\
& Ar$^{++}$ & --- & --- & --- & --- & 5.64 \\
\hline
$\epsilon$(X) & O & 8.50 & 8.49 & 8.58 & 8.58 & 8.43 \\
& N & 7.75 & 7.75 & 7.80 & 7.80 & 7.71 \\
& S & 6.93 & 6.90 & 7.00 & 6.96 & 6.88 \\
\hline
\end{tabular}
\begin{description}
\scriptsize
\item[] $^{(1)}$ $T_{\rm e}$, ionic, and total abundance values are based on the assumption
$n_{\rm e}$=$n_{\rm e}$({[O~{\sc ii}]}) as discussed in Sect. \ref{section46}.
\end{description}
\end{center}
\end{table}
\section{Summary and conclusions}
\label{section7}
M\,43 is a close-by, bright \hii\ region of simple
geometry, ionized by a single star. These characteristics makes M\,43 an
ideal object for investigating several topics of interest in the field
of \hii\ regions and massive stars.
In a series of two papers, we present a combined, comprehensive study of
the nebula and its ionizing star using as many observational constraints
as possible. In this first part of the study, we have introduced the observational
dataset, obtained the stellar parameters of HD\,37061, obtained useful information
from the nebular images, and analyzed the nebular spectra
extracted from apertures located at various distances from the central star.
All this information is then used in Paper II to construct a customized
photoionization model of the nebula. \\
In this paper, we have found observational evidence of a diffuse and extended emission
in the region where M\,43 is located that is not associated to material ionized
by HD\,37061. Our nebular observations have allowed us to ascertain the most likely
origin of this emission \citep[e.g.][and references therein]{Ode09, Ode10}, namely
that light emitted in the Huygens region (the central, brightest part of the Orion nebula)
is scattered by dust. We have also shown the importance of an adequate correction
of this scattered light from the imagery and spectroscopic observations of M\,43
to a proper determination of the total nebular {H$\alpha$}\ luminosity, the nebular physical
conditions, and chemical abundances. In particular, we illustrated that an overestimate of
$T_{\rm e}$\ by $\sim$\,400\,--\,500 K, hence an underestimate of abundances by
0.05\,--\,0.10 dex, result from the empirical analysis of our spectroscopic dataset
when the spectra are not corrected for the scattered light contribution. In
addition, the derived total nebular {H$\alpha$}\ luminosity may be overestimated by a factor
$\sim$\,1.5.
The quantitative analysis of the optical spectrum of HD\,37061 with the
stellar atmosphere code FASTWIND lead to \Teff\,=\,31000$\pm$500 K, \grav\,=\,4.2$\pm$0.1, and
log\,Q(H$^0$)\,=\,47.2\,$\pm$\,0.2 (assuming a distance of 400\,pc).
The analysis of the {H$\alpha$}\ and {H$\beta$}\ images indicate a non-constant extinction distribution
within the nebula that is well correlated with the dust features indicated by
\cite{Smi87}. Once the {H$\alpha$}\ image is corrected for extinction and diffuse emission,
a total nebular {H$\alpha$}\ luminosity of (3.0$\pm$1.1) x 10$^{35}$ erg\,s$^{-1}$ is obtained.
This value is compatible with the ionizing flux of the star, implying that the nebula is
(mostly) ionization bounded.
We extracted nine apertures from a long-slit located to the west of HD\,37061 (east-west direction)
to obtain the spatial distribution of the nebular physical conditions (temperature and density)
and ionic abundances (\ion{He}{$^+$}, \ion{O}{$^+$}, \ion{O}{$^{2+}$}, \ion{N}{$^+$},
\ion{S}{$^+$}, \ion{S}{$^{2+}$}, \ion{Ar}{$^{2+}$}). Since it is important to correct
these spectra for the contribution of scattered light from the Huygens region, we also consider
two apertures outside the nebula that are used for this aim.
For sulfur, oxygen, and nitrogen, we have been able to determied total abundances directly
from observable ions (no ICFs are needed). The derived abundances, 8.57\,$\pm$\,0.05,
6.97\,$\pm$\,0.03, and 7.80\,$\pm$\,0.04, respectively, are compared with previous
determinations in the Orion nebula. Although an overall agreement is found, our study illustrates
the importance of the atomic data and, specially in the case of M\,42, the considered ICFs.
\begin{acknowledgements}
SSD acknowledges the funds by the Spanish Ministerio de Educaci\'on y Ciencia
under the MEC/Fullbright postdoctoral fellowship program.
SSD, JGR, CE, and ARLS acknowledge finantial support by the Spanish MICINN under
projects AYA2008-06166-C03-01 and AYA2007-63030. SSD also acknowledges financial
support from UNAM/DGAPA PAPIIT IN 112708 (IP: M. Pe\~na) and CONACyT J49737
(IP: C. Morisset) and the members of the Instituto de Astronom{\'{\i}}a,
UNAM for their warm hospitality. JGR acknowledges the support of an UNAM postdoctoral grant.
Part of this work was done while CM in sabbatical at IAC founded by CONACyT and PASPA-UNAM grants.
This work has also been partially funded by the Spanish MICINN under the
Consolider-Ingenio 2010 Program grant CSD2006-00070: First Science with the GTC.
\end{acknowledgements}
|
\section{Introduction}
In this section, we present some definitions and known results
which are necessary to prove our main theorems. Throughout this
paper, $G=(V,E)$ is a finite simple graph. We use $\overline G$
for the complement of graph $G$. The distance between two
vertices $u$ and $v$, denoted by $d_G(u,v)$, is the length of a
shortest path between $u$ and $v$ in $G$. Also, $N_G(v)$ is the
set of all neighbors of vertex $v$ in $G$. We write these simply
$d(u,v)$ and $N(v)$, when no confusion can arise. The notations
$u\sim v$ and $u\nsim v$ denote the adjacency and none-adjacency
relation between $u$ and $v$, respectively. The symbols
$(v_1,v_2,\ldots, v_n)$ and $(v_1,v_2,\ldots,v_n,v_1)$ represent
a path of order $n$, $P_n$, and a cycle of order $n$, $C_n$,
respectively.
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. In this
case, we say set $W$ resolves $G$. Elements in a resolving set
are called {\it landmarks}. A resolving set $W$ for $G$ with
minimum cardinality is called a {\it basis} of $G$, and its
cardinality is the {\it metric dimension} of $G$, denoted by
$\beta(G)$. The concept of (metric) representation is introduced
by Slater~\cite{Slater1975} (see~\cite{Harary}). For more results
related to these concepts
see~\cite{trees,bounds,sur1,Discrepancies}.
\par We say an ordered set $W$ {\it 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 $T$. 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 $d(w,w)=0$.
\par Two distinct vertices $u,v$ are said {\it twins}
if $N(v)\backslash\{u\}=N(u)\backslash\{v\}$. It is called that
$u\equiv v$ if and only if $u=v$ or $u,v$ are twins.
In~\cite{extermal}, it is proved that ``$\equiv$" is an equivalent
relation. The equivalence class of vertex $v$ is denoted by
$v^*$. Hernando et al.~\cite{extermal} proved that $v^*$ is a
clique or an independent set in $G$. As in ~\cite{extermal}, we
say $v^*$ is of type (1), (K), or (N) if $v^*$ is a class of size
$1$, a clique of size at least $2$, or an independent set of size
at least $2$. We denote the number of equivalence classes of $G$
with respect to ``$\equiv$" by $\iota(G)$. We mean by
$\iota_{_K}(G)$ and $\iota_{_N}(G)$, the number of classes of
type (K) and type (N) in $G$, respectively. We also use $a(G)$ and
$b(G)$ for the number of all vertices in $G$ which have at least
an adjacent twin and a none-adjacent twin vertex in $G$,
respectively. On the other way, $a(G)$ is the number of all
vertices in the classes of type (K) and $b(G)$ is the number of
all vertices in the classes of type (N). Clearly,
$\iota(G)=n(G)-a(G)-b(G)+\iota_{_N}(G)+\iota_{_K}(G)$.
\begin{obs}~{\rm\cite{extermal}}\label{twins}
Suppose that $u,v$ are twins in a graph $G$ and $W$ resolves $G$.
Then $u$ or $v$ is in $W$. Moreover, if $u\in W$ and $v\notin W$,
then $(W\setminus\{u\})\cup \{v\}$ also resolves $G$.
\end{obs}
\begin{lem}~\rm\cite{Ollerman}\label{B=1,B=n-1}
Let $G$ be a connected 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}
\begin{lem}~\rm\cite{K dimensional,idea}\label{B(P_n) B(C_N)}
\begin{description}
\item (i) If $n\notin\{3,6\}$, then $\beta(C_n\vee
K_1)=\lfloor{{2n+2}\over 5}\rfloor$, \item (ii) If
$n\notin\{1,2,3,6\}$, then $\beta(P_n\vee K_1)=\lfloor{{2n+2}\over
5}\rfloor$.
\end{description}
\end{lem}
The metric dimension of cartesian product of graphs is studied by
Caseres et al. in ~\cite{cartesian product}. They obtained the
metric dimension of cartesian product of graphs $G$ and $H$,
$G\square H$, where $G,H\in \{P_n,C_n,K_n\}$.
\par The {\it lexicographic product} of graphs $G$ and $H$, denoted by
$G[H]$, is a graph with vertex set \linebreak $V(G)\times
V(H):=\{(v,u)~|~v\in V(G),u\in V(H)\}$, where two vertices
$(v,u)$ and $(v',u')$ are adjacent whenever, $v\sim v'$, or
$v=v'$ and $u\sim u'$. When the order of $G$ is at least $2$, it
is easy to see that $G[H]$ is a connected graph if and only if
$G$ is a connected graph.
This paper is aimed to investigate the metric dimension of
lexicographic product of graphs.
The main goal of Section~\ref{Adjacency Resolving Sets} is
introducing a new parameter, which we call it adjacency metric
dimension. In Section~\ref{Lexicographic product}, we prove some
relations to determine the metric dimension of lexicographic
product of graphs, $G[H]$, in terms of the order of $G$ and the
adjacency metric dimension of $H$. As a corollary of our main
theorems, we obtain the exact value of the metric dimension of
$G[H]$, where $G=C_n(n\geq 5)$ or $G=P_n (n\geq4)$, and
$H\in\{P_m,C_m,\overline P_m,\overline
C_m,K_{m_1,\ldots,m_t},\overline K_{m_1,\ldots,m_t}\}$.
\section{Adjacency Resolving Sets}\label{Adjacency Resolving Sets}
S. Khuller et al.~\cite{landmarks} have considered the
application of the metric dimension of a connected graph in robot
navigation. In that sense, a robot moves from node to node of a
graph space. If the robot knows its distances to a sufficiently
large set of landmarks, its position on the graph is uniquely
determined. This suggest the problem of finding the fewest number
of landmarks needed, and where should be located, so that the
distances to the landmarks uniquely determine the robot's position
on the graph. The solution of this problem is the metric dimension
and a basis of the graph.
\par Now let there exist a large
number of landmarks, but the cost of computing distance is much
for the robot. In this case, robot can determine its position on
the graph only by knowing landmarks which are adjacent to it.
Here, the problem of finding the fewest number of landmarks
needed, and where should be located, so that the adjacency and
none-adjacency to the landmarks uniquely determine the robot's
position on the graph is a different problem. The answer to this
problem is one of the motivations of introducing {\it adjacency
resolving sets} in graphs.
\begin{deff}\label{Adjacency metric dimension}
Let $G$ be a graph and $W=\{w_1,w_2,\ldots,w_k\}$ be an ordered
subset of $V(G)$. For each vertex $v\in V(G)$ the adjacency
representation of $v$ with respect to $W$ is $k$-vector
$$r_2(v|W):=(a_G(v,w_1),a_G(v,w_2),\ldots,a_G(v,w_k)),$$ where
$$a_G(v,w_i)=\left\{
\begin{array}{ll}
0 & {\rm if}~v=w_i, \\
1 & {\rm if}~v\sim w_i,\\
2 & {\rm if}~v\nsim w_i.
\end{array}\right.$$ If all distinct vertices of $G$ have distinct
adjacency representations, $W$ is called an adjacency resolving
set for $G$. The minimum cardinality of an adjacency resolving set
is called adjacency metric dimension of $G$, denoted by
$\beta_2(G)$. An adjacency resolving set of cardinality
$\beta_2(G)$ is called an adjacency basis of $G$.
\end{deff}
\par By the definition, if $G$ is a
connected graph with diameter $2$, then $\beta(G)=\beta_2(G)$. The
converse is false; it can be seen that
$\beta_2(C_6)=2=\beta(C_6)$ while, $diam(C_6)=3$.
\par In the following, we obtain some useful
results on the adjacency metric dimension of graphs.
\begin{pro}\label{B(H)<B2(H)} For every connected graph $G$, $\beta(G)\leq \beta_2(G)$.
\end{pro}
\begin{proof}{
Let $W$ be an adjacency basis of $G$. Thus, for each pair of
vertices $u,v\in V(G)$ there exist a vertex $w\in W$ such that,
$a_G(u,w)\neq a_G(v,w)$. Therefore, $d_G(u,w)\neq d_G(v,w)$ and
hence $W$ is a resolving set for $G$. }\end{proof}
\begin{pro}\label{B2(H)=B2(overlineH)}
For every graph $G$, $\beta_2(G)=\beta_2(\overline{G})$.
\end{pro}
\begin{proof}{Let $W$ be an adjacency basis of $G$. For each pair of
vertices $u,v\in V(G)$, there exist a vertex $w\in W$ such that
$a_G(u,w)\neq a_G(v,w)$. Without loss of generality, assume that
$a_G(u,w)< a_G(v,w)$. Thus, if $a_G(u,w)=0$, then
$a_{\overline{G}}(u,w)=0$ and $a_{\overline{G}}(v,w)>0$. Also, if
$a_G(u,w)=1$, then $a_G(v,w)=2$ and hence,
$a_{\overline{G}}(u,w)=2$ and $a_{\overline{G}}(v,w)=1$.
Therefore, $W$ is an adjacency resolving set for $\overline{G}$
and $\beta_2(\overline{G})\leq \beta_2(G)$. Since
$\overline{\overline{G}}=G$, we conclude that $\beta_2(G)\leq
\beta_2(\overline{G})$ and consequently, $\beta_2(G)=
\beta_2(\overline{G})$. }\end{proof} Let $G$ be a graph of order
$n$. It is easy to see that, $1\leq \beta_2(G)\leq n-1$. In the
following proposition, we characterize all graphs $G$ with
$\beta_2(G)=1$ and all graphs $G$ of order $n$ and
$\beta_2(G)=n-1$.
\begin{pro}\label{characterization 1, m-1}
If $G$ is a graph of order $n$, then
\begin{description}
\item (i) $\beta_2(G)=1$ if and only if
$G\in\{P_1,P_2,P_3,\overline{P}_2,\overline{P}_3\}$.
\item (ii) $\beta_2(G)=n-1$ if and only if $G=K_n$ or
$G=\overline{K}_n$.
\end{description}
\end{pro}
\begin{proof}{(i) It is easy to see that for
$G\in\{P_1,P_2,P_3,\overline{P}_2,\overline{P}_3\}$,
$\beta_2(G)=1$. Conversely, let $G$ be a graph with
$\beta_2(G)=1$. If $G$ is a connected graph, then by
Proposition~\ref{B(H)<B2(H)}, $\beta(G)\leq \beta_2(G)=1$. Thus,
by Theorem~\ref{B=1,B=n-1}, $G=P_n$. If $n\geq 4$, then
$\beta_2(P_n)\geq 2$. Hence, $n\leq3$. If $G$ is a disconnected
graph and $\beta_2(G)=1$, then $\overline{G}$ is a connected graph
and by Proposition~\ref{B2(H)=B2(overlineH)},
$\beta_2(\overline{G})=1$. Thus, $\overline{G}=P_n$,
$n\in\{2,3\}.$ Therefore, $G=\overline{P}_2$ or
$G=\overline{P}_3$.
\noindent (ii) By Proposition~\ref{B(H)<B2(H)}, we have
$n-1=\beta(K_n)\leq \beta_2(K_n)$. On the other hand,
$\beta_2(G)\leq n-1$. Therefore, $\beta_2(K_n)=n-1$ and by
Proposition~\ref{B2(H)=B2(overlineH)},
$\beta_2(\overline{K}_n)=\beta_2(K_n)=n-1$. Conversely, let $G$ be
a connected graph with $\beta_2(G)=n-1$. Suppose on the contrary
that $G\neq K_n$. Thus, $P_3$ is an induced subgraph of $G$. Let
$P_3=(x_1,x_2,x_3)$. Therefore, $a_G(x_2,x_1)=1$ and
$a_G(x_3,x_1)=2$. Consequently, $V(G)\backslash\{x_2,x_3\}$ is an
adjacency resolving set for $G$ of cardinality $n-2$. That is,
$\beta_2(G)\leq n-2$, which is a contradiction. Hence, $G=K_n$. If
$G$ is a disconnected graph with $\beta_2(G)=n-1$, then
$\overline{G}$ is a connected graph and by
Proposition~\ref{B2(H)=B2(overlineH)},
$\beta_2(\overline{G})=n-1$. Thus, $\overline{G}=K_n$.
}\end{proof}
\begin{lemma}\label{delta=n-1}
If $u$ is a vertex of degree $n(G)-1$ in a connected graph $G$,
then $G$ has a basis which does not include $u$.
\end{lemma}
\begin{proof}{Let $B$ be a basis of $G$ which contains $u$.
Thus, $r(u|B\backslash\{u\})=(1,\ldots,1)$. Since $B$ is a basis
of $G$, there exist two vertices $v,w\in
V(G)\backslash(B\backslash\{u\})$ such that,
$r(v|B\backslash\{u\})=r(w|B\backslash\{u\})$ and $d_G(u,v)\neq
d_G(u,w)$. If $u\notin\{v,w\}$, then $d(u,v)=d(u,w)=1$, which is a
contradiction. Hence, $u\in\{v,w\}$, say $u=v$. Therefore,
$r(w|B\backslash\{u\})=r(u|B\backslash\{u\})=(1,1,\ldots,1)$ and
for each $x,y\in V(G)\backslash\{u,w\}$,
$r(x|B\backslash\{u\})\neq r(y|B\backslash\{u\})$. Note that,
$r(w|B)=(1,1,\ldots,1)$, because $u\sim w$. Since $B$ is a basis
of $G$, $w$ is the unique vertex of $G$ which its representation
with respect to $B$ is entirely $1$. It implies that $w$ is the
unique vertex of $G\backslash B$ with
$r(w|B\backslash\{u\})=(1,1,\ldots,1)$. Therefore, the set
$(B\backslash\{u\})\cup\{w\}$ is a basis of $G$ which does not
contain $u$.
}\end{proof}
\begin{pro}\label{H join K_1}
For every graph $G$, $\beta(G\vee K_1)-1\leq \beta_2(G)\leq
\beta(G\vee K_1)$. Moreover, $\beta_2(G)=\beta(G\vee K_1)$ if and
only if $G$ has an adjacency basis for which no vertex has
adjacency representation entirely $1$ with respect to it.
\end{pro}
\begin{proof}{
Let $V(G)=\{v_1,v_2,\ldots,v_n\}$ and $V(K_1)=\{u\}$. Note that,
$d_{G\vee K_1}(v_i,v_j)=a_G(v_i,v_j)$, $1\leq i,j \leq n$. By
Lemma~\ref{delta=n-1}, $G\vee K_1$ has a basis
$B=\{b_1,b_2,\ldots,b_k\}$ such that $u\notin B$.
Therefore,$$r(v_i|B)=(d_{G\vee K_1}(v_i,b_1),d_{G\vee
K_1}(v_i,b_2),\ldots,d_{G\vee K_1}(v_i,b_k))=r_2(v_i|B)$$ for each
$v_i$, $1\leq i\leq n$. Thus, $B$ is an adjacency resolving set
for $G$ and $\beta_2(G)\leq \beta(G\vee K_1)$.
\par Now let $W=\{w_1,w_2,\ldots,w_t\}$ be an adjacency basis of $G$.
Since $d_{G\vee K_1}(v_i,w_j)=a_G(v_i,w_j)$, $1\leq i \leq n$,
$1\leq j\leq t$, we have $r(v_i|W)=r_2(v_i|W)$, $1\leq i\leq n$.
Hence, $W$ resolves $V(G\vee K_1)\backslash \{u\}$ and
$\beta(G\vee K_1)-1\leq \beta_2(G)$. On the other hand, $r(u|W)$
is entirely $1$. Therefore, $W$ is a resolving set for $G\vee K_1$
if and only if $r_2(v_i|W)$ is not entirely $1$ for every $v_i$,
$1\leq i\leq n$. Since $\beta_2(G)\leq \beta(G\vee K_1)$, we have
$\beta_2(G)=\beta(G\vee K_1)$ if and only if $r_2(v_i|W)$ is not
entirely $1$ for every $v_i$, $1\leq i\leq n$. }\end{proof}
\begin{pro}\label{B_2(P_m),B_2(C_m)}
If $n\geq 4$, then $\beta_2(C_n)=\beta_2(P_n)=\lfloor{{2n+2}\over
5}\rfloor$.
\end{pro}
\begin{proof}{
If $n\leq 8$, then by a simple computation, we can see that
$\beta_2(C_n)=\beta_2(P_n)=\lfloor{{2n+2}\over 5}\rfloor$. Now,
let $G\in\{P_n,C_n\}$, and $n\geq 9$. By Theorem~\ref{B(P_n)
B(C_N)}, $\beta(G\vee K_1)=\lfloor{{2n+2}\over 5}\rfloor\geq 4$.
Hence, by Proposition~\ref{H join K_1}, we have $\beta_2(G)\geq
3$. If $W$ is an adjacency basis of $G$, then for each vertex
$v\in V(G)$, $r_2(v|W)$ is not entirely $1$, because $v$ has at
most two neighbors. Therefore, by Proposition~\ref{H join K_1},
$\beta_2(G)=\beta_2(G\vee K_1)=\lfloor{{2n+2}\over 5}\rfloor$.
}\end{proof}
\begin{pro}\label{B_2 multipartite} If $K_{m_1,m_2,\ldots,m_t}$ is
the complete $t$-partite graph, then
$$\beta_2(K_{m_1,m_2,\ldots,m_t})=\beta(K_{m_1,m_2,\ldots,m_t})=\left\{
\begin{array}{ll}
m-r-1 & ~{\rm if}~r\neq t, \\
m-r & ~{\rm if}~r=t, \end{array}\right.$$ where $m_1,
m_2,\ldots,m_r$ are at least $2$, $m_{r+1}=\cdots=m_t=1$, and
$\sum_{i=1}^tm_i=m$.
\end{pro}
\begin{proof}{ Since $diam(K_{m_1,m_2,\ldots,m_t})=2$,
we have $\beta_2(K_{m_1,m_2,\ldots,m_t})=
\beta(K_{m_1,m_2,\ldots,m_t})$. Let $M_i$ be the partite set of
size $m_i$, $1\leq i\leq t$. For each $i,~1\leq i\leq r$, all
vertices of $M_i$ are none-adjacent twins. Also, all vertices of
$\cup_{i=r+1}^tM_i$ are adjacent twins. Let $x_i$ be a fixed
vertex in $M_i$, $1\leq i\leq r$. If $r=t$, then by
Observation~\ref{twins},
$\beta(K_{m_1,m_2,\ldots,m_t})\geq\sum_{i=1}^tm_i-r$. Also, the
set $\cup_{i=1}^tM_i\backslash\{x_1,x_2,\ldots,x_r\}$ is a
resolving set for $K_{m_1,m_2,\ldots,m_t}$ with cardinality
$\sum_{i=1}^tm_i-r$. Thus,
$\beta(K_{m_1,m_2,\ldots,m_t})=\sum_{i=1}^tm_i-r=m-r$. If $r\neq
t$, then $\cup_{i=r+1}^t M_i\neq\emptyset$. Let $x_{r+1}\in
\cup_{i=r+1}^t M_i$. Observation~\ref{twins} implies that
$\beta(K_{m_1,m_2,\ldots,m_t})\geq\sum_{i=1}^tm_i-r-1$. On the
other hand, the set
$\cup_{i=1}^tM_i\backslash\{x_1,x_2,\ldots,x_{r+1}\}$ is a
resolving set for $K_{m_1,m_2,\ldots,m_t}$ with cardinality
$\sum_{i=1}^tm_i-r-1=m-r-1$. }
\end{proof}
\section{Lexicographic Product of Graphs}\label{Lexicographic product}
Throughout this section, $G$ is a connected graph of order $n$,
$V(G)=\{v_1,v_2,\ldots,v_n\}$, $H$ is a graph of order $m$, and
$V(H)=\{u_1,u_2,\ldots,u_m\}$. Therefore, $G[H]$ is a connected
graph. For convenience, we denote the vertex $(v_i,u_j)$ of
$G[H]$ by $v_{ij}$. Note that, for each pair of vertices
$v_{ij},v_{rs}\in V(G[H])$,
$$d_{G[H]}(v_{ij},v_{rs})=\left\{
\begin{array}{ll}
d_G(v_i,v_r) & ~{\rm if}~v_i\neq v_r, \\
1 & ~{\rm if}~v_i=v_r~{\rm and}~u_j\sim u_s,\\
2 & ~{\rm if}~v_i=v_r~{\rm and}~u_j\nsim u_s.
\end{array}\right.$$On the other words, $$d_{G[H]}(v_{ij},v_{rs})=\left\{
\begin{array}{ll}
d_G(v_i,v_r) & ~{\rm if}~v_i\neq v_r, \\
a_H(u_j,u_s)& ~{\rm otherwise.}
\end{array}\right.$$
Let $S$ be a subset of $V(G[H])$. The {\it projection} of $S$ onto
$H$ is the set $\{u_j\in V(H)\ |\ v_{ij}\in S\}$. Also, the ith
{\it row} of $G[H]$, denoted by $H_i$, is the set $\{v_{ij}\in
V(G[H])\ |\ 1\leq j\leq m\}.$
\begin{lemma}\label{S i Resolves H_i} If\- $W\subseteq V(G[H])$ is a
resolving set for $G[H]$, then $W\cap H_i$ resolves $H_i$, for
each $i,~1\leq i\leq n$. Moreover, the projection of $W\cap H_i$
onto $H$ is an adjacency resolving set for $H$, for each $i,~1\leq
i\leq n$.
\end{lemma}
\begin{proof}{
Since $W$ resolves $G[H]$, for each pair of vertices
$v_{ij},v_{iq}\in H_i$, there exist a vertex $v_{rt}\in W$ such
that, $d_{G[H]}(v_{rt},v_{ij})\neq d_{G[H]}(v_{rt},v_{iq})$. If
$r\neq i$, then $d_{G[H]}(v_{rt},v_{ij})=d_G(v_r,v_i)=
d_{G[H]}(v_{rt},v_{iq})$, which is a contradiction. Therefore,
$i=r$ and $W\cap H_i$ resolves $H_i$.
\par Now, let $u_j,u_q\in
V(H)$. Since $W\cap H_i$ resolves $H_i$, there exist a vertex
$v_{it}\in W\cap H_i$ such that, $d_{G[H]}(v_{it},v_{ij})\neq
d_{G[H]}(v_{it},v_{iq})$. Hence,
$a_H(u_t,u_j)=d_{G[H]}(v_{it},v_{ij})\neq
d_{G[H]}(v_{it},v_{iq})=a_H(u_t,u_q)$. Consequently, the
projection of $W\cap H_i$ onto $H$ is an adjacency resolving set
for $H$. }
\end{proof}
By Lemma~\ref{S i Resolves H_i}, every basis of $G[H]$ contains
at least $\beta_2(H)$ vertices from each copy of $H$ in $G[H]$.
Thus, the following lower bound for $\beta(G[H])$ is obtained.
\begin{equation}\label{l bound}
\beta(G[H])\geq n\beta_2(H).
\end{equation}
\begin{thm}\label{B1 B2}
Let $G$ be a connected graph of order $n$ and $H$ be an arbitrary
graph. If there exist two adjacency bases $W_1$ and $W_2$ of $H$
such that, there is no vertex with adjacency representation
entirely~$1$ with respect to $W_1$ and no vertex with adjacency
representation entirely $2$ with respect to $W_2$, then
$\beta(G[H])=\beta(G[\overline H])=n\beta_2(H)$.
\end{thm}
\begin{proof}{
By Inequality~\ref{l bound}, we have $\beta(G[H])\geq
n\beta_2(H)$. To prove the equality, it is enough to provide a
resolving set for $G[H]$ of size $n\beta_2(H)$. For this sake, let
$$S=\{v_{ij}\in V(G[H])\ |\ v_i\in K(G),~u_j\in W_1\}\cup \{v_{ij}\in V(G[H])\
|\ v_i\notin K(G),~u_j\in W_2\},$$
where $K(G)$ is the set of all vertices of $G$ in equivalence
classes of type (K). On the other word, $K(G)$ is the set of all
vertices of $G$ which have adjacent twins. We show that $S$ is a
resolving set for $G[H]$.
Let $v_{rt},v_{pq}\in
V(G[H])\backslash S$ be two distinct vertices.
The following possibilities can be happened. \vspace{4mm}\\
1. $r=p$. Note that, $v_{rt}\neq v_{pq}$ implies that $t\neq q$.
Since $W_1$ and $W_2$ are adjacency resolving sets, there exist
vertices $u_j\in W_1$ and $u_l\in W_2$ such that,
$a_H(u_t,u_j)\neq a_H(u_q,u_j)$ and $a_H(u_t,u_l)\neq
a_H(u_q,u_l)$. If $v_r\in K(G)$, then $v_{rj}\in S$ and
$d_{G[H]}(v_{rt},v_{rj})=a_H(u_t,u_j)\neq
a_H(u_q,u_j)=d_{G[H]}(v_{pq},v_{rj})$. Similarly, if $v_r\notin
K(G)$, then $v_{rl}\in S$ and $d_{G[H]}(v_{rt},v_{rl})\neq
d_{G[H]}(v_{pq},v_{rl})$.\vspace{4mm}\\
2. $r\neq p$ and $v_r,v_p\in K(G)$. If $v_r$ and $v_p$ are not
twins, then there exist a vertex $v_i\in
V(G)\backslash\{v_r,v_p\}$ which is adjacent to only one of the
vertices $v_r$ and $v_p$. Hence, for each $u_j\in W_1$, we have
$v_{ij}\in S$ and $d_{G[H]}(v_{rt},v_{ij})=d_G(v_r,v_i)\neq
d_G(v_p,v_i)= d_{G[H]}(v_{pq},v_{ij})$. If $v_r$ and $v_p$ are
twins, then $v_r\sim v_p$, because $v_r,v_p\in K(G)$. Since
$r_2(u_t|W_1)$ is not entirely $1$, there exist a vertex $u_l\in
W_1$ such that, $a_H(u_t,u_l)=2$. Therefore, $v_{rl}\in S$ and
$d_{G[H]}(v_{rt},v_{rl})=a_H(u_t,u_l)=2$. On the other hand,
$d_{G[H]}(v_{pq},v_{rl})=d_G(v_p,v_r)=1$. Thus,
$d_{G[H]}(v_{rt},v_{rl})\neq d_{G[H]}(v_{pq},v_{rl})$. \vspace{4mm}\\
3. $r\neq p$, $v_r\in K(G)$, and $v_q\notin K(G)$. In this case,
$v_r$ and $v_p$ are not twins. Therefore, there exist a vertex
$v_i\in V(G)\backslash\{v_r,v_p\}$ which is adjacent to only one
of the vertices $v_r$ and $v_p$. Let $u_j$ be a vertex of $W_1\cup
W_2$, such that $v_{ij}\in S$. Hence,
$d_{G[H]}(v_{rt},v_{ij})=d_G(v_r,v_i)\neq d_G(v_p,v_i)=
d_{G[H]}(v_{pq},v_{ij})$. \vspace{4mm}\\
4. $r\neq p$ and $v_r,v_p\notin K(G)$. If $v_r$ and $v_p$ are not
twins, then there exist a vertex $v_i\in
V(G)\backslash\{v_r,v_p\}$ which is adjacent to only one of the
vertices $v_r$ and $v_p$. Thus, for each $u_j\in W_2$, we have
$v_{ij}\in S$ and $d_{G[H]}(v_{rt},v_{ij})=d_G(v_r,v_i)\neq
d_G(v_p,v_i)= d_{G[H]}(v_{pq},v_{ij})$. If $v_r$ and $v_p$ are
twins, then $v_r\nsim v_p$, because $v_r,v_p\notin K(G)$. Since
$r_2(u_t|W_2)$ is not entirely $2$, there exist a vertex $u_l\in
W_2$, such that $a_H(u_t,u_l)=1$. Therefore, $v_{rl}\in S$ and
$d_{G[H]}(v_{rt},v_{rl})=a_H(u_t,u_l)=1$. On the other hand,
$d_{G[H]}(v_{pq},v_{rl})=d_G(v_p,v_r)=2$, since $v_r$ and $v_p$
are none-adjacent twins in the connected $G$. Hence,
$d_{G[H]}(v_{rt},v_{rl})\neq d_{G[H]}(v_{pq},v_{rl})$.
\par Thus, $r(v_{rt}|S)\neq r(v_{pq}|S)$. Therefore, $S$ is a
resolving set for $G[H]$ with cardinality $n\beta_2(H)$.
\par Clearly, in $\overline H$, for each $u\in V(\overline
H)$, $r_2(u|W_1)$ is not entirely $2$ and $r_2(u|W_2)$ is not
entirely $1$. Since $\beta_2(H)=\beta_2(\overline H)$, by
interchanging the roles of $W_1$ and $W_2$ for $\overline H$, we
conclude $\beta(G[\overline H])=n\beta_2(\overline
H)=n\beta_2(H)$. }\end{proof} In the following three theorems, we
obtain $\beta(G[H])$, when $H$ does not satisfy the assumption of
Theorem~\ref{B1 B2}.
\begin{thm}\label{thm generalG[H]}
Let $G$ be a connected graph of order $n$ and $H$ be an arbitrary
graph. If for each adjacency basis $W$ of $H$ there exist vertices
with adjacency representations entirely $1$ and entirely $2$ with
respect to $W$, then $\beta(G[H])=\beta(G[\overline
H])=n(\beta_2(H)+1)-\iota(G).$
\end{thm}
\begin{proof}{ Let $B$ be a basis of $G[H]$ and $B_i$ be the
projection of $B\cap H_i$ onto $H$, for each $i$, $1\leq i\leq n$.
By Lemma~\ref{S i Resolves H_i}, $B_i$'s are adjacency resolving
sets for $H$. Therefore, $|B\cap H_i|=|B_i|\geq \beta_2(H)$ for
each $i$, $1\leq i\leq n$.
Let $I=\{i\ |\ |B_i|=\beta_2(H)\}$. We claim that
$|I|\leq\iota(G)$, otherwise by the pigeonhole principle, there
exist a pair of twin vertices $v_r,v_p\in V(G)$ such that,
$|B_r|=|B_p|=\beta_2(H)$. Since $B_r$ and $B_p$ are adjacency
bases of $H$, by the assumption there are vertices $u_t$ and
$u_q$ with adjacency representations entirely $1$ with respect to
$B_r$ and $B_p$, respectively. Also, there are vertices $u_t'$ and
$u_q'$ with adjacency representations entirely $2$ with respect to
$B_r$ and $B_p$, respectively.
Hence, for each $u\in B_r$ and $u^\prime\in B_p$, we have
$u_{t}\sim u,~u_{t^\prime}\nsim u,~u_{q}\sim u^\prime$, and
$u_{q^\prime}\nsim u^\prime$. If $v_r\sim v_p$, then for each
$v_{ij}\in B$ one of the following
cases can be happened.\vspace{4mm}\\
1. $i\notin\{r,p\}$. Since $v_r$ and $v_p$ are twins, we have
$d_G(v_r,v_i)=d_G(v_p,v_i)$. On the other hand,
$d_{G[H]}(v_{rt},v_{ij})=d_G(v_r,v_i)$ and
$d_{G[H]}(v_{pq},v_{ij})=d_G(v_p,v_i)$. Thus,
$d_{G[H]}(v_{rt},v_{ij})=d_{G[H]}(v_{pq},v_{ij})$.\vspace{4mm}\\
2. $i=p\neq r$. In this case,
$d_{G[H]}(v_{pq},v_{ij})=a_H(u_{q},u_j)$ and
$d_{G[H]}(v_{rt},v_{ij})=d_G(v_r,v_i)$. Since $v_i=v_p\sim v_r$,
we have $d_G(v_r,v_i)=1$. On the other hand $u_j\in B_p$ and
hence, $a_H(u_{q},u_j)=1$. Therefore,
$d_{G[H]}(v_{rt},v_{ij})=d_{G[H]}(v_{pq},v_{ij})$.\vspace{4mm}\\
3. $i=r\neq p$. Similar to previous case,
$d_{G[H]}(v_{rt},v_{ij})=a_H(u_{t},u_j)=1$ and
$d_{G[H]}(v_{pq},v_{ij})=d_G(v_p,v_i)=1$. Consequently,
$d_{G[H]}(v_{rt},v_{ij})=d_{G[H]}(v_{pq},v_{ij})$.\vspace{4mm}\\
4. $i=p=r$. In this case, $d_{G[H]}(v_{pq},v_{ij})=a_H(u_{q},u_j)$
and $d_{G[H]}(v_{rt},v_{ij})=a_H(u_{t},u_j)$. Since, $u_j\in
B_p=B_r$, we have $a_H(u_{q},u_j)=1=a_H(u_{t},u_j)$. Thus,
$d_{G[H]}(v_{rt},v_{ij})=d_{G[H]}(v_{pq},v_{ij})$.
\vspace{4mm}\\
Hence, $v_r\sim v_p$ implies that $r(v_{rt}|B)=r(v_{pq}|B)$, which
is a contradiction. Therefore, $v_r\nsim v_p$. Since $G$ is a
connected graph, none-adjacent twin vertices $v_r$ and $v_p$ have
at least one common neighbor and thus, $d_G(v_r,v_p)=2$.
Consequently, by a same method as the case $v_r\sim v_p$, we can
see that $r(v_{rt^\prime}|B)=r(v_{pq^\prime}|B)$, which
contradicts the assumption that $B$ is a basis of $G[H]$. Hence
$|I|\leq \iota(G)$. On the other hand, every basis of $G[H]$ has
at least $\beta_2(H)+1$ vertices in $H_i$, where $i\notin I$.
Therefore,
\begin{eqnarray*}
\beta(G[H])=|B|=|\cup_{i=1}^n(B\cap H_i)|&\geq &|I|\beta_2(H)+(n-|I|)(\beta_2(H)+1)\\
&=&n\beta_2(H)+n-|I|\\
&\geq & n(\beta_2(H)+1)-\iota(G).
\end{eqnarray*}
Now let $W$ be an adjacency basis of $H$. By assumption, there
exist vertices $u_1,u_2\in V(H)\backslash W$ such that, $u_1$ is
adjacent to all vertices of $W$ and $u_2$ is not adjacent to any
vertex of $W$. Also, let $K(G)$ be the set of all classes of
type (K), and $N(G)$ be the set of all classes of $G$ of type (N)
in $G$. Choose fixed vertex $v$ from $v^*$ for each $v^*\in
N(G)\cup K(G)$. We claim that the set
$$S=\{v_{ij}\in V(G[H])\,|\,u_j\in W\}\cup\{v_{t1}\,|\,v_t\in
\cup_{v^*\in K(G)}(v^*\backslash\{v\})\} \cup\{v_{t2}\,|\,v_t\in
\cup_{v^*\in N(G)}(v^*\backslash\{v\})\}$$ is a resolving set for
$G[H]$. Let $v_{rt},v_{pq}\in V(G[H])\backslash S$. Hence, one
of the following cases can be happened.\vspace{4mm}\\
1. $r=p$. Since $W$ is an adjacency basis of $H$, there exist a
vertex $u_j\in W$, such that $a_H(u_q,u_j)\neq a_H(u_t,u_j)$.
Therefore, $d_{G[H]}(v_{pq},v_{rj})=a_H(u_q,u_j)\neq
a_H(u_t,u_j)=d_{G[H]}(v_{rt},v_{rj})$. Consequently,
$r(v_{rt}|S)\neq r(v_{pq}|S)$. \vspace{4mm}\\
2. $r\neq p$ and $v_r,v_p$ are not twins. Hence, there exist a
vertex $v_i\in V(G)$ which is adjacent to only one of the vertices
$v_r$ and $v_p$. Thus, for each vertex $u_j\in W$,
$d_{G[H]}(v_{rt},v_{ij})=d_G(v_r,v_i)\neq
d_G(v_p,v_i)=d_{G[H]}(v_{pq},v_{ij})$.
This yields, $r(v_{rt}|S)\neq r(v_{pq}|S)$. \vspace{4mm}\\
3. $v_r$ and $v_p$ are adjacent twins. Therefore, at least one of
the vertices $v_{r1}$ and $v_{p1}$, say $v_{r1}$ belongs to $S$.
Since $v_{rt}\notin S$, we have $t\neq 1$. Hence, there exists a
vertex $u_j\in S$, such that $a_H(u_t,u_j)=2$, otherwise $t=1$.
Consequently, $d_{G[H]}(v_{rt},v_{rj})=a_H(u_t,u_j)=2$. On the
other hand, $d_{G[H]}(v_{pq},v_{rj})=d_G(v_p,v_r)=1$, because
$v_r\sim v_p$. This gives, $r(v_{rt}|S)\neq r(v_{pq}|S)$. \vspace{4mm}\\
4. $v_r$ and $v_p$ are none-adjacent twins. In this case, at least
one of the vertices $v_{r2}$ and $v_{p2}$, say $v_{r2}$ belongs to
$S$. Hence, $t\neq 2$ and there exists a vertex $u_j\in W$, such
that $a_H(u_t,u_j)=1$, otherwise $t=2$. Therefore,
$d_{G[H]}(v_{rt},v_{rj})=a_H(u_t,u_j)=1\neq2=d_G(v_p,v_r)=d_{G[H]}(v_{pq},v_{rj})$.
Thus, $r(v_{rt}|S)\neq r(v_{pq}|S)$. \vspace{4mm}\\
Consequently, $S$ is a resolving set for $G[H]$ with cardinality
$$|S|=n\beta_2(H)+a(G)-\iota_{_K}(G)+b(G)-\iota_{_N}(G)=n(\beta_2(H)+1)-\iota(G). $$ Since all adjacency bases of $H$ and
$\overline H$ are the same, $\overline H$ satisfies the condition
of the theorem. Hence, $\beta(G[\overline
H])=n(\beta_2(H)+1)-\iota(G)$ and the proof is completed.
}\end{proof}
\begin{thm}\label{nB2+a(G) K(G)}
Let $G$ be a connected graph of order $n$ and $H$ be an arbitrary
graph. If $H$ has the following properties
\begin{description}\item (i) for each adjacency basis of $H$
there exist a vertex with adjacency representation entirely $1$,
\item (ii) there exist an adjacency basis $W$ of $H$ such that
there is no vertex with adjacency representation entirely $2$ with
respect to $W$,
\end{description} then
$\beta(G[H])=n\beta_2(H)+a(G)-\iota_{_K}(G).$
\end{thm}
\begin{proof}{
Let $B$ be a basis of $G[H]$ and $B_i$ be the projection of $B\cap
H_i$ onto $H$, for each $i$, $1\leq i\leq n$. By Lemma~\ref{S i
Resolves H_i}, $B_i$'s are adjacency resolving sets for $H$.
Therefore, $|B\cap H_i|=|B_i|\geq \beta_2(H)$ for each $i$, $1\leq
i\leq n$.
Let $I=\{i\ |\ |B_i|=\beta_2(H)\}$. We claim that $|I|\leq
n-a(G)+\iota_{_K}(G)$, otherwise by the pigeonhole principle,
there exist a pair of adjacent twin vertices $v_r,v_p\in V(G)$,
such that $|B_r|=|B_p|=\beta_2(H)$. Since $B_r$ and $B_p$ are
adjacency bases of $H$, by assumption $(i)$ there exist
vertices $u_{t},u_q\in V(H)$ with adjacency representation
entirely $1$ with respect to $B_r$ and $B_p$, respectively.
Hence, for each $u\in B_r$ and each $u^\prime\in B_p$, we have
$u_{t}\sim u$, and $u_q\sim u'$. Since $v_r\sim v_p$, for each
$v_{ij}\in B$ one of the following
cases can be happened.\vspace{4mm}\\
1. $i\notin\{r,p\}$. Since $v_r$ and $v_p$ are twins, we have
$d_G(v_r,v_i)=d_G(v_p,v_i)$. On the other hand,
$d_{G[H]}(v_{rt},v_{ij})=d_G(v_r,v_i)$ and
$d_{G[H]}(v_{pq},v_{ij})=d_G(v_p,v_i)$. Thus,
$d_{G[H]}(v_{rt},v_{ij})=d_{G[H]}(v_{pq},v_{ij})$.\vspace{4mm}\\
2. $i=p\neq r$. In this case,
$d_{G[H]}(v_{pq},v_{ij})=a_H(u_{q},u_j)$ and
$d_{G[H]}(v_{rt},v_{ij})=d_G(v_r,v_i)$. Since $v_i=v_p\sim v_r$,
we have $d_G(v_r,v_i)=1$. On the other hand $u_j\in B_p$ and
hence, $a_H(u_{q},u_j)=1$. Therefore,
$d_{G[H]}(v_{rt},v_{ij})=d_{G[H]}(v_{pq},v_{ij})$.\vspace{4mm}\\
3. $i=r\neq p$. Similar to previous case,
$d_{G[H]}(v_{rt},v_{ij})=a_H(u_{t},u_j)=1$ and
$d_{G[H]}(v_{pq},v_{ij})=d_G(v_p,v_i)=1$. Consequently,
$d_{G[H]}(v_{rt},v_{ij})=d_{G[H]}(v_{pq},v_{ij})$.\vspace{4mm}\\
4. $i=p=r$. In this case, $d_{G[H]}(v_{pq},v_{ij})=a_H(u_{q},u_j)$
and $d_{G[H]}(v_{rt},v_{ij})=a_H(u_{t},u_j)$. Since, $u_j\in
B_p=B_r$, we have $a_H(u_{q},u_j)=1=a_H(u_{t},u_j)$. Thus,
$d_{G[H]}(v_{rt},v_{ij})=d_{G[H]}(v_{pq},v_{ij})$.
\vspace{4mm}\\
Hence, $r(v_{rt}|B)=r(v_{pq}|B)$, which is a contradiction.
Therefore, $|I|\leq n-a(G)+\iota_{_K}(G)$. On the other hand,
every basis of $G[H]$ has at least $\beta_2(H)+1$ vertices in
$H_i$, where $i\notin I$. Thus,
\begin{eqnarray*}
\beta(G[H])=|B|&\geq &|I|\beta_2(H)+(n-|I|)(\beta_2(H)+1)\\
&=&n\beta_2(H)+n-|I|\\
&\geq & n\beta_2(H)+a(G)-\iota_{_K}(G).
\end{eqnarray*}
Now let $K(G)$ be the set of all classes of type (K) in $G$ and
$v\in v^*$ be a fixed vertex for each class $v^*$ of type (K).
Also, let $u_1\in V(H)\backslash W$, such that $r_2(u_1|W)$ is
entirely $1$. Consider $$S=\{v_{ij}\in V(G[H])\,|\,u_j\in
W\}\cup\{v_{t1}\,|\,v_t\in \cup_{v^*\in
K(G)}(v^*\backslash\{v\})\}$$ and let $v_{rt},v_{pq}\in
V(G[H])\backslash S$. If $v_r$ and $v_p$ are not none-adjacent
twins, then similar to the proof of Theorem~\ref{thm generalG[H]},
we have $r(v_{rt}|S)\neq r(v_{pq}|S)$. Now, let $v_r$ and $v_p$ be
none-adjacent twin vertices of $G$. By assumption, there exists a
vertex $u_j\in W$, such that $a_H(u_t,u_j)=1$. Therefore,
$d_{G[H]}(v_{rt},v_{rj})=a_H(u_t,u_j)=1$. On the other hand,
$d_{G[H]}(v_{pq},v_{rj})=d_G(v_p,v_r)=2$, since $v_r$ and $v_p$
are none-adjacent twins in the connected graph $G$. Hence,
$r(v_{rt}|S)\neq r(v_{pq}|S)$. This implies that $S$ is a
resolving set for $G[H]$ with cardinality
$n\beta_2(H)+a(G)-\iota_{_K}(G)$.
}\end{proof}
By a similar proof, we have the following theorem.
\begin{thm}\label{nB2+b(G) N(G)}
Let $G$ be a connected graph of order $n$ and $H$ be an arbitrary
graph. If $H$ has the following properties
\begin{description}\item (i) for each adjacency basis of $H$
there exist a vertex with adjacency representation entirely $2$,
\item (ii) there exist an adjacency basis $W$ of $H$ such that
there is no vertex with adjacency representation entirely~$1$ with
respect to $W$,
\end{description} then
$\beta(G[H])=n\beta_2(H)+b(G)-\iota_{_N}(G).$
\end{thm}
\begin{cor}\label{no twin} If $G$ has no pair of twin
vertices, then $\beta (G[H])=n\beta_2(H)$.
\end{cor}
\begin{proof}{ The
adjacency bases of $H$ satisfy one of the conditions of Theorems
~\ref{B1 B2},~\ref{thm generalG[H]},~\ref{nB2+a(G) K(G)},
and~\ref{nB2+b(G) N(G)} . Now, if $G$ does not have any pair of
twin vertices, then $\iota(G)=n,~\iota_{_K}(G)=a(G)=0$, and
$\iota_{_N}(G)=b(G)=0$. Therefore, $\beta
(G[H])=n\beta_2(H)$.}\end{proof} By Theorems~\ref{B1 B2},~\ref{thm
generalG[H]},~\ref{nB2+a(G) K(G)}, and~\ref{nB2+b(G) N(G)} the
exact value of $\beta(G[H])$ of many graphs $G$ and $H$ can be
determined. In the following two corollaries, $\beta(G[H])$ for
some of the well known graphs are obtained.
\begin{cor}\label{G=P_n or C_n} Let $G=P_n$, $n\geq4$ or $G=C_n$,
$n\geq5$. Then, $G$ does not have any pair of twin vertices. Thus
by Corollary~\ref{no twin}, $\beta(G[H])=n\beta_2(H)$, for each
graph $H$. In particular, by
Propositions~\ref{B2(H)=B2(overlineH)} and
\ref{B_2(P_m),B_2(C_m)},
$\beta_2(P_m)=\beta_2(C_m)=\beta_2(\overline
P_m)=\beta_2(\overline C_m)=\lfloor{2m+2\over5}\rfloor$.
Therefore, $\beta(G[P_m])=\beta(G[C_m])=\beta(G[\overline
P_m])=\beta(G[\overline C_m])=n\lfloor{2m+2\over5}\rfloor$. Also,
by Propositions~\ref{B2(H)=B2(overlineH)} and \ref{B_2
multipartite}, we have
$$\beta(G[\overline K_{m_1,m_2,\ldots,m_t}])=\beta(G[K_{m_1,m_2,\ldots,m_t}])=\left\{
\begin{array}{ll}
n(m-r-1) & ~{\rm if}~r\neq t, \\
n(m-r) & ~{\rm if}~r=t, \end{array}\right.$$ where $m_1,
m_2,\ldots,m_r$ are at least $2$, $m_{r+1}=\cdots=m_t=1$, and
$\sum_{i=1}^tm_i=m$.
\end{cor}
\begin{cor}\label{H=K_m1,m2,..,mt} Let $H=K_{m_1,m_2,\ldots,m_t}$, where $m_1,
m_2,\ldots,m_r$ are at least $2$, $m_{r+1}=\cdots=m_t=1$, and
$\sum_{i=1}^tm_i=m$. Thus, for each adjacency basis of $H$ there
is no vertex of $H$ with adjacency representation entirely $2$.
\par If $r=t$, then for each adjacency basis of $H$ there is no
vertex of $H$ with adjacency representation entirely $1$.
Therefore, by Theorem~\ref{B1 B2}, $\beta(G[H])=n\beta_2(H)$ for
each connected graph $G$ of order $n$. If $r\neq t$, then for
each adjacency basis of $H$, there exist a vertex with adjacency
representation entirely~$1$. Thus, by Theorem~\ref{nB2+a(G)
K(G)}, $\beta(G[H])=n\beta_2(H)+a(G)-\iota_{_K}(G)$ for each
connected graph $G$ of order $n$.
\par In particular, if $G=K_n$,
then all vertices of $K_n$ are adjacent twins. Thus, $a(K_n)=n$
and $\iota_{_K}(K_n)=1$, hence, $\beta(K_n[H])=n\beta_2(H)+n-1$.
Therefore, by Proposition~\ref{B_2 multipartite},
$$\beta(K_n[H])=\left\{
\begin{array}{ll}
n(m-r)-1 & ~{\rm if}~r\neq t, \\
n(m-r) & ~{\rm if}~r=t. \end{array}\right.$$
\end{cor}
|
\section{Introduction and summary}
Over the last few years several new classes of ${\mathcal N}=8$ $d=3$ superconformal field theories have been discovered \cite{BL,G,ABJM}. Until then, it had been widely assumed that the only such theories are infrared limits of ${\mathcal N}=8$ super-Yang-Mills theories and therefore are infinitely-strongly coupled. The newly discovered theories are not of this type. Rather they are Chern-Simons-matter theories which are superconformal already on the classical level. First of all, there are BLG theories \cite{BL,G} which have gauge group $SU(2)\times SU(2)$ \cite{vanR} and an arbitrary Chern-Simons coupling. ${\mathcal N}=8$ supersymmetry in these theories is visible on the classical level. Then there are ${\mathcal N}=8$ ABJM theories \cite{ABJM} which have gauge group $U(N)\times U(N)$ and have Chern-Simons coupling $k=1$ or $k=2$. These theories have ${\mathcal N}=6$ supersymmetry on the classical level, and ${\mathcal N}=8$ supersymmetry arises as a quantum effect. ${\mathcal N}=8$ ABJM theories are strongly coupled, but they have a a weakly-coupled AdS-dual description in the large-N limit \cite{ABJM} and describe the physics of M2-branes.
In this paper we exhibit another class of ${\mathcal N}=8$ $d=3$ superconformal Chern-Simons-matter theories. The theories themselves are not new: they are a special class of ABJ theories describing fractional M2-branes \cite{ABJ}. The gauge group of ABJ theories is $U(M)\times U(N)$ with Chern-Simons couplings $k$ and $-k$ for the two factors. These theories have ${\mathcal N}=6$ superconformal symmetry on the classical level for all values of $M,N,$ and $k$. We will show that for $M=N+1$ and $k=\pm 2$ they have hidden ${\mathcal N}=8$ supersymmetry on the quantum level. The same kind of arguments were used by us in \cite{BK} to show that ABJM theories with gauge group $U(N)_k\times U(N)_{-k}$ and $k=1,2$ have hidden ${\mathcal N}=8$ supersymmetry.
At first sight it might seem unlikely that ABJ theories may have ${\mathcal N}=8$ supersymmetry for $N\neq M$. These theories are not parity-invariant on the classical level, while all hitherto known ${\mathcal N}=8$ $d=3$ theories are parity-invariant. On the other hand, we know of no reason why ${\mathcal N}=8$ supersymmetry should imply parity-invariance. We will see that $U(N+1)_2\times U(N)_{-2}$ theories do have hidden parity-invariance on the quantum level. The definition of the parity transformation involves a nontrivial duality on one of the gauge group factors.
ABJ theories with $M=N+1$ and $k=2$ have the same moduli space as $U(N)_2\times U(N)_{-2}$ ABJM theories. Nevertheless we show that at least for $N=1$ and $N=2$ (and presumably for higher $N$) these two ${\mathcal N}=8$ theories are not isomorphic. We do this by comparing superconformal indices \cite{BBMR} of both theories. The indices are computed using the localization method of \cite{Kim}.
The existence of two non-isomorphic ${\mathcal N}=8$ superconformal field theories with the moduli space $({\mathbb R}^8/{\mathbb Z}_2)^N/S_N$ is unsurprising from the point of view of M-theory. Such theories should describe $N$ M2-branes on an orbifold ${\mathbb R}^8/{\mathbb Z}_2$, and it is well-known that there are exactly two such orbifolds differing by G-flux taking values in $H^4({\mathbb R}{\mathbb P}^7,{\mathbb Z})={\mathbb Z}_2$ \cite{Sethi}.
The interpretation of Bagger-Lambert-Gustavsson theories in terms of M2-branes is unclear in general. However, for low values of $k$ it has been proposed that BLG theories describe systems of two M2-branes on ${\mathbb R}^8$ or ${\mathbb R}^8/{\mathbb Z}_2$ \cite{LT,Distleretal,LP}. Such systems of M2-branes are also described by ABJM and ABJ theories \cite{ABJM,ABJ}. Thus we may reinterpret these proposals in field-theoretic terms as isomorphisms between certain BLG theories and ABJM or ABJ theories. We test these proposals by computing the superconformal indices of BLG theories and comparing them with those of ABJM and ABJ theories. Based on this comparison, we propose that the following ${\mathcal N}=8$ theories are isomorphic on the quantum level:
\begin{itemize}
\item $U(2)_1\times U(2)_{-1}$ ABJM theory and $(SU(2)_1\times SU(2)_{-1})/{\mathbb Z}_2$ BLG theory
\item $U(2)_2\times U(2)_{-2}$ ABJM theory and $SU(2)_2\times SU(2)_{-2}$ BLG theory
\item $U(3)_2\times U(2)_{-2}$ ABJ theory and $(SU(2)_4\times SU(2)_{-4})/{\mathbb Z}_2$ BLG theory
\end{itemize}
The first two of these isomorphisms have been discussed in \cite{LP}.
We provide further evidence for the first of these dualities by showing that on the quantum level $(SU(2)_1\times SU(2)_{-1})/{\mathbb Z}_2$ BLG theory has a free sector realized by monopole operators with minimal GNO charge. This sector has ${\mathcal N}=8$ supersymmetry and can be thought of as a free ${\mathcal N}=4$ hypermultiplet plus a free ${\mathcal N}=4$ twisted hypermultiplet. Thus this BLG theory has not one but two copies of ${\mathcal N}=8$ supersymmetry algebra, one acting on the free sector and one acting on the remainder. This quantum doubling of the ${\mathcal N}=8$ supercurrent multiplet is required by duality, because $U(2)_1\times U(2)_{-1}$ theory also has such a doubling on the quantum level, as well as a free sector \cite{BK}. All these peculiar properties stem from the fact that the theory of $N$ M2-branes in flat space must have a free ${\mathcal N}=8$ sector describing the center-of-mass motion. In the ``traditional'' approach to the theory of $N$ M2-branes via the $U(N)$ ${\mathcal N}=8$ super-Yang-Mills theory, this decomposition is apparent on the classical level (one can decompose all fields into trace and traceless parts which then do not interact, with the trace part being free). In the ABJM description of the same system this decomposition arises only on the quantum level \cite{BK}. For $N=2$ we also have a BLG description of the same system, and the existence of a free sector is again a quantum effect.
Superconformal index provides a simple tool for distinguishing ${\mathcal N}=8$ theories which have the same moduli space. We can apply this method to other BLG theories which do not have an obvious interpretation in terms of M2-branes. For example, as noted in \cite{LP}, $SU(2)_k\times SU(2)_{-k}$ and $(SU(2)_{2k}\times SU(2)_{-2k})/{\mathbb Z}_2$ BLG theories have the same moduli space for all $k$ and one may wonder if they are in fact isomorphic. We compare the indices of these theories for $k=1,2$ and show that they are different. We also find that for $k=1$ both BLG theories have an extra copy of the ${\mathcal N}=8$ supercurrent multiplet realized by monopole operators. This indicates that each of these theories decomposes as a product two ${\mathcal N}=8$ SCFTs which do not interact with each other. For higher $k$ there is only one copy of the ${\mathcal N}=8$ supercurrent multiplet.
This work was supported in part by the DOE grant DE-FG02-92ER40701.
\section{The moduli space}
Consider the family of ${\mathcal N}=6$ Chern-Simons-matter theories constructed by Aharony, Bergman and Jafferis \cite{ABJ}. The gauge group of such a theory is $U(M)\times U(N)$, with Chern-Simons couplings $k$ and $-k$. If we regard it as an ${\mathcal N}=2$ $d=3$ theory, then the matter consists of two chiral multiplets $A_a$, $a=1,2$ in the representation $(M,{\bar N})$ and two chiral multiplets $B_{\dot{a}},{\dot{a}}=1,2$ in the representation $({\bar M},N)$. The theory has a quartic superpotential
$$
W=\frac{2\pi}{k}\epsilon^{ab}\epsilon^{{\dot{a}}{\dot{b}}} {\rm Tr} A_a B_{\dot{a}} A_b B_{\dot{b}}
$$
which preserves $SU(2)\times SU(2)$ symmetry as well as $U(1)_R$ R-symmetry. The chiral fields $A_a$ and $B_{\dot{a}}$ transform as $({\bf 2},{\bf 1})_1$ and $({\bf 1},{\bf 2})_1$ respectively. It was shown in \cite{ABJ} that the Lagrangian of such a theory has $Spin(6)$ symmetry which contains $Spin(4)=SU(2)\times SU(2)$ and $U(1)_R$ as subgroups. This implies that the action has ${\mathcal N}=6$ superconformal symmetry, and the supercharges transform as a ${\bf 6}$ of $Spin(6)$ R-symmetry.
We wish to explore the possibility that on the quantum level some of these theories have ${\mathcal N}=8$ supersymmetry. A necessary condition for this is that at a generic point in the moduli space of vacua the theory has ${\mathcal N}=8$ supersymmetry. The moduli space can be parameterized by the expectation values of the fields $A_a$ and $B_{\dot{a}}$. Let us assume $M\geq N$ for definiteness. The superpotential is such that the expectation values can be brought to the diagonal form \cite{ABJ}:
$$
\langle {A_a}^i_j\rangle=a_{ja} \delta^i_j,\quad \langle {B_{\dot{a}}}^j_i\rangle=b_{\dot{a}}^j \delta^i_j\quad i=1,\ldots,M,\quad j=1,\ldots,N.
$$
Thus the classical moduli space is parameterized by $2N$ complex numbers $a_{ja}$ and $2N$ complex numbers $b^j_{\dot{a}}$ which together parameterize ${\mathbb C}^{4N}$. Unbroken gauge symmetry includes a $U(M-N)$ factor which acts trivially on the moduli, as well as a discrete subgroup of $U(N)$. The low-energy effective action for the $U(M-N)$ gauge field is the Chern-Simons action at level $k'=k-{\rm sign}(k)(M-N)$. Thus along the moduli space the theory factorizes into a free theory describing the moduli and the topological $U(M-N)$ Chern-Simons theory at level $k'$. Note that for $M-N>|k|$ the sign of $k'$ is different from that of $k$. This has been interpreted in \cite{ABJ} as a signal that for $M-N>|k|$ supersymmetry is spontaneously broken on the quantum level, and that the classical moduli space is lifted. Therefore from now on we will assume $M-N\leq |k|$.
The putative ${\mathcal N}=8$ supersymmetry algebra must act trivially on the topological sector, so we need to analyze for which $M,N,$ and $k$ the free theory of the moduli has ${\mathcal N}=8$ supersymmetry. This theory is a supersymmetric sigma-model whose target space is the quotient of ${\mathbb C}^N$ by the discrete subgroup of $U(N)$ which preserves the diagonal form of the matrices $A_a$ and $B_b$. This discrete subgroup is a semi-direct product of the permutation group $S_N$ and the ${\mathbb Z}_k^N$ subgroup of the maximal torus of $U(N)$ \cite{ABJ}. Thus the target space is $({\mathbb C}^4/{\mathbb Z}_k)^N/S_N$. The action of ${\mathbb Z}_k$ on ${\mathbb C}^4$ is given by
$$
z_i\mapsto \eta z_i,\quad i=1,\ldots,4,\quad \eta^k=1.
$$
Here $z_{1,2}$ are identified with $a_{ia}$, $a=1,2$, while $z_{3,4}$ are identified with $b^j_{\dot{a}}$, ${\dot{a}}=1,2$.
Free ${\mathcal N}=2$ sigma-model with target ${\mathbb C}^4\simeq{\mathbb R}^8$ has ${\mathcal N}=8$ supersymmetry and $Spin(8)$ R-symmetry. Supercharges transform as ${\bf 8}_c$ of $Spin(8)$, while the moduli parameterizing ${\mathbb R}^8$ transform as ${\bf 8}_v$. The above ${\mathbb Z}_k$ action on ${\bf 8}_v$ factors through the $Spin(8)$ action on the same space, and for $|k|
>2$ its commutant with ${\mathbb Z}_k$ is $U(4)$. ${\mathbb Z}_k$ itself can be identified with the ${\mathbb Z}_k$ subgroup of the $U(1)$ subgroup of $U(4)$ consisting of scalar matrices. Under the $U(4)$ subgroup ${\bf 8}_c$ decomposes as ${\bf 6}_0+{\bf 1}_2+{\bf 1}_{-2}$, and therefore for $|k|>2$ only ${\bf 6}_0$ is ${\mathbb Z}_k$-invariant. Thus for $|k|>2$ the moduli theory has only ${\mathcal N}=6$ supersymmetry.
For $|k|=1,2$ the ${\mathbb Z}_k$ subgroup acts trivially on ${\bf 8}_c$, and therefore these two cases are the only ones for which the theory of moduli has ${\mathcal N}=8$ supersymmetry. In view of the above, for $|k|=1$ we may assume that $M-N\leq 1$ while for $|k|=2$ we may assume $M-N\leq 2$.
For $N=M$ and $|k|=1,2$ it has been argued in \cite{ABJM} that the full theory has ${\mathcal N}=8$ supersymmetry on the quantum level. The hidden symmetry currents are realized by monopole operators. This proposal has been proved using controlled deformation to weak coupling \cite{BK}; for other approaches see \cite{Gus,BKK,O}.
It remains to consider the case $0<M-N\leq |k|$ for $|k|=1,2$. Some of these theories are dual to the ${\mathcal N}=8$ ABJM theories with $N=M$ and $k=1,2$. Indeed, it has been argued in \cite{ABJ} that for $M-N\leq |k|$ the theory with gauge group $U(M)_k\times U(N)_{-k}$ is dual to the theory with gauge group $U(2N-M+|k|)_{-k}\times U(N)_k$. We will call it the ABJ duality.\footnote{Alternatively, the ABJ duality follows from the ${\mathcal N}=3$ version of the Giveon-Kutasov duality applied to the $U(M)$ factor \cite{KWY3}. One can also verify that the $S^3$ partition functions of the dual ABJ theories agree \cite{KWY3}.} It maps $M-N$ to $|k|-(M-N)$ and $k$ to $-k$. Hence the ABJ theory with gauge group $U(N+1)_1\times U(N)_{-1}$ is dual to the ABJ theory with gauge group $U(N)_{-1}\times U(N)_1$. Similarly, the ABJ theory with gauge group $U(N+2)_2\times U(N)_{-2}$ is dual to the ABJ theory with gauge group $U(N)_{-2}\times U(N)_2$.
The only remaining case is the ABJ theory with gauge group $U(N+1)_2\times U(N)_{- 2}$ and its parity-reversal. Each theory in this family is self-dual under the ABJ duality combined with parity. Put differently, the combination of naive parity and ABJ duality is a symmetry for all $N$, i.e. while these theories are not parity-invariant on the classical level, they have hidden parity on the quantum level. In the remainder of this paper we will argue that this family of theories in fact has hidden ${\mathcal N}=8$ supersymmetry and is not isomorphic to any other known family of ${\mathcal N}=8$ $d=3$ SCFTs. We will also present evidence that certain BLG theories with $k=1,2$ are isomorphic to ${\mathcal N}=8$ ABJ and ABJM theories for $N=1,2$.
\section{Monopole operators and hidden ${\mathcal N}=8$ supersymmetry}
In this section we will show that the ABJ theory with gauge group $U(N+1)_2\times U(N)_{-2}$ has hidden ${\mathcal N}=8$ supersymmetry. We will follow the method of \cite{BK} to which the reader is referred for details. The main step is to demonstrate the presence of protected scalars with scaling dimension $\Delta=1$ which live in the representation ${\bf 10}_{-1}$ of the manifest symmetry group $Spin(6)\times U(1)_T$. Here $U(1)_T$ is the topological symmetry of the ABJ theory whose current
$$
J^\mu=-\frac{k}{16\pi} \epsilon^{\mu\nu\rho}\left ({\rm Tr} F_{\nu\rho}+{\rm Tr} \tilde F_{\nu\rho}\right).
$$
is conserved off-shell. Once the existence of these scalars is established, acting on them with two manifest supercharges produces conserved currents with $\Delta=2$ transforming in the representation ${\bf 6}_{-1}$ of $Spin(6)\times U(1)_T$. Since conserved currents in any field theory form a Lie algebra, these currents together with their Hermitian-conjugate currents, $Spin(6)$ currents and the $U(1)_T$ current must combine into an adjoint of some Lie algebra containing $Spin(6)\times U(1)_T$ Lie algebra as a subalgebra. The unique possibility for such a Lie algebra is $Spin(8)$, which implies that the theory has ${\mathcal N}=8$ supersymmetry.
The existence of $\Delta=1$ scalars transforming in ${\bf 10}_{-1}$ is established using a controlled deformation of the theory compactified on $S^2$ to weak coupling. This deformation preserves $Spin(4)\times U(1)_R$ subgroup of $Spin(6)$ as well as $U(1)_T$. Decomposing ${\bf 10}_{-1}$ with respect to this subgroup, we find that it contains BPS scalars in $({\bf 3},{\bf 1})_{1,-1}$ of $Spin(4)\times U(1)_R\times U(1)_T$ and anti-BPS scalars in $({\bf 1},{\bf 3})_{-1,-1}$. Such BPS scalars cannot disappear as one changes the coupling (see appendix A for a detailed argument), so it is sufficient to demonstrate the presence of BPS scalars at extremely weak coupling. Note that the scaling dimension $\Delta$ of an operator is now reinterpreted as the energy of a state on $S^2$.
The BPS scalars we are looking for have nonzero $U(1)_T$ charge and therefore are monopole operators \cite{BKW}. At weak coupling monopole operators in ABJ theories are labeled by GNO ``charges'' $(m_1,\ldots,m_M)$ and $({\tilde m}_1,\ldots,{\tilde m}_N)$. GNO charges label spherically symmetric magnetic fields on $S^2$ and are defined up to the action of the Weyl group of $U(M)\times U(N)$ \cite{GNO}. They do not correspond to conserved currents and can be defined only at weak coupling. Their sum however is related to the $U(1)_T$ charge:
$$
Q_T=-\frac{k}{4}\left(\sum m_i+\sum \tilde m_i\right).
$$
Equations of motion of the ABJ theory imply that $\sum m_i=\sum \tilde m_i$, so $Q_T$ is integral for even $k$ but may be half-integral for odd $k$. We are interested in the case $Q_T=-1$, $k=2$, which implies
$$
\sum m_i=\sum \tilde m_i=1.
$$
Consider a bare BPS monopole, i.e. the vacuum state, with GNO charges $m_1=\tilde m_1=1$ and all other GNO charges vanishing. This state has $\Delta=0$ but because of Chern-Simons terms it is not gauge-invariant (does not satisfy the Gauss law constraint). One can construct a gauge-invariant state by acting on the bare BPS monopole with two creation operators corresponding to the fields ${\bar A}^{1\tilde 1}_a$, $a=1,2$. These states are completely analogous to the BPS scalars for the $U(N)_2\times U(N)_{-2}$ ABJM theory constructed in \cite{BK} (see eq. (13) in that paper). The resulting multiplet of states transforms as $({\bf 3},{\bf 1})_{1,-1}$ of $Spin(4)\times U(1)_R\times U(1)_T$. It also has $\Delta=1$ and zero spin, since the creation operators for the field ${\bar A}$ with lowest energy have $\Delta=1/2$ and zero spin.
Similarly, by starting from an anti-BPS bare monopole with the same GNO charges and acting on it with two creation operators belonging to the fields $B_a^{1\tilde 1}$ we obtain anti-BPS scalars which transform in $({\bf 1},{\bf 3})_{-1,-1}$. One can also check that no other GNO charges give rise to BPS scalars with $\Delta=1$. In view of the above discussion this implies that the $U(N+1)_2\times U(N)_{-2}$ ABJ theory has hidden ${\mathcal N}=8$ supersymmetry.
\section{Superconformal index and comparison with other ${\mathcal N}=8$ theories}
One may question if $U(N+1)_2\times U(N)_{-2}$ ABJ theories are genuinely distinct from other known ${\mathcal N}=8$ $d=3$ theories. The moduli space of such a theory is
$({\mathbb C}^4/{\mathbb Z}_2)^N/S_N$, which is exactly the same as the moduli space of the $U(N)_2\times U(N)_{-2}$ ABJ theory. They differ in that along the moduli space the former theory has an extra topological sector described by $U(1)$ Chern-Simons theory at level $1$. The latter theory is not quite trivial \cite{Wittensl2}, but it is very close to being trivial; for example, it does not admit any nontrivial local or loop observables. In any case, one could conjecture that even at the origin of the moduli space the two ${\mathcal N}=8$ $d=3$ theories differ only by this decoupled topological sector. Some evidence in support of this conjecture is that BPS scalars in the two theories are in 1-1 correspondence, as we have seen in the previous section.
Fortunately, in the last few years there has been substantial progress in understanding superconformal $d=3$ gauge theories which allows us to compute many quantities exactly. One such quantity is the partition function on $S^3$ \cite{KWY1}; another one is the superconformal index on $S^2\times S^1$ \cite{BBMR,Kim}. The superconformal index receives contribution from BPS scalars as well as from other protected states with nonzero spin. In what follows we will compute the index for several low values of $N$ and verify that it is different for the two families of ${\mathcal N}=8$ theories. The perturbative contribution to the superconformal index for ABJM theories has been computed in \cite{BM}; the contributions of sectors with a nontrivial GNO charge has been determined in \cite{Kim}. We will follow the approach of \cite{Kim}.
Bagger and Lambert \cite{BL} and Gustavsson \cite{G} constructed another infinite family of ${\mathcal N}=8$ $d=3$ superconformal Chern-Simons-matter theories with gauge group $SU(2)\times SU(2)$ and matter in the bifundamental representation. More precisely, as emphasized in \cite{ABJM,LP}, there are two versions of BLG theories which have gauge groups $SU(2)_k \times SU(2)_{-k}$ or $(SU(2)_k\times SU(2)_{-k})/{\mathbb Z}_2$ where $k$ is an arbitrary natural number. The moduli space is $({\mathbb C}^4\times{\mathbb C}^4)/{\mathbf D}_{2k}$ and $({\mathbb C}^4\times{\mathbb C}^4)/{\mathbf D}_k$ respectively, where ${\mathbf D}_k$ is the dihedral group of order $2k$ \cite{LT,Distleretal,LP}. For large enough $k$ the moduli space is different from the moduli space of ABJ theories and so BLG theories cannot be isomorphic to any of them. However, for low values of $k$ there are some coincidences between moduli spaces which suggest that perhaps some of BLG theories are isomorphic to ABJ theories.
One such case is $k=1$ and $G=(SU(2)\times SU(2))/{\mathbb Z}_2$. The moduli space is $({\mathbb C}^4\times{\mathbb C}^4)/{\mathbb Z}_2$ where ${\mathbb Z}_2$ exchanges the two ${\mathbb C}^4$ factors. It is natural to conjecture that this theory is isomorphic to $U(2)_1\times U(2)_{-1}$ ABJM theory. A derivation of this equivalence was proposed in \cite{LP}. Another special case is $k=2$ and $G=SU(2)\times SU(2)$. In that case the moduli space is isomorphic to $({\mathbb C}^4/{\mathbb Z}_2\times{\mathbb C}^4/{\mathbb Z}_2)/{\mathbb Z}_2$, where the first two ${\mathbb Z}_2$ factors reflect the coordinates on the two copies of ${\mathbb C}^4$, while the third one exchanges them \cite{LT,Distleretal,LP}. This is the same moduli space as that of $U(2)_2\times U(2)_{-2}$ ABJM theory and $U(3)_2\times U(2)_{-2}$ ABJ theory. It was conjectured in \cite{LP} that this BLG theory is isomorphic to the $U(2)_2\times U(2)_{-2}$ ABJM theory. Finally, one can take $k=4$ and $G=(SU(2)\times SU(2))/{\mathbb Z}_2$. The moduli space is the same as in the previous case, so one could conjecture that this BLG theory is isomorphic to either the $U(2)_2\times U(2)_{-2}$ ABJM theory or the $U(3)_2\times U(2)_{-2}$ ABJ theory.
Below we will first of all compute the superconformal index for the $U(N)_2\times U(N)_{-2}$ ABJM theories and $U(N+1)_2\times U(N)_{-2}$ ABJ theories for $N=1,2$ and verify that although these theories have the same moduli space, they have different superconformal indices and therefore are not isomorphic. We will also compute the index for the special BLG theories with low values of $k$ discussed above and test the proposed dualities with the ABJM and ABJ theories. We will see that certain BLG theories have an additional copy of the ${\mathcal N}=8$ supercurrent multiplet which is realized by monopole operators. In some cases this is predicted by dualities.
\subsection{${\mathcal N}=8$ ABJM vs. ${\mathcal N}=8$ ABJ theories}
The superconformal index for a supersymmetric gauge theory on $S^2\times{\mathbb R}$ is defined as
\begin{align}
{\cal I}(x, z_i)=Tr[(-1)^Fx^{E+j_3}\prod_{i}z_i^{F_i}]\label{ind}
\end{align}
where $F$ is the fermion number, $E$ is the energy, $j_3$ is the third component of spin and $F_i$ are flavor symmetry charges. The index receives contributions only from states satisfying $\{Q,Q^\dagger\}=E-r-j_3=0$, where $Q$ is one of the 32 supercharges and $r$ is a $U(1)$ $R$-charge. For details the reader is referred to \cite{BBMR,Kim}.
The localization method \cite{Kim} enables one to express the index in a simple form\footnote{The formula is written for the case of zero anomalous dimensions of all fields which is true for all theories with at least ${\mathcal N}=3$ supersymmetry.}
\begin{align}
{\cal I}(x, z_i)=\sum_{\{n_i\}}\int[da]_{\{n_i\}}x^{E_0(n_i)}e^{S_{CS}^0(n_i,a_i)}exp(\sum_{m=1}^\infty f(x^m, z_i^m, ma_i))
\end{align}
where the sum is over GNO charges, the integral whose measure depends on GNO charges is over a maximal torus of the gauge group, $E_0(n_i)$ is the energy of a bare monopole with GNO charges $\{n_i\}$, $S_{CS}^0(n_i)$ is effectively the weight of the bare monopole with respect to the gauge group and the function $f$ depends on the content of vector multiplets and hypermultiplets. For details see \cite{Kim}.
We computed the indices for the $U(2)_2\times U(1)_{-2}$ and $U(1)_2\times U(1)_{-2}$ theories up to the sixth order in $x$ and found the following pattern. In each topological sector the indices agree at the leading order in $x$ as a consequence of the identical spectra of BPS scalars of the lowest dimension. However, next-to-leading terms are different which signals nonequivalence of these theories. We summarize our results in tables 1 and 2 in Appendix B.
It is possible to single out contributions from different topological sectors by treating topological $U(1)_T$ symmetry as a flavor symmetry and introducing a new variable $z$ into the index. The result is a double expansion in $x$ and $z$ with powers of $z$ multiplying contributions of the appropriate topological charge . Alternatively, one can restrict summation over all GNO charges to those giving the desired topological charge. We used the second type of calculation.
We also compared the indices for the ABJ theory $U(3)_2\times U(2)_{-2}$ and the ABJM theory $U(2)_2\times U(2)_{-2}$ up to the fourth order in $x$. The contributions from different GNO sectors are summarized in tables 3 and 4 in Appendix B. Note that we count the contributions from the topological sectors $T\ge1$ twice because there is an identical contribution from the sectors with opposite topological charges. Starting at order $x^3$ the indices disagree, which means that these two ${\mathcal N}=8$ theories, despite having the same moduli space, are not equivalent.
\subsection{Comparison with BLG theories}
There are two BLG theories which have the same moduli space as $U(2)_2\times U(2)_{-2}$ ABJM and $U(3)_2\times U(2)_{-2}$ ABJ theories. They have gauge groups $SU(2)_2\times SU(2)_{-2}$ and $(SU(2)_4\times SU(2)_{-4})/{\mathbb Z}_2$. It is natural to conjecture that these four theories are pairwise isomorphic. Indeed, the moduli space is $({\mathbb C}^4/{\mathbb Z}_2\times{\mathbb C}^4/{\mathbb Z}_2)/{\mathbb Z}_2$ in all four cases, suggesting that all these theories describe two M2-branes on an ${\mathbb R}^8/{\mathbb Z}_2$ orbifold. It is well-known that there are two distinct ${\mathbb R}^8/{\mathbb Z}_2$ orbifolds in M-theory \cite{Sethi}, which means that there should be only two nonisomorphic ${\mathcal N}=8$ theories with this moduli space.
Comparison of the indices of the $U(2)_2\times U(2)_{-2}$ ABJM theory and the $SU(2)_2\times SU(2)_{-2}$ BLG theory (see Table 5) reveals their agreement up to the fourth order in $x$. Thus we conjecture that the two theories are equivalent.
This conjecture can be checked further by comparing contributions to the indices from individual topological sectors on the ABJM side and sectors parametrized by the corresponding $U(1)$ charge on the BLG side. Recall that the topological charge $Q_T$ on the ABJM side is a charge of a $U(1)$ subgroup of the $Spin(8)$ $R$-symmetry group. The commutant of this subgroup is $Spin(6)$ R-symmetry visible already on the classical level. Furthermore, the supercharge used in the deformation and the definition of the index is charged under a $U(1)$ subgroup of this $Spin(6)$. On the BLG side, the whole $Spin(8)$ R-symmetry is visible on the classical level. Recall that one can think of the BLG theory as a ${\mathcal N}=2$ field theory with gauge group $SU(2)\times SU(2)$ and four chiral multiplets in the bifundamental representation. In this description, there is a manifest $SU(4)=Spin(6)$ symmetry under which the four chiral superfields transform as ${\bf 4}$. The commutant of this $Spin(6)$ symmetry is $U(1)_R$ symmetry with respect to which all four chiral superfields have charge $1/2$ and the supercharge has charge $1$. The topological charge $Q_T$ on the ABJM side corresponds to the charge of a $U(1)$ subgroup of $Spin(6)$ which we denote as $U(1)_t$\footnote{We now adopt the notation $T\equiv\sum_im_i$ for the topological charge and normalize the $U(1)_t$ charge of fundamental scalars of the BLG theories to $\pm1$ for notational convenience. The $U(1)_R$ charges are not shown in what follows.}. Thus we should compare the ABJM index in a particular topological sector with the BLG index in a sector with a particular $U(1)_t$ charge. The four chiral fields of the BLG theory decompose as ${\bf 4}={\bf 2}_{1}+{\bf 2'}_{-1}$ under $U(1)_t\times Spin(4)$.
To keep track of $U(1)_t$ charges we introduce a new variable $z$ in accordance with (\ref{ind}). To the fourth order in $x$ only the $(\ket{0}\ket{0},\ket{1}\ket{1},\ket{2}\ket{2})$ GNO charges contribute. The two-variable index is
\begin{align}
& {\cal I}_{BLG,k=2}(x,z)=1+4x+21x^2+32x^3+53x^4+z^2(3x+16x^2+36x^3+48x^4)+\nonumber\\
& z^4(11x^2+36x^3+54x^4)+z^6(22x^3+64x^4)+45x^4z^8+ z^{-2}(3x+16x^2+36x^3+48x^4)+\nonumber\\
&z^{-4}(11x^2+36x^3+54x^4)+z^{-6}(22x^3+64x^4)+45x^4z^{-8}+O(x^5).
\end{align}
This is in a complete agreement with the index for the $U(2)_2\times U(2)_{-2}$ ABJM theory.
Similarly, we can compute the two-variable index for the $(SU(2)_4\times SU(2)_{-4})/{\mathbb Z}_2$ BLG theory. The difference compared to the $SU(2)\times SU(2)$ case is that the GNO charges are allowed to be half-integral, but their difference is required to be integral. The contributions of individual GNO charges are summarized in Table 6. We see that the total index agrees with that of the $U(3)_2\times U(2)_{-2}$ ABJ theory at least up to the fourth order in $x$. The two-variable index for this BLG theory is given by
\begin{align}
& {\cal I}'_{BLG}(x,z)=1+4x+21x^2+36x^3+39x^4+z^2(3x+16x^2+39x^3+40x^4)+\nonumber\\
& z^4(11x^2+36x^3+56x^4)+z^6(22x^3+64x^4)+45z^8x^4+z^{-2}(3x+16x^2+39x^3+40x^4)+\nonumber\\
& z^{-4}(11x^2+36x^3+56x^4)+z^{-6}(22x^3+64x^4)+45z^{-8}x^4+O(x^5)
\end{align}
and agrees with the two-variable index of the $U(3)_2\times U(2)_{-2}$ ABJ theory.
Lambert and Papageorgakis \cite{LP} argued that the $(SU(2)_1\times SU(2)_{-1})/{\mathbb Z}_2$ BLG theory is isomorphic to the $U(2)_1\times U(2)_{-1}$ ABJM theory. We can test this proposal in the same way by comparing the two-variable superconformal indices of the two theories. We find that they agree up to at least the fourth order in $x$. The contributions from different GNO charges are written down in Tables 7 and 8. They happen to match in each GNO sector separately. For a fixed topological charge on the ABJM side and the corresponding value of the $U(1)_t$ charge on the BLG side which manifests itself in the index as a power of $z$, the contribution to the index comes from a sum over different GNO charges, and the two sums happen to coincide term by term. For example, in the topological sector $T=1$ on the ABJM side the contribution from the GNO charge $\ket{n,1-n}\ket{n,1-n}$ equals the contribution from the GNO charge $\ket{n-1/2}\ket{n-1/2}$ with the first power of $z$ on the BLG side.
The index makes apparent a peculiar feature of these two theories: they have twice the number of BPS scalars needed to enhance supersymmetry from ${\cal N}=6$ to ${\cal N}=8$. The first set of scalars has vanishing GNO charge. The corresponding contribution to the index is $\Delta{\cal I}=4x+3xz^2+3xz^{-2}$. It represents the decomposition ${\bf 10}={\bf 4}_0+{\bf 3}_2+{\bf 3}_{-2}$ under $U(1)_t\times Spin(4)\subset Spin(6)$. The corresponding operators are gauge-invariant bilinear combinations of four chiral superfields present in the BLG model. The second set of ten BPS scalars comes from the GNO charge $\ket{1}\ket{1}$ and makes an identical contribution to the index. Ten BPS states are obtained by acting with ten scalar bilinears on the bare monopole to form gauge-invariant states $Q^{(i}Q^{j)}\ket{1}\ket{1}$. Here $Q^i$ is an off-diagonal component of the $i^{\rm th}$ complex scalar, $i=1,\ldots,4$. Among these ten states there are representations ${\bf (3,1)}_{1,-1}+{\bf (1,3)}_{1,1}$ of $Spin(4)\times U(1)_R\times U(1)_t$ with the normalization of the $U(1)_t$ charge as on page 4. Together with their Hermitian-conjugates, these BPS scalars lead to supersymmetry enhancement as in \cite{BK}.
The existence of two copies of the ${\mathcal N}=8$ supersymmetry algebra for the $U(2)_1\times U(2)_{-1}$ ABJM theory was noted in \cite{BK}. It was shown there that the extra copy arises because the theory has a free sector with ${\mathcal N}=8$ supersymmetry realized by monopole operators. The same is true about the $(SU(2)_1\times SU(2)_{-1})/{\mathbb Z}_2$ BLG theory, giving further support for the duality. The sector with the GNO charge $\ket{1/2}\ket{1/2}$ contains four gauge-invariant BPS scalars $Q^{i}\ket{1/2}\ket{1/2}$ with energy $\Delta=1/2$ whose contribution to the index is $\Delta{\cal I'}=2x^{1/2}z+2x^{-1/2}z$. This expression corresponds to the decomposition ${\bf 4}={\bf 2}_1+{\bf 2'}_{-1}$ under $U(1)_t\times Spin(4)\subset Spin(6)$. By virtue of state-operator correspondence these states correspond to four free fields with conformal dimension $\Delta=1/2$. Their bilinear combinations give rise to ten BPS scalars with GNO charge $\ket{1}\ket{1}$ discussed above. This is in a complete agreement with the structure of the $U(2)_1\times U(2)_{-1}$ ABJM theory explored in \cite{BK}.
We can also use superconformal index to test whether certain BLG theories with identical moduli spaces are isomorphic on the quantum level. It has been noted in \cite{LP} that the moduli spaces of $SU(2)_k\times SU(2)_{-k}$ and $(SU(2)_{2k}\times SU(2)_{-2k})/{\mathbb Z}_2$ BLG theories are the same (they are both given by $({\mathbb C}^4\times{\mathbb C}^4)/{\mathbf D}_{2k}$. We have seen above that for $k=2$ these two theories are not isomorphic. We also computed the index for $k=1$ and found that the indices disagree already at the second order in $x$ (Tables 9 and 10), so the theories are not equivalent. Examining BPS scalars, we find that neither of these theories has a free sector, but they both have two copies of the ${\mathcal N}=8$ supercurrent multiplet. One copy is visible on the classical level, while the BPS scalars of the other copy carry GNO charges, so it is intrinsically quantum-mechanical in origin. The presence of the second copy of ${\mathcal N}=8$ superalgebra indicates that on the quantum level both of these theories decompose into two ${\mathcal N}=8$ SCFTs which do not interact with each other. This phenomenon does not occur for higher $k$.
\section*{Appendix A}
Our method of detecting hidden supersymmetry is based on deforming the theory to weak coupling and analyzing the spectrum of BPS scalars.
In this appendix we provide a sufficient condition for BPS scalars to be protected as one deforms the coupling from weak to strong. In general, a local operator (or the corresponding state in the radial quantization) which lives in a short representation of the superconformal algebra can pair up with another short multiplet to form a long multiplet; quantum number of a long multiplet can change continuously as one deforms the coupling. We would like to show that this cannot happen for the cases of interest to us.
The kind of short multiplet we are interested in has a BPS scalar among its primaries. In the radial quantization such a state has energy $\Delta$ equal to its $U(1)_R$ charge $r$. To form a long multiplet there must be a short multiplet containing a spinor with energy $\Delta'=\Delta\pm 1/2$ and $R$-charge $r=r'\pm 1$. The option with $\Delta'=\Delta+1/2$ and $r'=r+1$ is ruled out by unitarity constraints \cite{BBMR}. These constraints also specify the short multiplet with the spinor. This is a so-called ``regular short multiplet" \cite{BBMR} with a scalar $\Delta''=\Delta-1$, $r''=r-2$ as the superconformal primary state satisfying $\Delta''=r''+1$. The zero-norm state is also a scalar, appears on the second level and has the quantum numbers of a BPS scalar $\Delta=r$. The spinor itself is on the first level.
We conclude that a necessary condition for a BPS scalar with quantum numbers $\Delta=r$ to pair up into a long multiplet and flow away is the existence of a "regular short multiplet" with quantum numbers $\Delta''=\Delta-1$ and $r''=r-2$.
In the particular case of a $U(N+1)_k\times U(N)_{-k}$ ABJ theory and $\Delta=1$ such ``regular short multiplets" do not exist because $\Delta''=0$ and all physical states have $\Delta\geq 1$.
\newpage
\section*{Appendix B: Superconformal indices for ${\mathcal N}=8$ ABJM, ABJ and BLG theories}
\begin{longtable}{|l|l|}
\hline
GNO charges & Index contribution\\
\hline
$T=0$ & $1+4x+2x^2+15x^4-16x^5+11x^6$\\
\hline
$\ket{0,0}\ket{0}$ & $1+4x+2x^2+15x^4-16x^5+2x^6$\\
$\ket{1,-1}\ket{0}$ & $9x^6$\\
\hline
$T=1$ & $3x+x^2-4x^3+20x^4-32x^5+24x^6$\\
\hline
$\ket{1,0}\ket{1}$ & $3x+x^2-4x^3+20x^4-32x^5+24x^6$\\
\hline
$T=2$ & $5x^2+4x^3-5x^4+4x^5-4x^6$\\
\hline
$\ket{2,0}\ket{2}$ & $5x^2+4x^3-5x^4+4x^5-4x^6$\\
\hline
$T=3$ & $7x^3+4x^4+x^6$\\
\hline
$\ket{3,0}\ket{3}$ & $7x^3+4x^4+x^6$\\
\hline
$T=4$ & $9x^4+4x^5$\\
\hline
$\ket{4,0}\ket{4}$ & $9x^4+4x^5$\\
\hline
$T=5$ & $11x^5+4x^6$\\
\hline
$\ket{5,0}\ket{5}$ & $11x^5+4x^6$\\
\hline
$T=6$ & $13x^6$\\
\hline
$\ket{6,0}\ket{6}$ & $13x^6$\\
\hline
total & $1+10x+14x^2+14x^3+71x^4-42x^5+39x^6$\\
\hline
\caption{$U(2)_2\times U(1)_{-2}$. $T$ stands for the topological charge.}
\end{longtable}
\begin{longtable}{|l|l|}
\hline
Topological charge & Index contribution\\
\hline
$T=0$ & $1+4x+x^2+4x^3+7x^4-12x^5+26x^6$\\
\hline
$T=1$& $3x+4x^2+8x^4-4x^5+8x^6$\\
\hline
$T=2$& $5x^2+4x^3=8x^5-4x^6$\\
\hline
$T=3$& $7x^3+4x^4+8x^6$\\
\hline
$T=4$ & $9x^4+4x^5$\\
\hline
$T=5$ & $11x^5+4x^6$\\
\hline
$T=6$ & $13x^6$\\
\hline
total & $1+10x+19x^2+26x^3+49x^4+26x^5+92x^6$\\
\hline
\caption{$U(1)_2\times U(1)_{-2}$}
\end{longtable}
\begin{longtable}{|l|l|}
\hline
GNO charges & Index contribution\\
\hline
$T=0$ & $1+4x+21x^2+36x^3+39x^4$\\
\hline
$\ket{0,0,0}\ket{0,0}$ & $1+4x+12x^2+12x^3+5x^4$\\
$\ket{1,0,-1}\ket{1,-1}$ & $9x^2+24x^3+10x^4$\\
$\ket{2,0,-2}\ket{2,-2}$ & $25x^4$\\
$\ket{1,0,-1}\ket{0,0}$ & $-x^4$\\
\hline
$T=1$ & $3x+16x^2+39x^3+40x^4$\\
\hline
$\ket{1,0,0}\ket{1,0}$ & $3x+16x^2+24x^3+8x^4$\\
$\ket{2,0,-1}\ket{2,-1}$ & $15x^3+32x^4$\\
\hline
$T=2$ & $11x^2+36x^3+56x^4$\\
\hline
$\ket{1,1,0}\ket{1,1}$ & $6x^2+12x^3+9x^4$\\
$\ket{2,0,0}\ket{2,0}$ & $5x^2+24x^3+26x^4$\\
$\ket{3,0,-1}\ket{3,-1}$ & $21x^4$\\
\hline
$T=3$ & $22x^3+64x^4$\\
\hline
$\ket{2,1,0}\ket{2,1}$ & $15x^3+32x^4$\\
$\ket{3,0,0}\ket{3,0}$ & $7x^3+32x^2$\\
\hline
$T=4$ & $45x^4$\\
\hline
$\ket{2,2,0}\ket{2,2}$ & $15x^4$\\
$\ket{3,1,0}\ket{3,1}$ & $21x^4$\\
$\ket{4,0,0}\ket{4,0}$ & $9x^4$\\
\hline
total & $1+10x+75x^2+230x^3+445x^4$\\
\hline
\caption{$U(3)_2\times U(2)_{-2}$.
$T$ stands for the topological charge.}
\end{longtable}
\begin{longtable}{|l|l|}
\hline
GNO charges & Index contribution\\
\hline
$T=0$ & $1+4x+21x^2+32x^3+53x^4$ \\
\hline
$\ket{0,0}\ket{0,0}$ & $1+4x+12x^2+8x^3+12x^4$\\
$\ket{1,-1}\ket{1,-1}$ & $9x^2+24x^3+16x^4$\\
$\ket{2,-2}\ket{2,-2}$ & $25x^4$\\
\hline
$T=1$ & $3x+16x^2+36x^3+48x^4$\\
\hline
$\ket{1,0}\ket{1,0}$ & $3x+16x^2+21x^3+16x^4$\\
$\ket{2,-1}\ket{2,-1}$ & $15x^3+32x^4$\\
\hline
$T=2$ & $11x^2+36x^3+54x^4$\\
\hline
$\ket{1,1}\ket{1,1}$ & $6x^2+12x^3+12x^4$\\
$\ket{2,0}\ket{2,0}$ & $5x^2+24x^3+21x^4$\\
$\ket{3,-1}\ket{3,-1}$ & $21x^4$\\
\hline
$T=3$ & $22x^3+64x^4$\\
\hline
$\ket{2,1}\ket{2,1}$ & $15x^3+32x^4$\\
$\ket{3,0}\ket{3,0}$ & $7x^3+32x^4$\\
\hline
$T=4$ & $45x^4$\\
\hline
$\ket{2,2}\ket{2,2}$ & $15x^4$\\
$\ket{3,1}\ket{3,1}$ & $21x^4$\\
$\ket{4,0}\ket{4,0}$ & $9x^4$\\
\hline
total & $1+10x+75x^2+220x^3+475x^4$\\
\hline
\caption{$U(2)_2\times U(2)_{-2}$.
$T$ stands for the topological charge.}
\end{longtable}
\begin{longtable}{|l|l|}
\hline
GNO charges & Index contribution\\
\hline
$\ket{0}\ket{0}$ & $1+10x+40x^2+76x^3+114x^4$\\
$\ket{1}\ket{1}$ & $35x^2+144x^3+196x^4$\\
$\ket{2}\ket{2}$ & $165x^4$\\
\hline
total & $1+10x+75x^2+220x^3+475x^4$\\
\hline
\caption{$SU(2)_2\times SU(2)_{-2}$}
\end{longtable}
\begin{longtable}{|l|l|}
\hline
GNO charges & Index contribution\\
\hline
$\ket{0}\ket{0}$ & $1+4x+12x^2+8x^3+12x^4+$\\
& $z^2(3x+8x^2+12x^3+8x^4)+z^{-2}(3x+8x^2+12x^3+8x^4)+$\\
& $z^4(6x^2+12x^3+12x^4)+z^{-4}(6x^2+12x^3+12x^4)+$\\
& $z^6(10x^3+16x^4)+z^{-6}(10x^3+16x^4)+15z^8x^4+15z^{-8}x^4$\\
\hline
$\ket{1/2}\ket{1/2}$ & $9x^2+28x^3+2x^4+$\\
& $z^2(8x^2+27x^3+8x^4)+z^{-2}(8x^2+27x^3+8x^4)+$\\
& $z^4(5x^2+24x^3+23x^4)+z^{-4}(5x^2+24x^3+23x^4)+$\\
& $z^6(12x^3+32x^4)+z^{-6}(12x^3+32x^4)+21z^8x^4+21z^{-8}x^4$\\
\hline
$\ket{1}\ket{1}$ & $25x^4+24z^2x^4+24z^{-2}x^4+24z^4x^4+24z^{-4}x^4+16z^6x^4+16z^{-6}x^4+9z^8x^4+9z^{-8}x^4$\\
\hline
\caption{$(SU(2)_4\times SU(2)_{-4})/{\mathbb Z}_2$}
\end{longtable}
\begin{longtable}{|l|l|}
\hline
GNO charges & Index contribution\\
\hline
$\ket{0}\ket{0}$ & $1+4x+12x^2+8x^3+12x^4+$\\
& $z^2(3x+8x^2+12x^3+8x^4)+z^{-2}(3x+8x^2+12x^3+8x^4)+$\\
& $z^4(6x^2+12x^3+12x^4)+z^{-4}(6x^2+12x^3+12x^4)+$\\
& $z^6(10x^3+16x^4)+z^{-6}(10x^3+16x^4)+15z^4x^4+15z^{-8}x^4$\\
\hline
$\ket{1/2}\ket{1/2}$ & $2z(x^{\frac12}+6x^{\frac32}+10x^{\frac52}+7x^{\frac72})+2z^{-1}(x^{\frac12}+6x^{\frac32}+10x^{\frac52}+7x^{\frac72})+$\\
& $2z^3(3x^{\frac32}+10x^{\frac52}+9x^{\frac72})+2z^{-3}(3x^{\frac32}+10x^{\frac52}+9x^{\frac72})+$\\
& $2z^5(6x^{\frac52}+14x^{\frac72})+2z^{-5}(6x^{\frac52}+14x^{\frac72})$\\
\hline
$\ket{1}\ket{0}$ & $-x^4$\\
\hline
$\ket{0}\ket{1}$ & $-x^4$\\
\hline
$\ket{1}\ket{1}$ & $4x+16x^2+16x^3+33x^4+$\\
& $z^2(3x+16x^2+19x^3+24x^4)+z^{-2}(3x+16x^2+19x^3+24x^4)+$\\
& $z^4(8x^2+24x^3+16x^4)+z^{-4}(8x^2+24x^3+16x^4)+$\\
& $z^6(15x^3+32x^4)+z^{-6}(15x^3+32x^4)+24z^8x^4+24z^{-8}x^4$\\
\hline
$\ket{3/2}\ket{3/2}$ & $2z(3x^{\frac32}+10x^{\frac52}+8x^{\frac72})+2z^{-1}(3x^{\frac32}+10x^{\frac52}+8x^{\frac72})+$\\
& $2z^3(2x^{\frac32}+10x^{\frac52}+10x^{\frac72})+2z^{-3}(2x^{\frac32}+10x^{\frac52}+10x^{\frac72})+$\\
& $2z^5(5x^{\frac52}+14x^{\frac72})+2z^{-5}(5x^{\frac52}+14x^{\frac72})+18z^7x^{\frac72}+18z^{-7}x^{\frac72}$\\
\hline
$\ket{2}\ket{2}$ & $9x^2+24x^3+16x^4+$\\
& $z^2(8x^2+24x^3+16x^4)+z^{-2}(8x^2+24x^3+16x^4)+$\\
& $z^4(5x^2+24x^3+21x^4)+z^{-4}(5x^2+24x^3+21x^4)+$\\
& $z^6(12x^3+32x^4)+z^{-6}(12x^3+32x^4)+21z^8x^4+21z^{-8}x^4$\\
\hline
$\ket{5/2}\ket{5/2}$ & $2z(6x^{\frac52}+14x^{\frac72})+2z^{-1}(6x^{\frac52}+14x^{\frac72})+$\\
& $2z^3(5x^{\frac52}+14x^{\frac72})+2z^{-3}(5x^{\frac52}+14x^{\frac72})+$\\
& $2z^5(3x^{\frac52}+14x^{\frac72})+2z^{-5}(3x^{\frac52}+14x^{\frac72})+14z^7x^{\frac72}+14z^{-7}x^{\frac72}$\\
\hline
$\ket{3}\ket{3}$ & $16x^3+32x^4+z^2(15x^3+32x^4)+z^{-2}(15x^3+32x^4)+$\\
& $z^4(12x^3+32x^4)+z^{-4}(12x^3+32x^4)+z^6(7x^3+32x^4)+z^{-6}(7x^3+32x^4)+$\\
& $16z^8x^4+16z^{-4}x^4$\\
\hline
$\ket{7/2}\ket{7/2}$ & $x^{\frac72}(20z+20z^{-1}+18z^3+18z^{-3}+14z^5+14z^{-5}+8z^7+8z^{-7})$\\
\hline
$\ket{4}\ket{4}$ & $x^4(25+24z^2+24z^{-2}+21z^{4}+21z^{-4}+16z^6+16z^{-6}+9z^8+9z^{-8})$\\
\hline
\caption{$(SU(2)_1\times SU(2)_{-1})/{\mathbb Z}_2$}
\end{longtable}
\begin{longtable}{|l|l||l|l|}
\hline
GNO charges & Index contribution & GNO charges & Index contribution\\
\hline
$T=0$ & & $T=5$ & \\
\hline
$\ket{0,0}\ket{0,0}$ & $1+4x+12x^2+8x^3+12x^4$ & $\ket{3,2}\ket{3,2}$ & $2(6x^{\frac52}+14x^{\frac72})$ \\
$\ket{1,-1}\ket{1,-1}$ & $4x+16x^2+16x^3+33x^4$ & $\ket{4,1}\ket{4,1}$ & $2(5x^{\frac52}+14x^{\frac72})$\\
$\ket{1,-1}\ket{0,0}$ & $-x^4$ & $\ket{5,0}\ket{5,0}$ & $2(3x^{\frac52}+14x^{\frac7/2})$\\
$\ket{0,0}\ket{1,-1}$ & $-x^4$ & $\ket{6,-1}\ket{6,-1}$ & $14x^{\frac72}$\\
$\ket{2,-2}\ket{2,-2}$ & $9x^2+24x^3+16x^4$ & &\\
$\ket{3,-3}\ket{3,-3}$ & $16x^3+32x^4$ & &\\
$\ket{4,-4}\ket{4,-4}$ & $25x^4$ & &\\
\hline
$T=1$ & & $T=6$ & \\
\hline
$\ket{1,0}\ket{1,0}$ & $2(x^{\frac12}+6x^{\frac32}+10x^{\frac52}+7x^{\frac72})$ & $\ket{3,3}\ket{3,3}$ & $10x^3+16x^4$\\
$\ket{2,-1}\ket{2,-1}$ & $2(3x^{\frac32}+10x^{\frac52}+8x^{\frac72})$ & $\ket{4,2}\ket{4,2}$ & $15x^3+32x^4$\\
$\ket{3,-2}\ket{3,-2}$ & $2(6x^{\frac52}+14x^{\frac72})$ & $\ket{5,1}\ket{5,1}$ & $12x^3+32x^4$\\
$\ket{4,-3}\ket{4,-3}$ & $20x^{\frac72}$ & $\ket{6,0}\ket{6,0}$ & $7x^3+32x^4$\\
& & $\ket{7,-1}\ket{7,-1}$ & $16x^4$\\
\hline
$T=2$ & & $T=7$ &\\
\hline
$\ket{1,1}\ket{1,1}$ & $3x+8x^2+12x^3+8x^4$ & $\ket{4,3}\ket{4,3}$ & $20x^{\frac72}$\\
$\ket{2,0}\ket{2,0}$ & $3x+16x^2+19x^3+24x^4$ & $\ket{5,2}\ket{5,2}$ & $18x^{\frac72}$\\
$\ket{3,-1}\ket{3,-1}$ & $8x^2+24x^3+16x^4$ & $\ket{6,1}\ket{6,1}$ & $14x^{\frac72}$\\
$\ket{4,-2}\ket{4,-2}$ & $15x^3+32x^4$ & $\ket{7,0}\ket{7,0}$ & $8x^{\frac72}$\\
$\ket{5,-3}\ket{5,-3}$ & $24x^4$ & &\\
\hline
$T=3$ & & $T=8$ & \\
\hline
$\ket{2,1}\ket{2,1}$ & $2(3x^{\frac32}+10x^{\frac52}+9x^{\frac72})$ & $\ket{4,4}\ket{4,4}$ & $15x^4$\\
$\ket{3,0}\ket{3,0}$ & $2(2x^{\frac32}+10x^{\frac52}+10x^{\frac72})$ & $\ket{5,3}\ket{5,3}$ & $24x^4$\\
$\ket{4,-1}\ket{4,-1}$ & $2(5x^{\frac52}+14x^{\frac72})$ & $\ket{6,2}\ket{6,2}$ & $21x^4$\\
$\ket{5,-2}\ket{5,-2}$ & $18x^{\frac72}$ & $\ket{7,1}\ket{7,1}$ & $16x^4$\\
& & $\ket{8,0}\ket{8,0}$ & $9x^4$\\
\hline
$T=4$ & \\
\hline
$\ket{2,2}\ket{2,2}$ & $6x^2+12x^3+12x^4$ & &\\
$\ket{3,1}\ket{3,1}$ & $8x^2+24x^3+16x^4$ & &\\
$\ket{4,0}\ket{4,0}$ & $5x^2+24x^3+21x^4$ & &\\
$\ket{5,-1}\ket{5,-1}$ & $12x^3+32x^4$ & &\\
$\ket{6,-2}\ket{6,-2}$ & $21x^4$ & &\\
\hline
\caption{$U(2)_1\times U(2)_{-1}$.
$T$ stands for the topological charge.}
\end{longtable}
\begin{longtable}{|l|l|}
\hline
GNO charges & Index contribution\\
\hline
$\ket{0}\ket{0}$ & $1+4x+12x^2+z^2(3x+8x^2)+z^{-2}(3x+8x^2)$\\
$\ket{1}\ket{1}$ & $4x+16x^2+z^2(3x+16x^2)+z^{-2}(3x+16x^2)$\\
$\ket{2}\ket{2}$ & $9x^2+8z^2x^2+8z^{-2}x^2$\\
\hline
\caption{$SU(2)_1\times SU(2)_{-1}$}
\end{longtable}
\begin{longtable}{|l|l|}
\hline
GNO charges & Index contribution\\
\hline
$\ket{0}\ket{0}$ & $1+4x+12x^2+z^2(3x+8x^2)+z^{-2}(3x+8x^2)$\\
$\ket{1/2}\ket{1/2}$ & $4x+17x^2+z^2(3x+16x^2)+z^{-2}(3x+16x^2)$\\
$\ket{1}\ket{1}$ & $9x^2+8z^2x^2+8z^{-2}x^2$\\
\hline
\caption{$(SU(2)_2\times SU(2)_{-2})/{\mathbb Z}_2$}
\end{longtable}
|
\section{What is a random sequence?}
About 30 years ago now, my friend and mentor J. Laurie Snell got
interested in the question of what constitutes an infinite random sequences
of $0$s and $1$s.
The model here is an infinite sequence of independent flips of a fair
coin,
with heads counting as $1$ and tails as $0$.
Or more technically, the standard product measure on
$\prod_{i=1}^\infty \{0,1\}$,
which up to a little fussing is the same as the standard Borel
measure on the unit interval $[0,1]$.
Some sequences are obviously not random:
\begin{itemize}
\item
$0000000000\ldots$;
\item
$1111111111\ldots$;
\item
$0101010101\ldots$;
\item
$1101001000100001\ldots$.
\end{itemize}
Other presumably non-random sequences are the binary expansion
of $1/\pi$, or $1/e$, or any sequence that can be printed out
by a computer program.
As there are only a countable number of computer programs,
this gives us only a countable set;
the complement is still an uncountable set of full measure.
Throwing out these computable sequences doesn't come near to doing the job.
Many more sequences need to be weeded out.
For example, a random sequence should exhibit the
strong law of large numbers: Asymptotically it should have half $0$s and half
$1$s.
There are uncountably many sequences that fail this test,
e.g. all those of the form $00a00b00c00d00e\ldots$
(On Beyond Zebra!).
And having half $0$s and half $1$s is far from enough.
Really we want the sequence to have all those properties dictated by
probability theory,
like say,
the law of the iterated logarithm.
This is a relative of the central limit theorem
which states---well, you can look it up.
What matters is that it is a statement about the sequence that
either is true or false,
and which according to probability theory is true \emph{almost surely}.
Now, some people have felt that a test for randomness should be effective in
some sense.
For example, maybe you should be able to make money from a non-random sequence,
as you could for example from a sequence of more than half $1$s by betting
on $1$ each time.
Looking into this led Laurie into a thicket of papers by Kolmogorov,
Martin-L\"of,
Schnorr, and Chaitin,
which are all very interesting and answer the question of what is random
from a certain perspective.
(This approach is by now a thriving industry:
See Downey et al.\ \cite{downeyetal:calibrating};
Nies
\cite{nies:randomness};
and
Downey and Hirschfeldt
\cite{downeyhirschfeldt:algorithmic}.)
At this point Jim Baumgartner, a logician who had been drawn into this
morass by Laurie, proposed the following:
\begin{definition} \label{def:random}
A sequence is \emph{non-random}
if and only if it belongs to some definable set of sequences
having measure $0$.
Here definable means uniquely definable by a formula in the language of first order set theory having only one free variable (no parameters).
And by measure $0$ we mean outer measure $0$,
in case the set is not Borel-measurable.
\end{definition}
Since there are only a countable number of formulas,
we're throwing out a countable collection of sets of measure $0$,
so the sequences that remain---the \emph{random} sequences,
should form a set of full measure.
Almost every sequence should be random.
I accepted this definition of random sequence for over 30 years.
Then about a month ago,
thanks to a stimulating colloquium talk by Johanna Franklin
and subsequent discussions with Rebecca Weber and Marcia Groszek,
I came to realize that
the argument that there are plenty of random sequences is bogus.
(Or perhaps it would be better to say that it is `suspect', in light
the fact that the same reasoning was used by Tarski---see section
\ref{sec:bogus} below.)
It is possible that
under this definition
\emph{there are no random sequences}!
The reason is that
the standard of axioms of set theory
(assuming they are consistent)
do not rule out the possibility that \emph{every set is definable}.
There are models of set theory, satisfying all the standard axioms,
where every set in the universe is definable.
(Cf.\ section \ref{sec:pdm} below.)
Of course these are countable models.
In these models, every sequence is definable,
and hence every singleton set consisting of a single sequence.
Since a singleton set has measure $0$,
every sequence belongs to a definable set of measure $0$, hence is
non-random.
So in such a model there are no random sequences.
Now we might have objected earlier,
`Of course there are no random sequences:
Given any sequence $\sigma$, the singleton set
$\{ \sigma \}$ has measure $0$.
A random sequence cannot be equal to any particular sequence.'
We thought we were avoiding this by only considering
definable sets of sequences:
There are only a countable number of definable sequences,
so there should be a full measure set of sequences left after we've ruled out
definable sequences.
And even after we've gone on to throw out
all definable sets of sequences of measure $0$,
there should remain a full-measure set of random sequences.
But now we're saying that in fact this objection might be
justified, after all:
For all we know, it might be the case that \emph{all sequences are definable}.
So maybe there really are no random sequences.
\section{How can this be?}
The answer is \emph{Skolem's paradox}.
We're dealing here with a countable model of set theory.
In any model of set theory, the collection of definable sets
will be countable from outside the model.
If the model is countable,
there is no obvious impediment to having every set be definable.
And in fact this turns out to be possible.
Look, the real problem here is that you can't define definability.
There is no formula in the language of set theory
characterizing a definable set,
because there is no formula characterizing a true formula.
(If there were, we'd be in real trouble.)
And so we can't write a formula characterizing
random sequences.
What we can do is write a formula characterizing random sequences
\emph{within a given model of set theory}.
Now, if the standard axioms of set theory are consistent,
then there is a model of set theory satisfying these axioms,
and in this model every set could be definable, which would
mean in particular that there are no random sequences in the model.
Note that we are not saying that it must be the case that there are
no random sequences.
There certainly are models where not all sequences are definable
(assuming set theory is consistent).
There presumably are models where the random sequences have full measure.
We're just saying that it \emph{may be}
that there are no random sequences.
\section{What to make of all this?}
One sensible response would be that we have missed the boat.
This definition of random sequence
is too restrictive.
We should be less demanding.
We should climb into the boat with Kolmogorov,
Martin-L\"of, Schnorr, and Chaitin.
And then we can talk not just about infinite sequences, but finite sequences
as well.
No finite sequences are completely random, of course, but clearly some are more
random than others.
We prefer to stick with Definition \ref{def:random},
and consider the possibility that there really are no random sequences.
Maybe the Old Man doesn't play dice with the universe of sets.
That is, assuming there really is a universe of sets.
In this connection, I can't resist quoting Abraham Robinson
\cite[p.\ 230]{robinson:64}:
\begin{quotation}
My position concerning the foundations of Mathematics is based
on the following two main points or principles.
(i) Infinite totalities do not exist in any sense of the word
(i.e., either really or ideally).
More precisely, any mention,
or purported mention,
of infinite totalities is, literally,
\emph{meaningless}.
(ii) Nevertheless, we should continue the business of Mathematics
``as usual,'' i.e., we should act \emph{as if}
infinite totalities really existed.
\end{quotation}
See also Cohen \cite{cohen:comments}.
\begin{comment}
who cautions:
\begin{quote}
Of course, good mathematics is beautiful,
while most philosophical discussion is barren and certainly not beautiful.
\end{quote}
\end{comment}
\section{Models of set theory for which all sets are definable} \label{sec:pdm}
I've found that Cohen's book `Set Theory and the Continuum Hypothesis'
\cite{cohen:continuum} is
a good place to look for general background on model theory.
The book is addressed to non-specialists, and `emphasizes
the intuitive motivations while at the same time giving as complete
proofs as possible'.
Here,
quoted verbatim from Cohen
\cite[pp.\ 104--105]{cohen:continuum},
are precise statements about models where all sets are definable.
\begin{theorem}
$\mathrm{ZF} + \mathrm{SM}$ implies the existence of a unique transitive model $M$
such that if $N$ is any standard model there is an $\in$-isomorphism
of $M$ into $N$. $M$ is countable.
\end{theorem}
Here $\mathrm{SM}$ is the statement that $\mathrm{ZF}$ has a standard model,
meaning one where the membership relation in the model coincides with
the `real world' membership relation $\in$.
The existence of a standard model is not provable because it implies
$\mathrm{Con}(\mathrm{ZF})$.
Cohen
\cite[p.\ 79]{cohen:continuum} says that $\mathrm{SM}$ is `most probably ``true"',
and gives an intuitive argument for accepting it as an axiom.
$\mathrm{SM}$ holds just if it is possible to quit early in the transfinite induction
that produces Godel's constructible universe $L$, and still have a model of
$\mathrm{ZF}$.
To get the minimal model $M$, we stop the construction at the earliest possible
ordinal.
$M$ satisfies $V=L$
(cf. \cite[p.\ 104]{cohen:continuum}),
hence also the axiom of choice, so in $M$ we have
a standard model of $\mathrm{ZFC} + (V=L)$.
\begin{theorem}
For every element $X$ in $M$ there is a formula $A(y)$ in $\mathrm{ZF}$ such
that $x$ is the unique element in $M$ satisfying $A_M(x)$.
Thus in $M$ every element can be `named'.
\end{theorem}
Models where every set is definable are called `pointwise definable'.
So according to these results, if $\mathrm{ZF}$ has a standard model, then
$\mathrm{ZFC}$ has a pointwise definable standard model.
The requirement that $\mathrm{ZF}$ has a standard model can be dispensed with,
if we don't care about winding up with a pointwise definable model
that is non-standard.
John Steel points out that
the definable sets within any model $N$ of $\mathrm{ZFC} + (V=L)$
constitute an elementary submodel $H$.
This implies that $H$ is a model of $\mathrm{ZFC} + (V=L)$,
and every set in $H$ is definable in $H$ (not just in $N$).
So, starting with any model for $\mathrm{ZF}$, we can restrict to a model
of $\mathrm{ZFC} + (V=L)$,
and within that find a pointwise definable model of $\mathrm{ZFC} +(V=L)$.
For much more about pointwise definable models of set theory, see
Hamkins, Linetsky, and Reitz \cite{hamkinsetal:pointwise}.
\section{What was wrong with the proof?} \label{sec:bogus}
So, what was wrong with the proof that there are plenty of random sequences?
Let's look at the proof, given by Tarski
\cite[p.\ 220]{tarski:definissable}
in 1931,
that there exist undefinable sets of real numbers.
This same proof method would show that there exist undefinable real
numbers,
or what is the same, undefinable random sequences.
Tarski's original paper
\cite{tarski:definissable} is in French.
Here,
from Tarski
\cite[p.\ 119]{tarski:definable},
is an English translation:
\begin{quote}
Moreover it is not difficult to show that the family of all definable
sets (as well as that of the functions which determine them) is only
denumerable, while the family of \emph{all} sets of numbers is not
denumerable. The existence of undefinable sets follows immediately.
\end{quote}
If anyone but Tarski had written this, I think we would say that
the author is confusing the system with the metasystem. In the
metasystem, we can talk about the family $D$ of definable sets of reals
of the system, and prove that it is denumerable \emph{in the metasystem}.
In the system itself, we can't talk about $D$, but we can talk about the
family $P$ of all sets of reals of the system, and prove that it is not
denumerable \emph{within the system itself}.
But from this we cannot
conclude that $D$ differs from $P$:
$D$ is denumerable in the metasystem;
$P$ is not denumerable in the system.
There is no contradiction here.
Tarski goes on to give a second proof:
\begin{quote}
`Plus encore' [the translation has `Also', but a better rendering
might be `And not only that'],
the definable sets can be arranged in an ordinary infinite sequence;
by applying the diagonal procedure it is possible \emph{to define
in the metasystem a concrete set which would not be definable in the
system itself}. In this there is clearly no trace of any antinomy,
and this fact will not appear at all paradoxical if we take proper
note of the relative character of the notion of definability.
\end{quote}
If anyone but Tarski had written this, I think we would say that the
author has failed to take proper note of the relative character of
\emph{being a set}.
This concrete set that we've defined in the metasystem
might not be a set in the system itself. So we have failed to produce
an undefinable set of the system.
Now, this critique of Tarski's reasoning has been written from the perspective
of model theory---a modern perspective based in large part
on work that Tarski did after writing down these proofs.
We interpret the `metasystem'
as a formal system in which statements about models (`systems')
of set theory can be formulated and proven.
If we choose as the metasystem ZFC, the usual formal system for set
theory, then in this metasystem we can prove a statement to the effect
that there exist models of ZFC where every set is definable.
This means that Tarski's arguments that there must be undefinable sets
cannot be correctly formalized in ZFC (assuming ZFC is consistent).
And we have pointed out where the problem lies in trying to formalize them.
So, does this mean that Tarski's arguments are \emph{wrong}?
They are if it is fair to recast them in the framework of model theory.
But it is not really clear that this is fair.
It has been suggested, for example, that Tarski is thinking of the
`standard model' rather than some arbitrary model.
That's all very well, but what does it mean concretely?
What special methods of proof apply to the `standard model'?
If we can make sense of Tarski's arguments,
then presumably we can salvage the argument
that the set of random sequences from Definition \ref{def:random}
has full measure.
Maybe it is false that maybe there are no random sequences.
\section*{Acknowledgement and disclaimer}
I'm grateful for help I've received from
Jim Baumgartner, Johanna Franklin,
Steven Givant, Marcia Groszek, Joel David Hamkins,
Laurie Snell, John Steel, and Rebecca Weber.
None of them is responsible for my opinions, or whatever errors
may be on display here.
Despite the impression I've been trying to give,
I'm aware that I know next to nothing about logic.
I've written this note because I find this whole business intriguing,
and I believe other people will too.
|
\section{Introduction}
In the last decades there has been a growing interest in
approaching Global Optimization problems by different numerical
techniques (see, for example, \cite{Bomze}--\cite{HansenE},
\cite{HorstPard}--\cite{Mockus}, \cite{Pinter,Ratschek},
\cite{SS2000}--\cite{Zhigljavsky} and references given therein).
Such an interest is motivated by a large number of real-life
applications where such problems arise (see, for example,
\cite{BaBo92}--\cite{Blondel},
\cite{Crivelli,FaPuSe99,FloPar,Grossman,HorstPard,Nemirovskii,Pard2000,PardShaXue
Pinter,Ritter,Sen,Se99,Shao}). These problems often lead to deal
with multiextremal non-differentiable objective function and
constraints. In such a context the Lipschitz condition becomes the
unique information about the problem.
It has been proved by Stephens and Baritompa~\cite{SteBar} that
if the only information about the objective function $\varphi(y)$
is that it belongs to the class of Lipschitz functions and the
Lipschitz constant is unknown, there does not exist any
deterministic or stochastic algorithm that, after a finite number
of evaluations of $\varphi(y)$, is able to provide an
underestimate of the global minimum $\varphi(y^*)$. Of course,
this result is very discouraging because usually in practice it is
difficult to know the constant. Nevertheless, the necessity to
solve practical problems remains. That is why in such problems
instead of the statement (S1) `{\it Find an algorithm able to stop
in a given time and provide an $\varepsilon$-approximation of the
global optimum $\varphi(y^*)$}' the statement (S2) `{\it Find an
algorithm able to stop in a given time and return the lowest
value of $\varphi(y)$ obtained by the algorithm}' is used. Under
this statement, either the computed solution (possibly improved by
a local analysis) is accepted by final users (engineers,
physicists, chemists, etc.) or the global search is repeated with
changed parameters of the algorithm.
Theoretical analysis of algorithms (depending on parameters) for
solving problems (S2) is similar to analysis of penalty methods.
It is proved for global optimization algorithms that for a {\it
fixed} problem there exists a parameter $P^*$ such that
parameters $P \ge P^*$ allow to solve the problem (obviously, the
parameter $P^*$ is problem-dependent). The problem `{\it How to
determine $P^*$?}' is not discussed in such an analysis, and in
every concrete case it is solved using additional information
about the problem. For example, in methods using a given
Lipschitz constant (see survey \cite{HanJau}) $P^*$ is the
Lipschitz constant and it is not discussed how to obtain it.
Other examples are diagonal methods (see \cite{Pinter}) and
information algorithms (see \cite{Se95,St78}) using similar
parameters. For all those methods it is possible to prove that if
a value $P \ge P^*$ is used as the parameter, then they converge
only to global minimizers. An alternative approach is represented
by methods converging to every point in the search domain (see
\cite{Jones,Torn}).
In this paper a constrained Lipschitz global minimization problem
is considered. In an informal way it can be stated as follows:
\begin{enumerate}
\item[(i)]
The objective function is multiextremal, non-differentiable,
'black box', and requires a high time to be evaluated;
\item[(ii)]
Constraints are non-convex (or even multiextremal) and
non-differentiable, leading to a complex feasible region
consisting of disjoint, non-convex sub-regions;
\item[(iii)]
Both the objective function and constraints are Lipschitz
functions with unknown Lipschitz constants;
\item[(iv)]
Both the objective function and constraints may be partially
defined, i.e., if a constraint is not satisfied at a point, the
rest of constraints and the objective function may be not defined
at that point.
\end{enumerate}
It can be seen from this statement that the problem belongs to the
class of problems considered in~\cite{SteBar}, thus statement
(S2) will be used hereinafter. One promising way to face problem
(i)--(iv) is the information approach introduced by Strongin in
\cite{St78,St85,St92}. It uses Peano type space-filling curves
(see~\cite{Bertsimas,Butz,PlatzBar,St78,SS2000} for examples of
usage of space-filling curves in mathematical programming) to
reduce the original Lipschitz multi-dimensional problem to a
H\"{o}lder univariate one (a comprehensive presentation of this
approach can be found in \cite{SS2000}). Global optimization of
H\"{o}lder functions (see \cite{Gourdin,Lera,Vanderbei}) has
given new tools for solving the reduced one-dimensional problem.
Peano curves avoid constructions of support (or auxiliary)
functions usually used in the multi-dimensional Lipschitz
optimization (see, for example,
\cite{HorstPard,Kearfott,Kelley,Pinter,Ratschek,SS2000} and
references given therein). Of course, if the user knows that the
objective function is differentiable and/or the problem is
convex, there is no sense to work with Peano curves and specific
methods explicitly using information about differentiability or
convexity can be applied. In contrast, when you deal with
non-differentiable 'black box' multiextremal problems, it is not
possible to work with sophisticated techniques using derivatives
(or other strong a priori suppositions) and such a reduction to
one dimension can help significantly.
In this paper, a novel algorithm belonging to the family of
information methods is proposed. It uses two powerful ideas for
solving problem (i)--(iv). The first one is the index scheme (see
\cite{SeMa95,St78,St85,StronMark,SS2000}), allowing to solve
Lipschitz problems where both the objective function $\varphi(y)$
and constraints $G_i(y),\, 1\le i\le m,$ may be multiextremal and
partially defined. Its importance increases in this case because
it is not clear how to solve such problems by using, for example,
the penalty approach. In fact, the latter requires that
$\varphi(y)$ and $G_i(y),\, 1\le i\le m,$ are defined over the
whole search domain. It seems that missing values can be simply
filled in with either a big number or the function value at the
nearest feasible point. Unfortunately, in the context of Lipschitz
algorithms, incorporating such ideas can lead to infinitely high
Lipschitz constants, causing degeneration of the methods and
non-applicability of the penalty approach. Thus, for problems
where the Lipschitz condition is almost a unique additional
information, the ability of the index scheme to work with
partially defined problems becomes crucial. Moreover, the index
scheme does not introduce any additional parameter and/or
variable, whereas the penalty approach requiring determination of
the penalty coefficient.
The second idea being at the basis of the new method is the local
tuning on the behaviour of the objective function and constraints
(see~\cite{Se95,Se98a,Se99,SeMa95}). Original information methods
work with {\it global} adaptive estimates of the Lipschitz
constants, i.e., the same estimates are used over the {\it whole}
search region for $\varphi(y)$ and functions $G_i(y),\, 1\le i\le
m$. However, global estimates (adaptive or given a priori) of the
Lipschitz constants may provide a poor information about the
behavior of the objective function over every small sub-region of
$D$. It has been shown in~\cite{Pinter,Se95,Se98a,Se99,SeMa95},
for different classes of global optimization problems, that local
estimates for different sub-regions of $D$ can accelerate the
search significantly. Of course, it is necessary to balance
global and local information about $\varphi(y)$ obtained by the
method during the search. Such a balancing is very important
because using only the local information can lead to missing the
global solution (see~\cite{Se98b,SteBar}).
The next section contains a formal statement of the problem
(i)--(iv) and presents the new algorithm. Convergence conditions
of the new method are established in Section~3. Numerical
experiments collected in Section~4 show a satisfactory
performance of the new algorithm in comparison with two global
optimization techniques. Finally, Section~5 gives concluding
remarks.
\section{Theoretical background and the index information\\
algorithm with local tuning}
We start by formulating the Lipschitz optimization problem
satisfying requirements (i)--(iii). Find the constrained global
minimum $\varphi^*$ and at least one minimizer $y^*$ such that
\begin{equation}
\varphi^* = \varphi(y^*) = \min\{\varphi(y): y\in S, \quad
G_i(y)\le 0,\quad 1\le i\le m\}, \label{f1}
\end{equation}
where the search domain is the hyperinterval
\[
S = \{y\in \mathbb{R}^N: a_j\le y_j\le b_j,\quad 1\le j\le N\},
\]
$\mathbb{R}^N$ is the $N$-dimensional Euclidean space, and
the objective function $\varphi(y)$ (henceforth denoted
as $G_{m+1}(y)$) and the functions $G_i(y),\, 1\le i\le m$, of
the constraints can be multiextremal and non-differentiable,
satisfying the Lipschitz condition with unknown constants
$0<\widehat{L}_i<\infty,\,1\le i\le m+1$, i.e.,
\begin{equation}
|G_i(y')-G_i(y'')|\le \widehat{L}_i \, ||y'-y''||,\quad 1\le i\le
m+1,\quad y',y''\in S. \label{f2}
\end{equation}
Without loss of generality, we shall consider the search domain
$S=D$, where
\begin{equation}
D=\{y\in \mathbb{R}^N: -2^{-1}\le y_j\le 2^{-1},\quad 1\le j\le
N\}. \label{f3}
\end{equation}
Formulation (\ref{f1})--(\ref{f3}) assumes that all the functions
$G_i(y),\, 1\le i\le m+1,$ can be evaluated in the whole region
$D$. In order to incorporate requirement (iv), we shall assume
that each function $G_i(y)$ is defined and computable only in the
corresponding subset $Q_i\subset D$, where
\begin{equation}
Q_1=D,\quad Q_{i+1}=\{y\in Q_i:G_i(y)\le 0\},\quad 1\le i\le m.
\label{f4}
\end{equation}
The above assumption also imposes the order in which the functions
$G_i(y), 1\le i\le m$, are evaluated. In many applications this
order is determined by the nature of the problem. In other cases
the user introduces a specific order suit for some reasons (for
example, first verify easier computable constraints).
In view of (\ref{f4}), the initial problem
(\ref{f1})--(\ref{f3}) is rewritten as
\begin{equation}
\varphi(y^*)=\min\{G_{m+1}(y): y\in Q_{m+1}\}, \label{f5}
\end{equation}
\begin{equation}
|G_i(y')-G_i(y'')|\le L_i \, ||y'-y''||, \quad y',y''\in Q_i,
\quad 1\le i\le m+1, \label{f6}
\end{equation}
where $L_i\le\widehat{L}_i$.
In order to start the description of the method it is necessary to
remind the idea of the space-filling curves. Such curves were
first introduced by Peano in~\cite{Peano} and Hilbert
in~\cite{Hilbert} and are fractal objects constructed on the
principle of self-similarity. They possess the property `to fill'
any cube $D$ in $\mathbb{R}^N$, i.e., they pass through every
point of $D$. An example of construction of such a
curve (see ~\cite{Sa94,SS2000} for details) is given in Fig.~\ref{Fig_1}.
Naturally, in numerical algorithms, approximations of the curve
are used (Fig.~\ref{Fig_1} presents approximations of levels one
to four).
\begin{figure}[t]
\framebox[\columnwidth]{\centerline{\psfig{figure=Vinc_1.ps,width=0.80\columnwidth,silent=yes}}}
\caption{Approximations of levels one to four to the Peano
(Hilbert) curve in two dimensions.}\label{Fig_1}
\end{figure}
It has been shown in~\cite{Butz,St78,St92} (see
also~\cite{SS2000}) that the multi-dimensional problem
\begin{equation}
\varphi^{*} = \varphi(
y^{*} ) = \min \{ \varphi(y):\; y \in D \} , \label {f2.1}
\end{equation}
where $\varphi(y)$ is a Lipschitz function with constant $L$,
$0 < L < \infty,$ can be reduced by a Peano-type space-filling
curve $y(x)$ to the one-dimensional problem
\begin{equation}
f^*=f(x^*) = \min \{ f(x),\; x \in [ 0,1 ]\} , \label {f2.2}
\end{equation}
where the notation $f(x) = \varphi(y(x))$ is used for the obtained
reduced one-dimensional function. Moreover, (see Theorem~8.1 in
\cite{SS2000}), the function $f(x)$ satisfies the H\"{o}lder
condition
\begin{equation}
| f(x') - f(x'') | \leq H \, | x' - x''|^{1/N}, \,
\quad x',x'' \in [ 0,1], \label {f2.3}
\end{equation}
in the H\"{o}lder metric
\begin{equation}
\rho(x',x'') = | x' - x''|^{1/N}, \label {f2.4}
\end{equation}
where $N$ is from (\ref{f3}) and $H=2\,L\,\sqrt{N+3}$.
By applying the same curve $y(x)$ to the functions $G_i(y),\; 1\le
i\le m+1$, and using designations $g_i(x)=G_i(y(x))$, the
multi-dimensional problem (\ref{f5})--(\ref{f6}) is reduced to
the following constrained one-dimensional problem
(\ref{f2.5})--(\ref{f2.7}):
\begin{equation}
g_{m+1}^*=g_{m+1}(x^*)=\min\{g_{m+1}(x): x\in q_{m+1}\},
\label{f2.5}
\end{equation}
where the region $q_{m+1}$ is defined by the relations
\begin{equation}
q_1=[0,1],\quad q_{i+1}=\{x\in q_i:g_i(x)\le 0\},\quad 1\le i\le
m. \label{f2.6}
\end{equation}
and the reduced functions $g_i(x), 1\le i\le m+1$, satisfy the
corresponding H\"{o}lder conditions
\begin{equation}
|g_i(x')-g_i(x'')|\le H_i \,|x'-x''|^{1/N}, \quad x',x''\in q_i,
\quad 1\le i\le m+1, \label{f2.7}
\end{equation}
with the constants $H_i=2\,L_i\,\sqrt{N+3}$, where $L_i, 1\le
i\le m+1$, are from (\ref{f6}). Note that for the regions $Q_i$
from (\ref{f4}) and $q_i$ from (\ref{f2.6}) the following relation
holds:
\[
Q_i=\{y(x):x \in q_i\}, \quad 1\le i\le m+1.
\]
This problem may be rewritten by using the index scheme
proposed in \cite{St85,StronMark} (see also \cite{St92}). The
scheme is an alternative to traditional penalty methods. Instead
of combining the objective and constraint functions into a penalty
one (and the need to define in a proper way penalty coefficients),
the scheme does not introduce any additional parameter or
variable and evaluates constraints one at a time at every point
where it has been decided to try to evaluate $g_{m+1}(x)$. The
function $g_i(x)$ is calculated only if all inequalities
\[
g_{j}(x)\le 0,\quad 1\le j<i,
\]
have been satisfied. In its turn, the objective function
$g_{m+1}(x)$ is evaluated only for those points where all the
constraints have been satisfied.
The index scheme juxtaposes to every point of the interval $[0,1]$
an index
\[
\nu =\nu (x),\quad 1\le \nu \le m+1,
\]
which is defined by the conditions
\begin{equation}
g_{j}(x)\le 0,\quad 1\le j\le \nu -1,\quad g_{\nu }(x)>0,
\label{f2.8}
\end{equation}
where for $\nu =m+1$ the last inequality is omitted. Thus,
$\nu(x)$ represents the number of the first constraint not
satisfied at~$x$ when $\nu(x)<m+1$. Then $\nu(x) =m+1$ means
that all constraints were satisfied at~$x$ and the objective
function $g_{m+1}(x)$ may be evaluated at this point.
Let us now define an auxiliary function $\Phi (x)$ as follows
\begin{equation}
\Phi (x) = g_{\nu }(x) - \left\{ \begin{array}{ll}
0, & \nu (x)<m+1,\\[5pt]
g^{*}_{m+1}, & \nu (x)=m+1,
\end{array}
\right. \label{f2.9}
\end{equation}
where $g^{*}_{m+1}$ is the solution of the problem
(\ref{f2.5})--(\ref{f2.7}). Due to (\ref{f2.8}) and (\ref{f2.9}),
the function $\Phi (x)$ has the following properties:
\begin{itemize}
\item[i.]
$\Phi (x)>0$, when $\nu (x)<m+1$;
\item[ii.]
$\Phi (x)\ge 0$, when $\nu (x)=m+1$;
\item[iii.]
$\Phi (x)=0$, when $\nu (x)=m+1$ and $g_{m+1}(x)=g^{*}_{m+1}$;
\item[iv.]
$\Phi (x)$ is not continuous at the a priori unknown boundary
points of the sets $q_i,2\le i\le m+1$.
\end{itemize}
Thus, the global minimizer of the constrained problem
(\ref{f2.5})--(\ref{f2.7}) coincides with the so\-lu\-tion
$x^{*}$ of the following unconstrained problem
\begin{equation}
\Phi (x^{*})=\min \{ \Phi (x): x\in [0,1] \}. \label{f2.10}
\end{equation}
Of course, instead of the fractal Peano curve $y(x)$ its $d$-level
approximation $y_{d}(x)$ is used for evaluation of $\Phi (x)$
(approximations of levels one to four are shown in
Fig.~\ref{Fig_1}, from where it can be seen how a $d$-level
approximation is obtained).
The new index information global optimization algorithm with local
tuning presented below generalizes and evolves two methods. On the
one hand, the information global optimization algorithm with local
tuning proposed in \cite{Se95} for solving the problem
(\ref{f2.1}) and, consequently, the problem
(\ref{f2.2})--(\ref{f2.3}). On the other hand, the index
algorithm with local tuning proposed in~\cite{SeMa95} for solving
the one-dimensional problem
\[
\min \{f(x): x \in [a, b],\; g_i(x) \le 0, \; 1 \le i \le m \}
\]
where both $f(x)$ and $g_i(x), 1 \le i \le m,$ are Lipschitz
continuous functions.
During every iteration of the algorithm a point $x \in [0,1]$ is
chosen and the value $\Phi (x)$ is evaluated (hereinafter such an
evaluation will be called a {\it trial} and the corresponding
point $x$ a {\it trial point}). Suppose now that $k$ iterations
of the algorithm have already been executed (two initial trials
are done at the end points $x^{0} = 0$ and $x^{1} = 1$). The
point $x^{k+1}, k \ge 1$, is determined by the following
algorithm.
\begin{description}
\item[\bf Step 1.]
The points $x^{0}, \ldots, x^{k}$ of the previous iterations are
renumbered by subscripts as follows
\[ 0 = x_{0} < x_{1} < \cdots < x_i < \cdots < x_{k} = 1. \]
\item[{\bf Step 2.}]
To each point $x_i$ associate the index $\nu_i=\nu(x_i)$, and the
value
\[
z_i=g_{\nu _i}(x_i) -\left\{ \begin{array}{ll}
0, & \nu_i<m+1,\\[5pt]
z_k^{*}, & \nu_i=m+1,
\end{array}
\right.
\]
where
\begin{equation}
z_k^{*}=\min \{ g_{m+1}(x_i): 0 \le i \le k, \nu(x_i)=m+1 \}.
\label{f2.11}
\end{equation}
The value $z_k^{*}$ estimates the unknown value $g^{*}_{m+1}$
from~(\ref{f2.9}) on the basis of the available data.
\item[\bf Step 3.]
Calculate lower bounds $\mu_{j}$ for the global H\"older constants
$H_{j}$ of the functions $g_j(x), 1 \le j \le
m+1, $ as follows
\begin{equation}
\mu_{j}= \max \Big\{ \frac{| z_{p}-z_{q} |}{ (x_{p}-x_{q})^{1/N}}:
0 \le q < p \le k, \nu _{p} = \nu_{q} = j \Big\}. \label{f2.12}
\end{equation}
Whenever $\mu_j$ can not be calculated, set $\mu_{j} = 0$.
\item[\bf Step 4.]
For each interval $[x_{i-1},x_i], i = 1, \ldots, k$, calculate the
values
\begin{equation}
M_i= \max \{ \lambda _i, \gamma _i, \xi \} \label{f2.13}
\end{equation}
that estimate the local H\"older constant over the interval
$[x_{i-1},x_i]$. The values $\lambda_i$ and $\gamma_i$ reflect
the influence on $M_i$ of the local and global information
obtained during the previous $k$ iterations, $\xi $ is a small
number - parameter of the method. The values $\lambda_i$ and
$\gamma_i$ are defined by
\begin{equation}
\lambda _i= \max \{ l_i, c_i, r_i\}, \label{f2.14}
\end{equation}
where
\[
\ba{ll} c_i &= \left\{\!\!\! \begin{array}{cl}
\displaystyle{\frac{| z_i-z_{i-1} |}{ (x_i-x_{i-1})^{1/N}}}, & \mbox{if} \; \nu_i = \nu_{i-1} \\[15pt]
0, & $ otherwise$
\end{array}
\right.\\[25pt]
l_i &= \left\{\!\!\! \begin{array}{cl}
\displaystyle{\frac{| z_{i-1}-z_{i-2}|}{ (x_{i-1}-x_{i-2})^{1/N}}}, & \mbox{if} \; i\ge 2,\quad \nu_{i-2} = \nu_{i-1}, \nu_{i-1}\ge\nu_i \\[15pt]
0, & \mbox{otherwise}
\end{array}
\right.\\[25pt]
r_i &= \left\{\!\!\! \begin{array}{cl}
\displaystyle{\frac{| z_{i+1}-z_i |}{ (x_{i+1}-x_i)^{1/N} }}, & \mbox{if} \; i\le k-1,\quad \nu_{i+1} = \nu_i, \nu_i \ge \nu_{i-1} \\[15pt]
0, & $ otherwise$
\end{array}
\right.
\ea
\]
and
\begin{equation}
\gamma_i= \frac{\mu_{j}(x_i-x_{i-1})^{1/N} }{X^{\rm max}_{j}},
\hspace{6mm} j=\max\{\nu _i,\nu_{i-1}\}, \label{f2.15}
\end{equation}
where $X^{\rm max}_{j}= \!\max \{ (x_i-x_{i-1})^{1/N}
\; \max\{\nu_i,\nu_{i-1} \} = j, \quad 1 \le i \le k \}$.
\item[\bf Step 5.]
For each interval $[x_{i-1},x_i], i = 1, \ldots, k$, calculate the
{\it characteristic} of the interval
\begin{equation}
R_i = \left\{ \begin{array}{ll}
\Delta_i + \,\displaystyle{ \frac{(z_i-z_{i-1})^{2} }{ (r M_i)^2 \Delta_i }
- \, \frac{2\, (z_i+z_{i-1})}{r M_i}}, & \nu_i=\nu_{i-1} \\[10pt]
2\, \Delta_i - \, \displaystyle{\frac{4\, z_i}{r M_i}}, & \nu_i>\nu_{i-1}\\[10pt]
2\, \Delta_i - \, \displaystyle{\frac{4\, z_{i-1}}{r M_i}}, & \nu_{i-1}>\nu_i
\end{array}
\right.
\label{f2.16}
\end{equation}
where $\Delta_i=(x_i-x_{i-1})^{1/N}$ and $r>1$ is a real value --
the reliability parameter of the method.
\item[\bf Step 6.] Choose the interval $[x_{t-1},x_t]$ having the
maximal characteristic as follows
\begin{equation}
t = \min \{ \arg\max \{ R_i: 1 \le i \le k\}
\}. \label{f2.17}
\end{equation}
\item[\bf Step 7.] If the interval $[x_{t-1},x_t]$ is such that
\[
(x_t - x_{t-1})^{1/N} > \delta,
\]
where $\delta$ is a given tolerance, go to {\bf Step 8}. Otherwise
{\bf Stop}.
\item[\bf Step 8.]
Execute the $(k+1)$-th iteration at the point
\begin{equation}
x^{k+1} \!= \frac{x_t + x_{t-1}}{2} - \frac{{\rm sign} (z_t -
z_{t-1})|z_t - z_{t-1}|^N}{2\, r\,M_t^N},
\label{f2.18}
\end{equation}
if $\nu_t = \nu_{t-1}$. In all other cases do it at the point
\begin{equation}
x^{k+1} \!= (x_t + x_{t-1})/2. \label{f2.19}
\end{equation}
Set $k=k+1$ and go to {\bf Step 1}.
\end{description}
If trial points with the index $m+1$ have been generated by the
algorithm then the value~$z_k^{*}$ from (\ref{f2.11}) can be taken
as an estimate of the global minimum $\varphi^*$ from (\ref{f1})
and the corresponding point $y_{d}(x_k^{*})$ such that
$z_k^{*}=g_{m+1}(x_k^{*})$ as an estimate of the point $y^*$. If
no points with the index $m+1$ have been generated by the
algorithm then it is necessary to continue the search with
changed parameters of the method.
Let us give a few remarks on the algorithm introduced above. The
information algorithms are derived as optimal statistical decision
functions within the framework of a stochastic model representing
the function to be optimized as a sample of a random function. The
characteristic $R_i$ in terms of the information approach (see
\cite{St78,SS2000}) may be interpreted (after normalization) as
the probability of finding a global minimizer within the interval
$[x_{i-1},x_{i}]$ based on the data available during the current
iteration. The method uses in its work four parameters: $d, r,
\xi ,$ and $\delta$. Their choice will be discussed in Section~4
while presenting the numerical experiments.
For every sub-interval $[x_{i-1},x_{i}], 1\le i\le k$, {\it
global} estimates $\mu _{j}$ of the {\it global} H\"older
constants $H_{j}, 1\le j\le m+1$, from (\ref{f2.7}) are not used.
In contrast, {\it local} estimates $M_{i}, 1\le i\le k$, from
(\ref{f2.13}) are adaptively determined. The values $\lambda _{i}$
and $\gamma _{i}$ reflect the influence on $M_{i}$ of the local
and global information obtained during the previous $k$
iterations. When the interval $[x_{i-1},x_{i}]$ is small, then
$\gamma _{i}$ is small too (see (\ref{f2.15})) and, due to
(\ref{f2.13}), the local information represented by $\lambda
_{i}$ has major importance. The value $\lambda _{i}$ is
calculated by considering the intervals $[x_{i-2},x_{i-1}], \;
[x_{i-1},x_{i}],$ and $ [x_{i},x_{i+1}]$ (see (\ref{f2.14})) as
those which have the strongest influence on the local estimate.
When the interval $[x_{i-1},x_{i}]$ is very wide, the local
information is not reliable and the global information
represented by $\gamma _{i}$ is used. Thus, local and global
information are balanced in the values $M_{i}, 1\le i\le k$. Note
that the method uses the local information over the {\it whole}
search region $[0,1]$ (and, consequently, over the whole
multi-dimensional domain $D$) {\it during} the global search both
for the objective function and constraints (being present in a
implicit form in the auxiliary function $\Phi (x)$).
\section{Convergence conditions}
In the further theoretical consideration it is assumed that each
region $Q_i, 2\le i\le m+1,$ from (\ref{f4}) is a union of $T_i$
disjoint sub-regions $Q_i^j, \; 1\le j\le T_i,$ having positive
volume (in general, the numbers $T_i$ are unknown to the user)
where each sub-region $Q_i^j, 1\le j\le T_i, 2\le i\le m+1,$ is
robust and can be non-convex. It is assumed that the feasible
region $Q_{m+1}$ is nonempty. Recall that a set $B$ is robust if
for each point $\beta$ belonging to the boundary of $B$ and for
any $ \varepsilon > 0$ there exists an $\varepsilon$-neighborhood
$\varepsilon(\beta)$ such that $\varepsilon(\beta) \bigcap B$ has
a positive volume. Supposition about robustness of the sets
$Q_i^j$ is quite natural in engineering applications (for example,
in optimal control) where the optimal solution must have an
admissible neighborhood of positive volume. This requirement is a
consequence of the inevitable inaccuracy of the physical systems,
and ensures that small changes of the parameters of the system
will not lead the solution to leave the feasible region.
Theoretical results presented in this section have the spirit of
existence theorems. They show (similarly to the penalty approach)
that for a given problem there exist parameters of the method
allowing to solve this {\it fixed} problem. Recall that (see
\cite{SteBar}) the information available from the statement
(\ref{f5})--(\ref{f6}) is not sufficient for establishing concrete
values of the parameters. These should be chosen from an
additional information about the nature of the practical problem
under consideration (see the next section).
The first result links solutions of the original
multi-dimensional problem (\ref{f5})--(\ref{f6}) to the solutions
of the corresponding one-dimensional problem
(\ref{f2.5})--(\ref{f2.7}) reduced via the curve $y_{d}(x)$. It
shows that for any given $\varepsilon>0$ there exists a level
$d$ such that the approximation $y_{d}(x)$ passes through the
$\varepsilon$-neighborhood of the global solution $y^*$ of the
problem (\ref{f5})--(\ref{f6}). Moreover, the global solution
$y_d^*$ of the reduced problem (\ref{f2.5})--(\ref{f2.7}) will
also belong to the same $\varepsilon$-neighborhood.
\begin{theorem}
\label{t1} For every problem (\ref{f5})--(\ref{f6}) having the
unique solution $y^*$, and a given accuracy $\varepsilon>0$,
there exists a subdivision level $d$ such that in the
$\varepsilon$-neighborhood $\varepsilon(y^*)$ of the global
minimizer $y^*$ there exists a segment $h$ of the level-$d$
approximating curve $y_{d}(x)$ with the following properties:
\begin{enumerate}
\item
the segment $h$ belongs to $Q_{m+1} \cap \varepsilon(y^*)$ and has
a finite positive length;
\item
the segment $h$ contains the global minimizer $y_d^*$ of the
problem
\begin{equation}
\varphi(y_d^*)=\min\{G_{m+1}(y_d(x)):\; y_d(x)\in Q_{m+1}, x
\in[0,1]\}. \label{f3.1}
\end{equation}
\end{enumerate}
\end{theorem}
{\bf Proof.} We have assumed that the feasible region $Q_{m+1}$ is
the union of $T_{m+1}$ robust sub\-regions $Q_{m+1}^j, \; 1\le
j\le T_{m+1}$. Suppose that $y^* \in Q_{m+1}^{p}$, where
$Q_{m+1}^{p}$ is one of those subregions. Then $Q_{m+1}^{p} \cap
\varepsilon(y^*)$ is a robust set with a nonzero volume. This
means that by increasing the accuracy of approximation it is
possible to find an approximation level $d$ such that the curve
$y_{d}(x)$ will have a segment $h \subset Q_{m+1}^{p} \cap
\varepsilon(y^*)$ with the required properties.
An illustration to Theorem~\ref{t1} is given in Fig.~\ref{Fig_2},
where a region $Q_{m+1}^{p}$ is shown in grey color.
Analogously, if problem (\ref{f5})--(\ref{f6}) has more than one
solution, an approximation level $d$ providing existence of such
intervals having finite positive lengths can be found for each of
global minimizers. Of course, (\ref{f3.1}) may be satisfied only
for one interval but, anyway, all these intervals will contain
$\varepsilon$-approximations of the global solution. Thus,
Theorem~\ref{t1} allows us to concentrate the further theoretical
investigation on the behaviour of the method during the solution
of problem (\ref{f2.10}).
\begin{figure}[t]
\centerline{\psfig{file=Vinc_4.ps,width=80mm,height=80mm,silent=yes}}
\caption{Illustration of existence of a subdivision
level $d$ allowing an $\varepsilon$-approximation $y_{d}^*$ of
the global solution $y^*$ such that the point $y_{d}^*$ is the
global minimizer of the objective function $G_{m+1}(y_d(x))$
on $Q_{m+1} \cap y_{d}(x)$.}
\label{Fig_2}
\end{figure}
\begin{lemma}
\label{l1} Let $\bar{x}$ be a limit point of the infinite
sequence $\{x^{k}\}$ generated by the algorithm and let $i = i(k)$
be the number of an interval $[x_{i-1},x_{i}]$ containing this
point during the $k$-th iteration. Then
\begin{equation}
\lim_{k \rightarrow \infty}(x_{i(k)} - x_{i(k)-1})^{1/N} = 0,
\label{f3.2}
\end{equation}
and for every $\sigma > 0$ there exists an iteration number
$n(\sigma )$, such that the inequality
\begin{equation}
R_{i(k)} < \sigma \label{f3.3}
\end{equation}
holds for all $k \ge n(\sigma )$.
\end{lemma}
{\bf Proof.} The new trial point $x^{k+1}$ from (\ref{f2.18}) or
(\ref{f2.19}) falls into an interval $[x_{i-1},x_{i}]$ (where
$i(k)=t(k)$ is determined by (\ref{f2.17})), and divides this one
into two sub-intervals
\[
[x_{i-1},x^{k+1}],\quad [x^{k+1},x_{i}].
\]
Let us show that there exists a number $\alpha$, independent of
the iteration number $k$, such that
\begin{equation}
\max \{ x_{i}-x^{k+1}, x^{k+1}-x_{i-1} \} \le \alpha
(x_{i}-x_{i-1}), \quad 0.5\le \alpha <1,
\label{f3.4}
\end{equation}
holds for these intervals. In the case $\nu _{t}\neq \nu _{t-1}$
the estimate (\ref{f3.4}) is obtained immediately from
(\ref{f2.19}) by taking $\alpha_1 =0.5$. In the opposite case,
from (\ref{f2.13})--(\ref{f2.15}) it follows that
\begin{equation}
\frac{\mid z_{t}-z_{t-1}\mid ^N}{M_{t}^N} \le x_{t}-x_{t-1},
\quad \nu _{t}=\nu _{t-1}. \label{f3.5}
\end{equation}
From this inequality, (\ref{f2.18}), and the fact that $r > 1$ we
can conclude that in the case $\nu _{t}=\nu _{t-1}$ (\ref{f3.4})
is true for $ \alpha_2 = (r+1)/(2r)$. Since $ \alpha_2>\alpha_1$,
(\ref{f3.4}) is proved by taking $ \alpha = \alpha_2$.
Now, considering (\ref{f3.4}) together with the existence of a
sequence converging to $\bar{x}$ (this point is a limit point of
$\{x^{k}\}$), we can deduce that (\ref{f3.2}) holds. Note that in
the case when two intervals containing the point $\bar{x}$ there
exist (i.e., when $\bar{x}\in \{x^{k}\}$) the number $i=i(k)$ is
juxtaposed to the interval for which (\ref{f3.2}) takes place.
Let us prove (\ref{f3.3}). From (\ref{f2.16}) and (\ref{f3.5}) we
obtain, for intervals with indexes $\nu _{i}=\nu _{i-1}$,
\[
R_{i} \le (1+r^{-2})(x_{i} - x_{i-1})^{1/N} \le 2(x_{i} -
x_{i-1})^{1/N},
\]
taking into account that $\Phi (x)\ge 0$ and $r>1$. The last
estimate holds also for intervals with $\nu _{i} \ne \nu _{i-1}$.
This inequality and (\ref{f3.2}) lead to (\ref{f3.3}).
\begin{theorem}
\label{t2} Let $x^{*}$ be any solution to the problem
(\ref{f2.10}) and $j=j(k)$ be the number of an interval
$[x_{j-1},x_{j}]$ containing this point during the $k$-th
iteration. Then, if for $k\ge k^{*}$ the condition
\begin{equation}
rM_{j}>\left\{ \begin{array}{ll}
2^{1-1/N}C_{j}+\sqrt{2^{2-2/N}C_{j}^{2}-D_{j}^{2}}, & \nu_{j-1}=\nu_{j},\\
2C_{j}, & \nu_{j-1}\neq \nu_{j},
\end{array}
\right. \label{f3.6}
\end{equation}
holds, the point $x^{*}$ will be a limit point of
$\{x^{k}\}$. The values $C_{j}$ and $D_{j}$ used in (\ref{f3.6})
are
\begin{equation}
C_{j}=\left\{ \begin{array}{ll}
z_{j-1}/(x^{*}-x_{j-1})^{1/N}, & \nu_{j-1}>\nu_{j},\\
\max \{ z_{j-1}/(x^{*}-x_{j-1})^{1/N}, z_{j}/(x_{j}-x^{*})^{1/N} \},& \nu_{j-1}=\nu_{j},\\
z_{j}/(x_{j}-x^{*})^{1/N}, & \nu_{j-1}<\nu_{j},
\end{array}
\right. \label{f3.7}
\end{equation}
\begin{equation}
D_{j}=\left\{ \begin{array}{ll}
\mid z_{j}-z_{j-1}\mid /(x_{j}-x_{j-1})^{1/N}, & \nu_{j-1}=\nu_{j},\\
0, & \mbox{\rm otherwise}.
\end{array}
\right. \label{f3.8}
\end{equation}
\end{theorem}
{\bf Proof.} Consider the case $\nu _{j-1}=\nu _{j}$. Due to
(\ref{f3.7}), we can write
\begin{equation}
z_{j-1}\le C_{j}(x^{*}-x_{j-1})^{1/N}, \label{f3.9}
\end{equation}
\begin{equation}
z_{j}\le C_{j}(x_{j}-x^{*})^{1/N}. \label{f3.10}
\end{equation}
Now, by using (\ref{f3.9}), (\ref{f3.10}), and the designation
\[
\beta = (x^{*} - x_{j-1} ) / ( x_{j} - x_{j-1} ),
\]
we deduce
\[
\begin{array}{ll}
z_{j-1} + z_{j} & \leq C_{j} ( ( x^{*} - x_{j-1} )^{1/N} + ( x_{j}
- x^{*} )^{1/N} ) \\
& =C_{j} ( \beta^{1/N} + ( 1 - \beta )^{1/N} ) ( x_{j} - x_{j-1}
)^{1/N} \\
& \leq C_{j} ( x_{j} - x_{j-1} )^{1/N} \max \{ \beta^{1/N} + (1 -
\beta )^{1/N} : 0 \leq \beta \leq 1 \} \\
& = 2^{1-1/N} C_{j} ( x_{j} - x_{j-1} )^{1/N}.
\end{array}
\]
By using this estimate and (\ref{f3.6}), (\ref{f3.8}), we obtain
\[
\ba{ll} R_{j(k)} & =( x_{j} - x_{j-1} )^{1/N} +
\displaystyle{\frac{( z_{j-1}\!-\! z_{j} )^{2}}{ ( r M_{j} )^{2}
(x_{j}\!-\!x_{j-1} )^{1/N}} - \frac{2( z_{j-1}\!+\!z_j)}{r M_{j}}}\\
\\
& \ge (x_j\!-\!x_{j-1})^{1/N} (1\!+\!D^{2}_{j} (r M_{j}
)^{-2}\!-\!2^{2-1/N} C_j(r M_{j} )^{-1}).
\ea
\]
Now, due to (\ref{f3.6}), we can conclude that
\begin{equation}
R_{j(k)} > 0. \label{f3.11}
\end{equation}
In the case $\nu _{j-1}> \nu _{j}$ the estimate (\ref{f3.9})
takes place due to (\ref{f3.7}). From (\ref{f2.16}) it follows
that
\[
R_{j(k)} \ge 2(x_{j}-x_{j-1})^{1/N}(1-2C_{j}(rM_{j})^{-1}),
\]
and, consequently, by taking into consideration (\ref{f3.6}), the
inequality (\ref{f3.11}) holds in this case also. Truth of
(\ref{f3.11}) for $\nu _{j-1}< \nu _{j}$ is demonstrated by
analogy.
Assume now that $x^{*}$ is not a limit point of the sequence
$\{x^{k}\}$. Then, there exists a number $K_s$ such that for all
$k \ge K_s$ the interval $[x_{j-1},x_{j}]$, $j=j(k)$, is not
changed, i.e., new points will not fall into this interval.
Consider again the interval $[x_{i-1},x_{i}]$ from Lemma~\ref{l1}
containing a limit point $\bar{x}$. It follows from (\ref{f3.3})
that there exists a number $K_p$ such that
\[
R_{i(k)} < R_{j(k)}
\]
for all $k \ge k^{*}=\max \{K_s,K_p\}$. This means that, starting
from the iteration $k^{*}$, the characteristic of the interval
$[x_{i-1},x_{i}], i = i(k), k\ge k^{*}$, is not maximal. Thus, a
trial will fall into the interval $[x_{j-1},x_{j}]$. But this fact
contradicts our assumption that $x^{*}$ is not a limit point.
\begin{corollary}
\label{c1}
Given the conditions of Theorem~\ref{t2}, Theorem~\ref{t1} ensures
that trial points generated by the algorithm will fall in the
feasible sub-region $Q_{m+1} \cap \varepsilon(y^*)$.
\end{corollary}
Theorems~\ref{t1},~\ref{t2} generalize for the constrained case
(\ref{f2.5})--(\ref{f2.7}) results established in~\cite{Se95}
for the Lipschitz global optimization problems with box
constraints. Usually, Lipschitz global optimization algorithms
need an overestimate (adaptive or a priori given) of the {\it
global} Lipschitz constant for the {\it whole} search region
(see, for example, survey \cite{HanJau} and references given
therein). The new algorithm does not need knowledge of the
precise Lipschitz constants for the functions $G_i(y),\, 1\le
i\le m+1,$ over the whole search region. In fact, the point $y^*$
(as all points in $D$) has up to $2^N$ images on the Peano curve.
To obtain an $\varepsilon$-approximation of $y^*$ it is
sufficient the fulfillment of the condition (\ref{f3.6}) (which is
significantly weaker than the Lipschitz condition) for only one
image of $y^*$ on a segment $h$ of the curve such that $h \subset
Q_{m+1} \cap \varepsilon(y^*)$. In the rest of the region $D$ the
constants $L_i,\, 1\le i\le m+1,$ can be underestimated.
In the preceding it has been established that, if condition
(\ref{f3.6}) is satisfied, the trial sequence $\{x^{k}\}$
generated by the algorithm converges to the global minimizer $x^*$
of the function $\Phi (y)$ and, consequently, the corresponding
sequence $\{y_d(x^{k})\}$ converges to the global minimizer $y^*$
of the function $\varphi (y)$. In the following theorem it is
proved that for any problem (\ref{f5})--(\ref{f6}) there exists a
continuum of values of $r$ satisfying (\ref{f3.6}).
\begin{theorem}
\label{t3} For any problem (\ref{f5})--(\ref{f6}) a value $r^{*}$
exists such that, for all $r > r^{*}$, condition (\ref{f3.6}) is
satisfied for all iteration numbers $k \ge 1$.
\end{theorem}
{\bf Proof.} Let us choose
\begin{equation}
r^{*} = 2^{3-1/N}\xi^{-1}\sqrt{N+3} \; \max \{ L_i:\; 1\le i\le
m+1\}, \label{f3.12}
\end{equation}
and take a value $r>r^{*}$. Since, due to (\ref{f2.7}),
\[
H_i = 2\, \sqrt{N+3}\; L_i,\quad 1\le i\le m+1,
\]
the value $r^{*}$ from (\ref{f3.12}) can be rewritten as
\[\begin{array}{ll}
r^{*} &= 2^{2-1/N}\xi^{-1} H,\\[5pt]
H &= \max \{ H_i: 1\le i\le m+1\}. \end{array}
\]
For all $k \ge 1$, for the estimates $M_{j(k)}$ and the values
$C_{j(k)}$ from (\ref{f3.7}) the inequalities
\[
\xi \le M_{j(k)}, \hspace{5mm}C_{j(k)} \le H,
\]
hold. Then, when the interval $[x_{j-1},x_{j}]$ has $\nu_j =
\nu_{j-1}$, it follows from (\ref{f3.12}) that
\[
\begin{array}{lll}
rM_{j(k)} &> r^{*}M_{j(k)} &\ge r^{*}\xi = 2^{2-1/N}H \\[5pt]
&> 2^{2-1/N}C_{j} &\ge
2^{1-1/N}C_{j} + \sqrt{2^{2-2/N}C_{j}^{2}-D_{j}^{2}}.
\end{array}
\]
To complete the proof it is sufficient to note that in the case
$\nu_j \ne \nu_{j-1}$ the last estimate can be substituted by
\[
2^{2-1/N}C_{j} > 2\,C_{j},
\]
because $N \ge 2$. Thus, for the chosen $r>r^{*}$, condition
(\ref{f3.6}) is satisfied.
Theorem~\ref{t2} establishes sufficient convergence conditions for
the method. Theorem~\ref{t3} ensures that there exists a
continuum of values of the parameter $r$ satisfying these
conditions. However, these theorems just prove {\it existence} of
such values and can not be considered as an instrument for
determining the value $r^{*}$. This value is problem-dependent
and can not be found without additional information about the
objective function and constraints. In presence of such
information it is possible to solve the problem (S1), otherwise
only the problem (S2) can be faced.
Condition (\ref{f3.6}) gives us a suggestion how to choose a
reasonable value of $r$ to start optimization in the case (S2).
Note that the value $D_{j}$ from (\ref{f3.9}) can be equal to
zero. The role of $M_j$ is to estimate $C_j$. Thus, if $D_{j}=0$
and $M_j=C_j$, it follows from (\ref{f3.6}) that $r$ should be
greater than $2^{2-1/N}$. Some practical advises for the choice of
the method parameters will be given in the next section.
\section{Numerical experiments}
This section presents numerical experiments that investigate
the performance of the proposed algorithm in solving some test
problems. The first three problems are taken from the
literature; the other problems are proposed by the authors. All
the methods have been implemented in MATLAB~\cite{Matlab} and the
experiments have been executed at a PC with Pentium III $733$MHz
processor.
{\bf Experiment 1.} In the first experiment the new algorithm is
compared with two methods: (i) the information algorithm for
solving problems with box constraints proposed in~\cite{St85},
combined with the penalty approach; (ii) the original index
information algorithm proposed by Strongin and Markin
in~\cite{St85}--\cite{StronMark}, that does not use the local
tuning. These methods have been chosen for comparison because all
of them have similar computational cost for a single iteration (of
course, if the cost required for the search of the penalty
coefficient allowing to solve the problem is not taken in
consideration for the method (i)), they use Peano curves, and have
the same stopping rule. Since the penalty approach needs the
objective function and constraints defined over the whole search
region, test problems enjoying this property have been chosen.
{\it Problem 1.} The first problem (see~\cite{StronMark}) is to
minimize the function
\begin{equation} \begin{array}{ll}
\varphi(y) = &
-1.5 \, y_1^2\exp(1-y_1^2-20.25\,(y_1-y_2)^2)\\[3pt]
& - \, (0.5\,(y_1-1)(y_2-1))^4\exp(2-(0.5\,(y_1-1))^4-(y_2-1)^4)
\end{array} \label{P1}
\end{equation}
over the rectangle $D = [0,4]\times[-1,3]$, under the constraints
\begin{displaymath} \begin{array}{l}
G_1(y) = 0.01 \,((y_1-2.2)^2 + (y_2-1.2)^2 - 2.25) \le 0,\\[3pt]
G_2(y) = 100 \,(1-(y_1-2)^2/1.44 - (0.5\,y_2)^2) \le 0,\\[3pt]
G_3(y) = 10 \,(y_2 - 1.5 - 1.5\sin(6.283\,(y_1-1.75))) \le 0.
\end{array} \end{displaymath}
The feasible region (see Fig.~\ref{Fig_3}) is the intersection of
the areas inside the circle $G_1(y) = 0$, outside the ellipses
$G_2(y) = 0$, and below the sinusoide $G_3(y) = 0$. It consists of
three disjoint and non-convex pieces with non-smooth boundaries
shown in Fig.~\ref{Fig_3} by the gray color. Here the solution is
$y^*=(0.942,0.944)$, giving the value
$\varphi(y^*)=-1.489$~\cite{SS2000}.
\begin{figure}[t]
\centerline{\psfig{file=penalty.eps,width=\textwidth,silent=yes}}
\caption{Trial points generated by the information algorithm
combined with the penalty approach with $\delta = 10^{-3}$ for
Problem~1}
\label{Fig_5}
\end{figure}
\begin{figure}[t]
\centerline{\psfig{file=sm.eps,width=\textwidth,silent=yes}}
\caption{ Trial points generated by the Strongin-Markin
algorithm with $\delta = 10^{-3}$ for Problem~1}
\label{Fig_4}
\end{figure}
\begin{figure}[ht]
\centerline{\psfig{file=new.eps,width=\textwidth,silent=yes}}
\caption{Trial points generated by the new algorithm with $\delta
= 10^{-3}$ for Problem~1}
\label{Fig_3}
\end{figure}
{\it Problem 2.} In this problem (see~\cite{St92}) the same
function as in the first experiment, over the same domain $D$, is
minimized subject to the constraints
\begin{equation} \begin{array}{l}
G_1(y) = -(y_1-2.2)^2-(y_2-1.2)^2 \le -1.21,\\[3pt]
G_2(y) = (y_1-2.2)^2+(y_2-1.2)^2 \le 1.25,
\end{array} \label{P2}
\end{equation}
that define a narrow annulus centered at the point $C \equiv
(2.2,1.2)$. Here the solution is $y^*=(1.088,1.088)$, giving the
value $\varphi(y^*)=-1.477$~\cite{SS2000}.
{\it Problem 3.} This is a minimization problem with four
constraints (see Section 9.2 in~\cite{Himmelblau}); the
expression of the objective function is reported in the appendix;
the search region is $D = [0,80] \times [0,80]$ and the
constraints are
\begin{equation} \begin{array}{l}
G_1(y) = 450 - y_1 y_2 \le 0,\\[3pt]
G_2(y) = (0.1\,y_1-1)^2-y_2\le 0,\\[3pt]
G_3(y) = 8\,(y_1-40)-(y_2-30)(y_2-55)\le 0,\\[3pt]
G_4(y) = (y_1-35)(y_1-30)/125+y_2-80\le 0.
\end{array} \label{P3}
\end{equation}
The feasible region for this problem is connected but not convex and has a
non-smooth boundary. The solution is the point
$y^*=(77.19,64.06)$, and the value of the function at this point
is $\varphi(y^*)=-59.59$~\cite{SS2000}.
{\it Problem 4.} The fourth test problem, proposed by the
authors, is defined by the search domain $D = [0,2\,\pi] \times
[0,2\,\pi]$, the non-smooth objective function
\begin{equation}
\varphi(y) = - |\sin(y_1) \sin(2\,y_2)| + 0.01\,(y_1\,y_2 +
(y_1-\pi)^2 + 3\,(y_2-\pi)^2)\,, \label{P4}
\end{equation}
and the non-smooth constraints
\begin{displaymath} \begin{array}{l}
G_1(y) = 1 - y_2 + \pi/2 - |\sin(2\,y_1)| + y_1/3 \le 0,\\[3pt]
G_2(y) = y_2 - 3\,\pi/2 + 4\,|(\sin(y_1+\pi)|+y_1/3-1.9 \le 0.
\end{array} \end{displaymath}
The feasible region is made of two disjoint pieces,
each of which is non-convex and has a non-smooth boundary. The
best solution, found by a fine grid over the search region, is
$y^*=(1.247,2.392)$, yielding $\varphi(y^*)= - 0.864$.
\newcommand{\xx}{\hspace*{-1mm}}
\begin{center}
\begin{table}[t]\centering
\caption{Results for Experiment 1 with $\delta = 10^{-3}$} \label{table_1}\footnotesize
\begin{tabular}{|c|rrrrrr|rrrrrr|}\hline
& \multicolumn{6}{c|}{Strongin-Markin method} & \multicolumn{6}{c|}{New method}\\
\xx Problem \xx\xx & $n_1$ & $n_2$ & $n_3$ & $n_4$ & $n_\varphi$ & $\varphi_c$ & $n_1$ & $n_2$ & $n_3$ & $n_4$ & $n_\varphi$ & $\varphi_c$ \\[1pt] \hline
1 & 1712 &\ \@ 953&\ \@ 673 & - & 279 & -1.489 & 289 & 201 & 119 & - &\@\@60 & -1.455 \\
2 & 1631 & 672 & - & - & 246 & -1.477 & 316 & 142 & - & - & 31 & -1.438 \\
3 & 351 & 348 & 316 & 140 & 88 & -59.59 & 106 & 103 & 90 & 65 & 36 & -59.34 \\
4 & 771 & 359 & - & - & 163 & -0.864 & 115 & 60 & - & - & 29 & -0.863 \\
\hline
\end{tabular}
\end{table}
\end{center}
\begin{center}
\begin{table}[t]\centering
\caption{Results for Experiments 1 with $\delta = 10^{-4}$}\label{table_2} \footnotesize
\begin{tabular}{|c|rrrrrr|rrrrrr|}\hline
& \multicolumn{6}{c|}{Strongin-Markin algorithm} & \multicolumn{6}{c|}{New method}\\
\xx Problem \xx\xx & $n_1$ & $n_2$ & $n_3$ & $n_4$ & $n_\varphi$ & $\varphi_c$ & $n_1$ & $n_2$ & $n_3$ & $n_4$ & $n_\varphi$ & $\varphi_c$ \\[1pt] \hline
1 & 4494 & 2636 & 1487 & - & 826 & -1.489 & 362 & 274 & 162 & - & 102 & -1.489 \\
2 & 4926 & 1941 & - & - & 781 & -1.478 & 376 & 170 & - & - & 44 & -1.439 \\
3 & 1073 & 1070 & 1032 & 655 & 359 & -59.59 & 115 & 112 & 99 & 74 & 45 & -59.34 \\
4 & 1867 & 787 & - & - & 355 & -0.864 & 135 & 72 & - & - & 40 & -0.863 \\ \hline
\end{tabular}
\end{table}
\end{center}
In all the experiments the maximum number of iterations was set to
5000, the level of approximation of the Peano curve
to $d = 10$, the value $\xi$ from Step 4 to $\xi = 10^{-8}$, and
the value of the reliability parameter to $r = 2.2$. Only the
points $0$ and $1$ have been used as starting points.
In the experiments with the penalty approach the constrained
problems were reduced to the unconstrained ones as follows
\[ \varphi_{P}(y)=\varphi(y) + P \cdot
\max\left\{G_1(y),G_2(y),\dots,G_{m}(y),0\right\}, y \in D.
\]
The coefficient $P$ has been computed by the rules:
\begin{description}
\item
-- the coefficient $P$ has been chosen equal to $0.1$ for all
the problems and it has been checked whether the found solution
for each problem belongs or not to the feasible subregions;
\item
-- if it does not belong to the feasible subregions, the
coefficient $P^*$ has been iteratively increased by $0.1$
until a feasible solution has been found.
\end{description}
Figures \ref{Fig_5}--\ref{Fig_3} present trial points generated by
the three methods during Experiment~1 with $\delta = 10^{-3}$.
Figure~\ref{Fig_5} shows the trial points generated by the penalty
approach. Remind that both the objective function and all the
constraints have been evaluated at every point.
Figures~\ref{Fig_4} and~\ref{Fig_3} show the points where the
algorithms using the index approach have chosen to evaluate the
constraints and eventually the target function. Points where the
target function has been computed are denoted by stars; those
where only the first constraint has been evaluated (and found
violated) are denoted by crosses; circles denote points where the
first constraint was satisfied and the second was not; squares
denote points were the first two constraints were satisfied and
the third was not.
\begin{center}
\begin{table}[t]\centering
\caption{Results obtained by the information algorithm combined
with the penalty approach Experiment 1 }\label{table_x}
\footnotesize
\begin{tabular}{|c|l|rr|rr|}\hline
Test Problem & $P^*$ & \multicolumn{2}{c|}{$\delta = 10^{-3}$} &
\multicolumn{2}{c|}{$\delta = 10^{-4}$}\\
& & $n$ & $\varphi_c$ & $n$ & $\varphi_c$ \\
\hline
1 & $0.1$ & $2298$ & $-1.489$ & ${\bf 5000}$ & $-1.489$ \\
2 & $0.3$ & $428$ & $-1.478$ & $1621$ & $-1.478$ \\
3 & $0.8$ & ${\bf 5000}$ & $-59.59$ & ${\bf 5000}$ & $-59.59$ \\
4 & $0.5$ & $283$ & $-0.864$ & $789$ & $-0.864$ \\ \hline
\end{tabular}
\end{table}
\end{center}
\begin{center}
\begin{table}[t]\centering
\caption{Speed-up obtained in Experiment 1 for Problems 1--4}\label{table_4} \footnotesize \centering
\begin{tabular}{|c|rrrr|rrrr|}\hline
Problem & 1 & 2 & 3 & 4 & 1 & 2 & 3 & 4 \\
& \multicolumn{4}{c|}{$\delta = 10^{-3}$} & \multicolumn{4}{c|}{$\delta = 10^{-4}$} \\
\hline
&&&&&&&& \\
$S_1$ & 4.29 & 5.13 & 3.74 & 4.03 & 12.25 & 10.84 & 14.68 & 8.83 \\
$S_2$ & 4.65 & 7.94 & 2.44 & 5.62 & 8.10 & 17.75 & 7.98 & 8.88 \\
$S_3$ & 5.41 & 5.21 & 3.11 & 6.34 & 10.49 & 12.96 & 9.41 &
11.27 \\
$\hat{S}_1$ & 7.95 & 1.35 & 47.17 & 2.46 & 13.81 & 4.31 & 43.48 & 5.84 \\
$\hat{S}_2$ & 38.30 & 13.81 & 138.89 & 9.76 & 49.02 & 36.84 & 111.11 & 19.73 \\
$\hat{S}_3$ & 13.74 & 2.63 & 62.50 & 4.16 & 22.47 & 8.24 & 56.18
& 9.58 \\
\hline
\end{tabular}
\end{table}
\end{center}
Table \ref{table_1} compares performance of the proposed
algorithm to the Strongin-Markin me\-thod for $\delta = 10^{-3}$.
The data reported are the computed solution $\varphi_c$, the
number of evaluations of each constraint $n_i$, and of the target
function $n_\varphi$. To investigate the sensitivity of both
algorithms to the value of $\delta$ the tests have been repeated
with $\delta = 10^{-4}$; results are reported in
Table~\ref{table_2}. Finally, Table~\ref{table_x} presents results
obtained by the information algorithm combined with the penalty
approach on Problems $1-4$, where $n$ is the number of iterations
executed by the algorithm. Cases where the algorithm was not able
to stop in 5000 iterations are shown in bold.
Table \ref{table_4} illustrates acceleration reached by the new
algorithm in comparison with the two methods tested in
Experiment~1. Three speed-up indexes have been used: the first
index is the ratio $S_1$ of iterations $n_1$ of the
Strongin-Markin method over the new one; the second is the ratio
$S_2$ of $n_\varphi$ of the objective function evaluations;
finally, $S_3$ is the ratio of the sums $n_\varphi + \sum_{k=1}^m
n_k$, being the number of summary evaluations of the objective
function and constraints. The same quantities for the penalty
approach are indicated by $\hat{S}_i, 1 \le i \le 3$. It can be
seen from Table~\ref{table_4} that the new method significantly
outperforms both methods tested. Moreover, the computational
burden increases very slowly for decreasing $\delta$ for it,
while the Strongin-Markin algorithm and the penalty approach show
a fast increase of iterations when a better accuracy is required.
Since Table \ref{table_4} shows an evident superiority of the new
method over the penalty algorithm, the latter is not used in the
further experiments (also because it is difficult to determine the
correct penalty coefficients for problems used in these
experiments).
{\bf Experiment 2.} This experiment aims at investigating the
sensitivity of both algorithms to the value of the reliability
parameter $r$. We have generated twenty pairs of random numbers
in the range $(-1,1)$, and have added any of them to the center
$C \equiv (2.2,1.2)$ of the annulus of Problem~2, thus obtaining
twenty new problems, whose solutions have been found with a fine
grid over the domain $D$.
They have been solved with both methods for either $r = 2.5$ and
$r = 3$. With $r = 2.5$ the proposed algorithm has found the
solution in 18 cases out of 20, the Strongin-Markin algorithm was
successful in 19 cases out of 20. Note that the proposed
algorithm has found the solution in the case the Strongin-Markin
algorithm has failed. With $r = 3$, both algorithms have found
the solution in all cases. The results (number of successes,
average number of computation of the constraints and of the
objective function) are reported in Table~\ref{table_3}. With $r
= 3$, in one case the Strongin-Markin algorithm has reached the
maximum number of iterations, but anyway it has found a good
approximation to the solution. It can be seen from
Table~\ref{table_3} that the computational burden increases very
slowly for increasing $r$ for the proposed algorithm, while the
Strongin-Markin algorithm shows a fast increase of iterations.
\begin{center}
\begin{table}[t]\centering
\caption{Results for Experiment 2}\label{table_3} \footnotesize
\begin{tabular}{|c|cc|cc|}\hline
& \multicolumn{2}{c|}{$r = 2.5$} & \multicolumn{2}{c|}{$r = 3$}\\
& Strongin-Markin method& New method & Strongin-Markin method & New method \\[1pt] \hline
Successes & $19/20$ & $18/20$ & $20/20$ & $20/20$ \\[3pt]
$n_1$ average & $2177$ & $359$ & $2939$ & $651$ \\
$n_2$ average & $874$ & $179$ & $1150$ & $330$ \\
$n_\varphi$ average & $282$ & $46$ & $312$ & $88$
\\
$S_1$ & -- & 6.06 & -- & 4.51 \\
$S_2$ & -- & 6.13 & -- & 3.55 \\
$S_3$ & -- & 5.71 & -- & 4.12 \\ \hline
\end{tabular}
\end{table}
\end{center}
{\bf Experiment 3.} The last experiment involves a problem having
changeable dimension $N$ and three constraints. The objective
function is
\begin{equation} \varphi_{N}(y) = y_1 +
e^{\rho_{N}-|\rho_{N}^2-5\,\rho_{N}+4|}, \hspace{5mm} \rho_{N}=
(\sum_{i=1}^{N}y_i^2)^\frac{1}{2}.
\label{P5}
\end{equation}
The search region is $D = [-2,8] \times [-6,4]\times \ldots \times
[-6,4]$ and the constraints are
\begin{displaymath} \begin{array}{l}
G_1(y) = (y_1-2.5)^2-6.25 + \sum_{i=2}^{N}y_i^2 \le 0,\\[3pt]
G_2(y) = -(y_1-2)^2 + 2.25 - \sum_{i=2}^{N}y_i^2 \le 0,\\[3pt]
G_3(y) = y_2 - 3\,\pi/2 + 4\,|(\sin(y_1+\pi)|+y_1/3+0.8 \le 0.\\[3pt]
\end{array}
\end{displaymath}
Numerical experiments have been executed with $2 \le N \le 6$
with different values of $r$ and $\delta$. The level of
approximation of the Peano curve was set to $d = 10$ for $2 \le N
\le 5$ and to $d = 8$ for $ N =6$, the value $\xi = 10^{-8}$. In
all the experiments the maximum number of iterations was set to
$40000$. The global solution is $y^*=(0,\ldots,0)$, yielding
$\varphi_{N}(y^*)= \exp(-4) \simeq 0.0183$.
The obtained results are shown in Tables~\ref{table_5}
and~\ref{table_6}. In all the experiments the new algorithm is
significantly faster. It can be seen from the Table~\ref{table_6}
that starting from dimension $N=3$ the Strongin-Markin algorithm
was not able to stop in less than $40000$ iterations.
Let us comment the choice of parameters of the new method. As it
has been mentioned in the Introduction, the statement of the
problem does not allow to create any deterministic or stochastic
algorithm that, after a finite number of evaluations of
$\varphi(y)$, is able to provide an underestimate of the global
minimum $\varphi(y^*)$. These means that it is impossible to say
without some additional information about the problem which
values of parameters ensure that the method finds the global
minimizer. However, the executed experiments can advice some
recommendations.
First of all, the parameters $d$ and $\xi$ are quite technical
and can be chosen easily. To have a better accuracy of the
solution it is necessary to have a better approximation of the
Peano curve determined by $d$. Since two points in the
multidimensional space have two images at the interval $[0,1]$,
then in a computer realization (see, for example, \cite{SS2000})
for a correct work of the method these images should be
represented by two different numbers. The mapping is such that
for a dimension $N$ and a level $d$ two numbers $a,b \in [0,1]$
will be considered different if $|a-b|>2^{-dN}$. For example, in
double precision the minimal representable number is $2^{-52}$.
Thus, if one works with a simple realization of the Peano curve
using double precision, the product $dN$ should be less than
$52$. Of course, more sophisticated implementations of the Peano
curve realizing numbers with more digits allow to increase this
number.
The second parameter is $\xi$ and it is chosen equal to a small
number (see experiments). It has been introduced in the method to
ensure that it works correctly in the case $l_i=c_i=r_i=0$ where
$l_i, c_i, r_i$ are from (\ref{f2.14}). The case when the real
H\"{o}lder constants are less than $\xi$ is degenerous for the
method because the local tuning is not used and the method works
slower.
The parameters $r$ and $\delta$ are chosen by the user on the
basis of additional information about the problem or just by
increasing $r$ and decreasing $\delta$ (see Table~\ref{table_5})
trying to obtain a satisfactory result (see statement (S2) in
Introduction). It can be also seen from Table~\ref{table_5} that
by increasing dimension of the problem it is necessary to change
$r$ and $\delta$ in the same way. This happens because, in spite
of increasing dimension, the reduced problem is always determined
over the interval $[0,1]$ independently on the dimension $N$. As a
consequence, the reduced problem has more local minima, and to
find the global one it is necessary to increase reliability of
the method (by increasing $r$) and to make the search more
accurate (by decreasing $\delta$).
\begin{table}
\centering \caption{Results for the new method on
Problems~(\ref{P5}) with dimension two to six} \label{table_5}
\footnotesize
\hspace*{-2cm}\begin{tabular}{|lr|llllll|rrrr|c|}\hline
$r$ & $\delta$ & $y_1$ & $y_2$ & $y_3$ & $y_4$ & $y_5$ & $y_6$ & $n_1$ & $n_2$ &
$n_3$ & $n_{\varphi}$ & $\varphi_c$ \\
\hline
2.35 & $10^{-3}$ & 0.0034 & 0.0596 & - & - & - & - &
130 & 51 & 41 & 36 &
0.0295 \\
2.35& $10^{-4}$ & 0.0007 & 0.0596 & - & - & - &
- & 147 & 57 & 47 & 42 & 0.0268\\
2.45 & $10^{-3}$ & 0.0068 & 0.1185& 0.1182 & - & - & - &
819 &
470 & 377& 249 & 0.0555\\
2.45 & $10^{-4}$ & 0.0052& -0.0088 & 0.0107 & - & - & - &
5847 & 4723 & 3975 &
3313 & 0.0252\\
2.7 & $10^{-3}$& 0.2540& -0.1553& -0.0674 & 0.4600 & - & -
& 708 & 533 & 364 & 249 &
0.6247\\
2.7 & $10^{-4}$ & 0.0264 & -0.0771& 0.0400 & -0.0679 & - &
- & 14441& 12485&10081 & 7410 & 0.0621\\
3.3 & $10^{-3}$ & 0.5146& -0.1650& -0.0771& -0.0924& -0.3799 & -
& 620 & 206 &164 &
111& 1.1704\\
3.3 & $10^{-4}$ & 0.0654 & 0.0303& -0.2920& 0.0596&
-0.0692 & - & 18535 & 12289 & 10761 & 7344 & 0.1748\\
3.35 &$10^{-3}$ & 0.5977 & -0.1992 & -0.0039& -0.3298&1.0898&
-0.6289 &
445 & 39 & 35 & 25 & 1.9579\\
3.35 &$10^{-4}$ & 0.3193 & 0.1523& -0.3945& 0.1523& 0.3477
&0.1523& 6610 & 2486 & 2292 &
1534 & 0.9688\\
3.35 & $5 \cdot 10^{-5} $ & 0.0508& -0.0430& 0.0742& -0.0430&
-0.1992& 0.0504& 19105 & 9871 & 8989 & 6233 & 0.1208 \\
\hline
\end{tabular}
\end{table}
\begin{table}
\centering \caption{Results for the Strongin-Markin algorithm on
Problems~(\ref{P5}) with dimension two to six}\label{table_6}
\footnotesize
\hspace*{-2cm}\begin{tabular}{|lr|llllll|rrrr|c|}\hline $r$ &
$\delta$ & $y_1$ & $y_2$ & $y_3$ & $y_4$ & $y_5$
& $y_6$ & $n_1$ & $n_2$ &
$n_3$ & $n_{\varphi}$ & $\varphi_c$ \\
\hline
2.35 & $10^{-3}$ & 0.0025 & 0.0010 & - & - & - & - & 5071 & 3811 & 3314 & 3066 & 0.0211\\
2.35 & $10^{-4}$ & 0.0000 & 0.0010 & - & - & - & - & 39883 & 33277 & 31108 & 30423 & 0.0185\\
2.45 & $10^{-3}$ & 0.0388 & 0.0010 & 0.0498 & - & - & - & 11003 & 5734 & 4911 & 3736 & 0.0654\\
2.45 & $10^{-4}$ & 0.0068 & 0.0400 & -0.0351 & - & - & - & {\bf 40000} & 22591 & 19901 & 16023 & 0.0320\\
2.7 & $10^{-3}$ & 0.1729 & 0.0693 & -0.1162 & -0.1860 & - & - & 5067 & 1833 & 1680 & 1037 & 0.2676\\
2.7 & $10^{-4}$ & 0.0459 & 0.1210 & -0.2139 & 0.0693 & - & - & {\bf 40000} & 17699 & 15402 & 10566 & 0.1271\\
3.3 & $10^{-3}$ & 0.5635 & 0.3037 & -0.0381 & 0.3916 & 0.0034 & - & 2661 & 432 & 381 & 220 & 1.5087\\
3.3 & $10^{-4}$ & 0.1436 & 0.3330 & -0.2627 & -0.0869 & 0.0492 & - & {\bf 40000} & 10487 & 9661 & 6137 & 0.3763\\
3.35 & $10^{-3}$ & 0.4756 & -0.2920 & -0.1846 & -0.0186 & -0.4482 & -1.1709 & 650 & 23 & 22 & 18 & 1.9356\\
3.35 & $10^{-4}$ & 0.3877 & 0.1377 & 0.4111 & 0.2061 & -0.3115 & -0.1260 & {\bf 40000} & 5413 & 5091 & 3251 & 1.1450\\
3.35 & $5\cdot 10^{-5}$ & 0.3877 & 0.1377 & 0.4111 & 0.2061 & -0.3115 & -0.1260 & {\bf 40000} & 5413 & 5091 & 3251 & 1.1450\\
\hline
\end{tabular}
\end{table}
\section{Conclusions}
In this paper, a novel global optimization algorithm for solving
multi-dimensional Lipschitz global optimization problems with
multiextremal partially defined constraints and objective function
has been presented. The new algorithm uses Peano type
space-filling curves and the index scheme to reduce the original
constrained problem to a H\"{o}lder one-dimensional problem. Local
tuning on the behaviour of the objective function and constraints
is executed during the work of the global optimization procedure
in order to accelerate the search. The new method works without
introducing any penalty coefficients and/or auxiliary variables.
Convergence conditions of a new type have been established for
the algorithm.
The new algorithm enjoys the following properties:
\begin{description}
\item
\hspace{8mm}{--} in order to guarantee convergence to a global
minimizer $y^{*}$ it is not necessary to know the precise
Lipschitz constants for the objective function and constraints
for the whole search region; on the contrary, only fulfillment of
condition (\ref{f3.6}) for a segment of the Peano curve
containing one of the images of $y^{*}$ is required;
\item
\hspace{8mm}{--} the usual problem of determining the moment to
stop the global procedure in order to start a local search does
not arise because local information is taken into consideration
throughout the duration of the global search;
\item
\hspace{8mm}{--} local information is taken into account not only
in the neighborhood of a global minimizer but also in the whole
search region, thus permitting a significant acceleration of the
search;
\item
\hspace{8mm}{--} thanks to usage of the index scheme, the new
algorithm does not introduce any additional parameters and/or
variables. Constraints are evaluated at every point one at a time
until the first violation of one of them, after that the rest of
constraints and the objective function are not evaluated at this
point. In its turn, the objective function is evaluated only for
that points where all the constraints have been satisfied.
\end{description}
The algorithm has been tested on a number of problems taken from
literature. Numerical results show the good performance of the new
technique in comparison with the two methods taken from
literature.
\section*{Appendix} \
The target function for Problem 3 is
\begin{displaymath}
\begin{array}{ll}
\varphi(y) = & - ( B_1 + B_2\, y_1 + B_3\, y_1^2 + B_4\, y_1^3 +
B_5\,
y_1^4 + B_6\, y_2 + B_7\, y_1 y_2 + B_8\, y_1^2 y_2\\[3pt]
& +\, B_9\, y_1^3 y_2 + B_{10}\, y_1^4 y_2 + B_{11}\, y_2^2 +
B_{12}\, y_2^3
+ B_{13}\, y_2^4 + B_{14}/(1+y_2) \\[3pt]
&+\, B_{15}\, y_1^2 y_2^2 + B_{16}\, y_1^3 y_2^2 + B_{17}\, y_1^3
y_2^3 +
B_{18}\, y_1 y_2^2 + B_{19}\, y_1 y_2^3 \\[3pt]
& +\, B_{20} \exp(0.0005\, y_1 y_2))
\end{array}
\end{displaymath}
where coefficients $B_1, \ldots, B_{20}$ are
\small \noindent $B_1= 75.1963666677;
B_2= -3.8112755343;
B_3= 0.1269366345;
B_4= -0.0020567665;\\
B_5= 0.0000103450;
B_6= -6.8306567613;
B_7= 0.0302344793;
B_8= -0.0012813448;\\
B_9= 0.0000352559;
B_{10}= -0.0000002266;
B_{11}= 0.2564581253;
B_{12}= -0.0034604030;\\
B_{13}= 0.0000135139;
B_{14}= -28.1064434908;
B_{15}= -0.0000052375;
B_{16}= -0.0000000063;\\
B_{17}0 = 0.0000000007;
B_{18}= 0.0003405462;
B_{19}= -0.0000016638;
B_{20}= -2.8673112392.$
\normalsize
|
\section{introduction}
Let $K = k(x,y)$ be the skew field of rational functions in the non-commutative
variables $x$ and $y$, where the ground field $k$ is $\mathbb{Q}$, $\mathbb{Q}(q)$ or any field containing either one. For any positive integer $r$, let $F_r$ be the Kontsevich automorphism of $K$, which is defined by
\begin{equation}\label{Kont_map}F_r : \left\{\begin{array}{l}x \mapsto xyx^{-1} \\ y \mapsto (1+y^r)x^{-1}.\end{array}\right. \end{equation}
Let
$$x_n=F_r^n(x)=\begin{array}{c}\underbrace{F_r\circ F_r \circ \cdots \circ F_r}(x)\\n\,\,\,\,\,\,\,\,\,\end{array}\text{ and } y_n =F_r^n( y),$$
for every integer $n$.
\begin{conj}[Kontsevich]\label{Kontconj}
Both $x_n$ and $y_n$ are non-commutative Laurent polynomials of $x$ and $y$ with non-negative integer coefficients.
\end{conj}
Usnich \cite{U} showed, by using derived categories, that $x_n$ and $y_n$ are non-commutative Laurent polynomials of $x$ and $y$. Independently, Berenstein and Retakh found an elementary proof for the Laurent phenomenon in \cite{BR} (with $F_r^n$ replaced by $\cdots F_{r_1}F_{r_2}F_{r_1}$). The Kontsevich conjecture was verified by Di Francesco and Kedem for $(r_1, r_2) \in \{(2, 2), (4, 1), (1, 4)\}$ in \cite{DK}, with explicit combinatorial expressions. The case of $r_1r_2\leq 3$ is elementary.
Conjecture~\ref{Kontconj} is related to the positivity conjecture for commutative cluster algebras. Cluster algebras were invented by Fomin and Zelevinsky \cite{FZ} in the Spring of 2000 as a tool for studying dual canonical bases and total positivity in semisimple
Lie groups. Since then, they have found applications and connections in a wide range of subjects including quiver representation theory, Poisson geometry, Teichm\"uller theory, tropical geometry and so forth. A cluster algebra is a commutative $\mathbb{Q}$-algebra with a family of distinguished generators (the \emph{cluster variables}) grouped into overlapping subsets (the \emph{clusters}) which are constructed by mutations. One of major results in the theory of cluster algebras is the Laurent phenomenon of cluster variables, which was proved by Fomin and Zelevinsky in \cite{FZ}. Their theorem says that any cluster variable is a Laurent polynomial of cluster variables in any cluster, with integer coefficients. In addition, they conjectured that the coefficients are non-negative. We mention some of known cases. For cluster algebras of finite type, the positivity conjecture with respect to a bipartite seed follows from a result of Fomin and Zelevinsky \cite[Corollary 11.7]{FZ4}. Musiker, Schiffler, and Williams \cite{MSW} found combinatorial formulas for the Laurent expansion of any cluster variable in any cluster algebra coming from a triangulated surface (with or without punctures) with respect to an arbitrary seed, which imply positivity of the Laurent expansion. For acyclic cluster algebras, Caldero and Reineke \cite{CR} made a significant progress, and Qin \cite{Q} and independently Nakajima \cite{N} who used the idea of Hernandez and Leclerc \cite{HL} proved the positivity conjecture for the special case of an initial seed. For rank 2 cluster algebras, Dupont \cite{D} showed the positivity by using the results for acyclic cluster algebras. For acyclic rank 2 cluster algebras, neither elementary proof nor combinatorial formula for the positive coefficients has been known except for the case of $r=2$ \cite{SZ}, \cite{MP}, \cite{CZ}.
In this note we prove Conjecture~\ref{Kontconj}. Actually we show that for $r\geq 2$ and $n>0$, $x_n$ is a non-commutative Laurent polynomial of $x$ and $y$ with non-negative integer coefficients. The case of $n<0$ can be taken care of by considering the anti-automorphism of $k\langle x_1^{\pm1}, x_2^{\pm1}\rangle$ as in \cite[Lemma 7]{BR}. Since $F_r^{-1}$ is obtained by interchanging $x$ and $y$ in (\ref{Kont_map}), the statement for $y_n$ easily follows from the one for $x_n$. Once we establish an explicit formula for $x_n$ (see Theorem~\ref{mainthm}), proofs are straightforward. Furthermore, it is not hard to show that the formula implies that every nonzero coefficient is equal to 1, hence it gives a combinatorial formula for the specialization of $x_n$ at $xy=q^eyx$ for any integer $e$ including 0. Our proof is completely elementary.
\noindent\emph{Acknowledgement.} We are grateful to Professors Sergey Fomin and Rob Lazarsfeld for their valuable advice. We also thank Professors Philippe Di Francesco, David Hernandez, Vladimir Retakh and Ralf Schiffler for their useful suggestions. Retakh pointed out that the argument also works for $r=2$ and suggested to compare with known formulas. Di Francesco clarified the definition of the Kontsevich automorphism and Definition 3 and 5.
\section{Main Result}
Throughout the notes, we fix $r\geq 2$ and let $F=F_r$. We will always use the following presentation for monomials in $K$:
$$y^{\beta_0}x^{\alpha_1}y^{\beta_1}x^{\alpha_2}y^{\beta_2}\cdots x^{\alpha_{m-1}}y^{\beta_{m-1}} x^{\alpha_m}\longleftrightarrow \left(\begin{array}{cccccc} \text{ }&\alpha_1 & \alpha_2 & \cdots & \alpha_{m-1}& \alpha_m\\ \beta_0 & \beta_1 & \beta_2 & \cdots & \beta_{m-1} & \text{ } \end{array} \right).$$
These expressions are not necessarily minimal, i.e. some of $\alpha_i$ or $\beta_i$ are allowed to be zeroes. As a matter of fact, zeroes serve as part of a potential for cancellation with $(1+y^r)^{-1}$, hence we almost never use minimal presentations in these notes. The multiplication of two monomials are given in the obvious way:
\tiny{$$\left(\begin{array}{ccccc} \text{ }&\alpha_1 & \cdots & \alpha_{m-1}& \alpha_m\\ \beta_0 & \beta_1 & \cdots & \beta_{m-1} & \text{ } \end{array} \right)\left(\begin{array}{ccccc} \text{ }&\alpha_{m+1} & \cdots & \alpha_{m+l-1}& \alpha_{m+l}\\ \beta_m & \beta_{m+1} & \cdots & \beta_{m+l-1} & \text{ } \end{array} \right)=\left(\begin{array}{ccccc} \text{ }&\alpha_1 & \cdots & \alpha_{m+l-1}& \alpha_{m+l}\\ \beta_0 & \beta_1 & \cdots & \beta_{m+l-1} & \text{ } \end{array} \right).$$}
\normalsize{Occasionally} we think of $\left(\begin{array}{ccccc} \text{ }&\alpha_1 & \cdots & \alpha_{m+l-1}& \alpha_{m+l}\\ \beta_0 & \beta_1 & \cdots & \beta_{m+l-1} & \text{ } \end{array} \right)$ as
$$\left(\begin{array}{ccccc} \text{ }&\alpha_1 & \cdots & \alpha_{m-1}& \alpha_m\\ \beta_0 & \beta_1 & \cdots & \beta_{m-1} & \beta_{m}\end{array} \right)\left(\begin{array}{cccc} \alpha_{m+1} & \cdots & \alpha_{m+l-1}& \alpha_{m+l}\\ \beta_{m+1} & \cdots & \beta_{m+l-1} & \text{ } \end{array} \right).$$
If we have a map $G : (x, y) \mapsto (xyx^{-1}, y^rx^{-1})$ and define $(z_n, w_n):=G^n(x, y)$, then it is easy to see that $z_n$ is one of the terms in $x_n$. Since we will express $x_n$ as the sum of monomials with the same number of columns as $z_n$ has, we want to find an expression for $z_n$, which requires the following definition.
\begin{defn}
Let $\{c_n\}$ be the sequence defined by the recurrence relation $$c_n=rc_{n-1} -c_{n-2},$$ with the initial condition $c_1=0$, $c_2=1$. Actually, when $r>2$, it is easy to see that
$$
c_n= \frac{1}{\sqrt{r^2-4} }\left(\frac{r+\sqrt{r^2-4}}{2}\right)^n - \frac{1}{\sqrt{r^2-4} }\left(\frac{r-\sqrt{r^2-4}}{2}\right)^n = \sum_{i\geq 0} (-1)^i { {n-2-i} \choose i }r^{n-2-2i}.
$$ \qed
\end{defn}
Then we have
\begin{equation}\label{z_n}z_n=\left(\begin{array}{ccccccc} \text{ }&1 & -1 & -1 & \cdots & -1& -1\\ 0&1& b_{n,1} & b_{n, 2} & \cdots & b_{n, c_{n}} & \text{ }\end{array} \right)\end{equation}
for some integers $b_{n,1},\cdots, b_{n, c_{n}}$, which are defined as follows.
\begin{defn}
The sequence $\{b_{i,j}\}_{i,j\in \mathbb{Z}_{>0}}$ is recursively defined by:
$$b_{i,j}=0\text{ for }j> c_i,$$
$$b_{2,1}=r-1, \text{ and}$$
$b_{n,1}=r-1,\, b_{n, p}=r \,\,\,\, (2+\sum_{i=1}^{j-1} b_{n-1,i}\leq p\leq\sum_{i=1}^j b_{n-1,i}),\, b_{n, 1+\sum_{i=1}^j b_{n-1,i}}=r-1$ for $n\geq 3$ and $1\leq j\leq c_{n-1}.$
It is easy to see that $1+\sum_{i=1}^{c_{n-1}} b_{n-1,i}=c_n$. The reason that (\ref{z_n}) holds is due to the following calculation:$$G(y^t x^{-1})=(y^rx^{-1})^{t}xy^{-1}x^{-1}=(y^rx^{-1})^{t-1}y^{r-1}x^{-1}\text{ for }t\geq 1.$$
\qed\end{defn}
\begin{exmp}
Let $r=3$. Then
$$
z_1=\left(\begin{array}{ccc} \text{ }&1 & -1 \\ 0&1& \text{ }\end{array} \right),
$$
$$z_2=\left(\begin{array}{cccc} \text{ }&1 & -1 & -1\\ 0&1& 2 & \text{ }\end{array} \right),$$
$$z_3=\left(\begin{array}{cccccc} \text{ }&1 & -1 & -1& -1 & -1\\ 0&1& 2 & 3 & 2 & \text{ }\end{array} \right), \text{ and }$$
$$z_4=\left(\begin{array}{ccccccccccc} \text{ }&1 & -1 & -1& -1 & -1 & -1 & -1& -1 & -1& -1 \\ 0&1& 2 & 3 & 2 & 3 & 3 & 2 & 3 & 2 &\text{ }\end{array} \right).$$ \qed
\end{exmp}
We want to express $x_n$ as
$$
x_n = \sum_{\alpha_i, \beta_i} \left(\begin{array}{ccccccc} \text{ }&1 & \alpha_1 & \alpha_2 & \cdots & \alpha_{c_n}& -1\\ 0&1& \beta_{1} & \beta_{2} & \cdots & \beta_{c_{n}} & \text{ }\end{array} \right),
$$
where the summation runs over $\alpha_i, \beta_i$ satisfying certain conditions. We believe that this expression is better than the minimal expression. We need to be able to locate zeroes in a systematic way. For this purpose, we will use the following notations.
\begin{defn}
Define a transformation $g$ on the set $\{(w_1,\cdots,w_j)\,| \, j\in \mathbb{Z}_{>0}, w_i=r\text{ or }r-1 \}$ of strings by
$\,\,\,\,\,\,\,\,\,\,\,\,\,\begin{array}{cc} g(r)= & \underbrace{(r,\cdots,r,r-1)} \\ \text{ } & r\text{ numbers}\end{array}$, \,\,\,\,\,\, $\begin{array}{cc} g(r-1)= & \underbrace{(r,\cdots,r,r-1),} \\ \text{ } & (r-1)\text{ numbers}\end{array}$, \,\,\,\,\,\, and
$$g^s(w_1,\cdots,w_i, w_{i+1},\cdots,w_j)=g^s(w_1,\cdots,w_i)g^s(w_{i+1},\cdots,w_j)\text{ for any integers }i,j,s\geq 0,$$where $g^s=\begin{array}{c}\underbrace{g \circ g \circ \cdots \circ g}\\ s \end{array}$.
Then we define the set of exceptional strings by $$\text{Exc}:=\{g^{s_1}(r)\cdots g^{s_j}(r) \,|\, j\in \mathbb{Z}_{>0}, \, s_1,...,s_j\in \mathbb{Z}_{\geq 0} \}.$$
\end{defn}
\begin{rem}\label{Excrem}
For any $\mathbf{w}=(w_1,\cdots,w_j)\in \text{Exc}$ and any $i$ $(1\leq i\leq j)$, its substring $(w_1,\cdots,w_i)$ consisting of the first $i$ numbers in $\mathbf{w}$ also belongs to $\text{Exc}$.\end{rem}
\begin{exmp}
Let $r=3$. Then $g^0(3)=(3),\, g^0(3)g^0(3)=(3,3), \,g(3)=(3,3,2),\,g(3)g(3)=(3,3,2,3,3,2),\,\text{ and }g^2(3)=(3,3,2,3,3,2,3,2)$ belong to $\text{Exc}$. But $(2)$ and $(3,3,2,3,2)$ do not.
\qed\end{exmp}
Let $[m]$ denote the set $\{1,2,\cdots,m\}$ for any nonnegative integer $m$.
\begin{defn}\label{def_of_f}
We need a function $f$ from $\{\text{subsets of } [c_{n-1}]\}$ to $\{\text{subsets of } [c_{n}]\}$. For each subset $V\subset [c_{n-1}]$, we define $f(V)$ as follows.
If $V=\emptyset$ then $f(\emptyset)=\emptyset$. If $V\neq\emptyset$ then we write $V=\cup_{i=1}^j\{e_i,e_i+1,\cdots, e_i+l_i-1\}$ with $l_i>0$ $(1\leq i\leq j)$ and $e_i+l_i<e_{i+1}$ $(1\leq i\leq j-1)$. For each $1\leq i\leq j$, define $f_i(V)$ by
$$
f_i(V):=\left\{ \begin{array}{ll}
[1+\sum_{k=1}^{e_i+l_i-1}b_{n-1,k}]\setminus [1+\sum_{k=1}^{e_i-1}b_{n-1,k}], & \text{ if }(b_{n-1,e_i},\cdots,b_{n-1,e_i+l_i-1})\in \text{Exc} \\
\text{ } & \text{ }\\
{[}1+\sum_{k=1}^{e_i+l_i-1}b_{n-1,k}{]}\setminus [\sum_{k=1}^{e_i-1}b_{n-1,k}], & \text{ otherwise.} \end{array} \right.
$$
Then $f(V)$ is obtained by taking the union of $f_i(V)$'s:
$$
f(V):=\cup_{i=1}^j f_i(V).
$$
Note that not every subset of $[c_n]$ can be realized as $f(V)$.
\qed\end{defn}
\begin{exmp}
Let $r=3$. If $n=3$ then $[c_2]=\{1\}$, and $f(\emptyset)=\emptyset\subset [c_3]$ and $f(\{1\})=\{1,2,3\}$.
If $n=4$ then $[c_3]=\{1,2,3\}$. In this case, $f(\{2\})=\{4,5,6\}$ since $(b_{3,2})=(3)\in \text{Exc}$. On the other hand, $f(\{1\}\cup\{3\})=\{1,2,3\}\cup\{6,7,8\}\subset [c_4]$.
\qed\end{exmp}
\begin{defn}\label{restrict_to_term}
Let $$T=\sum_{\delta_1\in H_1}\cdots\sum_{\delta_{c_n}\in H_{c_n}} \left(\begin{array}{ccccccc} \text{ }&1 & \alpha_1& \alpha_2 & \cdots & \alpha_{c_n}& -1\\ 0&1& \beta_{1}-\delta_{1}r & \beta_{2}-\delta_{2}r & \cdots & \beta_{c_{n}}-\delta_{c_n}r & \text{ }\end{array} \right),$$ where each of $H_i$ is either $\{0\}$ or $\{0,1\}$. Let $V$ be any subset of $[c_n]$. Then we define
$$T|_V:=\left\{\begin{array}{ll} 0, & \text{ if } H_i=\{0\} \text{ for some } i\in V \\
\text{ } & \text{ }\\
\begin{array}{l}\left(\begin{array}{ccccccc} \text{ }&1 & \alpha_1 & \alpha_2 & \cdots & \alpha_{c_n}& -1\\ 0&1& \beta_{1}-\delta_{1}r & \beta_{2}-\delta_{2}r & \cdots & \beta_{c_{n}}-\delta_{c_n}r & \text{ }\end{array} \right) \\\text{ where }\delta_{i}=1\text{ for }i\in V\text{ and }\delta_{i}=0\text{ for }i\not\in V, \end{array}
& \text{ otherwise.}\end{array}\right.$$
Note that $T=\sum_{V\subset [c_n]} T|_V$.
In the proof of Lemma~\ref{mainlemma}, we will use the following minor modification. We let $T=T_1T_2$, where
$$T_1=\sum_{\delta_1\in H_1}\cdots\sum_{\delta_{m}\in H_{m}} \left(\begin{array}{ccccccc} \text{ }&1 & \alpha_1 & \alpha_2 & \cdots & \alpha_{m}&\alpha_{m+1}\\ 0&1& \beta_{1}-\delta_{1}r & \beta_{2}-\delta_{2}r & \cdots & \beta_{m}-\delta_{m}r \end{array} \right)$$and
$$T_2=\sum_{\delta_{m+1}\in H_{m+1}}\cdots\sum_{\delta_{c_n}\in H_{c_n}} \left(\begin{array}{ccccc} & \alpha_{m+2} & \cdots & \alpha_{c_n}& -1\\ \beta_{m+1}-\delta_{m+1}r & \beta_{m+2}-\delta_{m+2}r & \cdots & \beta_{c_{n}}-\delta_{c_n}r & \text{ }\end{array} \right).$$ Let $V$ be any subset of $[c_n]$ with $V=V_1\cup V_2$ $(V_1\subset [m], V_2 \subset [c_n]\setminus [m])$. By abuse of notation, we say that
$$
T|_V=T_1|_{V_1}T_2|_{V_2}.
$$
\qed\end{defn}
It is easy to see that
$$F(z_{n-2})=\sum_{\delta_1=0}^1\cdots\sum_{\delta_{c_{n-1}}=0}^1\left(\begin{array}{ccccccc} \text{ }&1 & -1 & -1 & \cdots & -1& -1\\ 0&1& b_{n-1,1}-\delta_1r & b_{n-1, 2}-\delta_2r & \cdots & b_{n-1, c_{n-1}}-\delta_{c_{n-1}}r & \text{ }\end{array} \right).$$ Some examples of $F(z_{n-2})|_V$ are given in Example~\ref{main_example1} and~\ref{main_example2}.
For monomials $A$ and $A_i$ in $K$, we let $\prod_{i=1}^m A_i$ denote $A_1A_2\cdots A_m$, and $A^m$ denote $\begin{array}{c}\underbrace{AA\cdots A}\\m \end{array}$.
\begin{defn}\label{def_of_tildeF}
Let $$w={F}(z_{n-1})|_V= \left(\begin{array}{ccccccc} \text{ }&1 & -1 & -1 & \cdots & -1& -1\\ 0&1& b_{n,1}-\delta_{1}r & b_{n,2}-\delta_{2}r & \cdots & b_{n,c_{n}}-\delta_{c_n}r & \text{ }\end{array} \right),$$
where $\delta_{i}=1$ for $i\in V$ and $\delta_{i}=0$ for $i\not\in V$. We will define $\tilde{F}(w)$, which is a modification of taking $F$, where we do not allow $(1+y^r)^{-1}$ to appear. In fact, if none of $b_{n,i}-\delta_{i}r$ is negative, then $\tilde{F}(w)=F(w)$. Remember that $b_{n,1}=r-1$ and that $b_{n,i}$ $(1<i\leq c_n)$ is either $r$ or $r-1$. In general, $\tilde{F}(w)$ can be defined by letting
$$\tilde{F}(w)=\tilde{F}\left(\begin{array}{ccc} \text{ }&1 & -1 \\ 0&1& r-1-\delta_{1}r \end{array} \right)\left[\prod_{i=2}^{c_n} \tilde{F}\left(\begin{array}{c} -1\\b_{n,i}-\delta_{i}r \end{array} \right)\right]\left(\begin{array}{c} -1\\ \text{ } \end{array} \right)$$
and using the following rules:
\tiny{$$\aligned &\tilde{F}\left(\begin{array}{ccc} \text{ }&1 & -1\\ 0&1 & r-1-\delta_1 r \end{array} \right)\\&:=\sum_{\delta^{(1)}=0}^{1-\delta_{1}}\cdots\sum_{\delta^{(r-1)}=0}^{1-\delta_{1}}\sum_{\delta^{(r)}=0}^{\Delta_{V,1}(1-\delta_{1})}\begin{array}{c}\underbrace{\left(\begin{array}{cccccccc} \text{ }&1& -1& \delta_1-1 & \cdots & \delta_1-1& 2\delta_1-1\\ 0& 1& (1-\delta^{(1)})r-1 & (1-\delta^{(2)})r& \cdots & (1-\delta^{(r-1)})r & \Delta_{V,1}(1-\delta^{(r)})r-1 \end{array}\right)},\\ r+2 \text{ columns} \end{array}\endaligned$$}
$$
\tilde{F}\left(\begin{array}{c} -1\\b_{n,i}-\delta_{i}r \end{array} \right):=\sum_{\delta^{(1)}=0}^{1-\delta_{i}}\cdots\sum_{\delta^{(b_{n,i}-1)}=0}^{1-\delta_{i}}\sum_{\delta^{(b_{n,i})}=0}^{\Delta_{V,i}(1-\delta_{i})}\begin{array}{c} \underbrace{\left( \begin{array}{cccc} \delta_i-1 & \cdots & \delta_i-1 & (r-b_{n,i}+1)\delta_i-1 \\ (1-\delta^{(1)})r &\cdots& (1-\delta^{(b_{n,i}-1)})r & \Delta_{V,i}(1-\delta^{(b_{n,i})})r -1\end{array} \right)} \\ b_{n,i} \text{ columns }\end{array} \text{ for }1<i\leq c_n,
$$
where
$\tiny{\Delta_{V,i}=\left\{\begin{array}{ll} 0, &\text{ if }\delta_i=0, \delta_{i+1}=1,\text{ and }(b_{n,i+1}, \cdots, b_{n,i+e})\not\in \text{Exc}\text{ with }e
\text{ satisfying }\{i+1,\cdots, i+e\}\subset V\text{ and }i+e+1\not\in V\\
1, & \text{ otherwise}. \end{array} \right.}$
\end{defn}
We are ready to state our main result.
\begin{thm}\label{mainthm}
Let $n\geq 3$. Let $z_{n-2}$ be the monomial corresponding to $$\left(\begin{array}{ccccccc} \text{ }& 1 & -1 & -1 & \cdots & -1& -1\\ 0&1& b_{n-2,1} & b_{n-2,2} & \cdots & b_{n-2,c_{n-2}} & \text{ }\end{array} \right).$$Then
\begin{equation}\label{maineq}x_n=\sum_{V\subset [c_{n-1}]} \left(\sum_{W\subset [c_n]\setminus f(V)} \tilde{F}({F}(z_{n-2})|_V)|_W\right)=\sum_{V\subset [c_{n-1}]} \left( \tilde{F}({F}(z_{n-2})|_V)\right). \end{equation}
\end{thm}
In addition to our theoretical proof, this formula is also checked by a computer up to the limit of practical computation ($r\leq 4$ and $n\leq 6$).
\begin{exmp}\label{main_example1}
Let $r=3$. Then $$F(z_1)|_{\emptyset}=\left(\begin{array}{cccc}\text{ }& 1 & -1 & -1\\0& 1&2 & \text{ } \end{array} \right)\,\,\,\,\,\,\text{ and }\,\,\,\,\,\,\,F(z_1)|_{\{1\}}=\left(\begin{array}{cccc}\text{ }& 1 & -1 & -1\\0& 1&-1 & \text{ } \end{array} \right),$$ which yield
$$\tilde{F}(F(z_1)|_{\emptyset})|_{\emptyset}=\left(\begin{array}{cccccc}\text{ }& 1 & -1 & -1&-1&-1\\0& 1&2 &3&2& \text{ } \end{array} \right),\,\,\,\,\,\,\,\,\,\,\,\,\,\tilde{F}(F(z_1)|_{\emptyset})|_{\{1,2\}}=\left(\begin{array}{cccccc}\text{ }& 1 & -1 & -1&-1&-1\\0& 1&-1 &0&2& \text{ } \end{array} \right),$$
$$\tilde{F}(F(z_1)|_{\emptyset})|_{\{1\}}=\left(\begin{array}{cccccc}\text{ }& 1 & -1 & -1&-1&-1\\0& 1&-1 &3&2& \text{ } \end{array} \right),\,\,\,\,\,\,\,\,\,\,\,\,\,\tilde{F}(F(z_1)|_{\emptyset})|_{\{1,3\}}=\left(\begin{array}{cccccc}\text{ }& 1 & -1 & -1&-1&-1\\0& 1&-1 &3&-1& \text{ } \end{array} \right),$$
$$\tilde{F}(F(z_1)|_{\emptyset})|_{\{2\}}=\left(\begin{array}{cccccc}\text{ }& 1 & -1 & -1&-1&-1\\0& 1&2 &0&2& \text{ } \end{array} \right),\,\,\,\,\,\,\,\,\,\,\,\,\,\tilde{F}(F(z_1)|_{\emptyset})|_{\{2,3\}}=\left(\begin{array}{cccccc}\text{ }& 1 & -1 & -1&-1&-1\\0& 1&2 &0&-1& \text{ } \end{array} \right),$$
$$\tilde{F}(F(z_1)|_{\emptyset})|_{\{3\}}=\left(\begin{array}{cccccc}\text{ }& 1 & -1 & -1&-1&-1\\0& 1&2 &3&-1& \text{ } \end{array} \right),\,\,\,\,\,\,\,\,\,\,\,\,\,\tilde{F}(F(z_1)|_{\emptyset})|_{\{1,2,3\}}=\left(\begin{array}{cccccc}\text{ }& 1 & -1 & -1&-1&-1\\0& 1&-1 &0&-1& \text{ } \end{array} \right),$$
$$\text{ and }\tilde{F}(F(z_1)|_{\{1\}})|_{\emptyset}=\left(\begin{array}{cccccc}\text{ }& 1 & -1 &0&1&-1\\0& 1&-1 &0&-1& \text{ } \end{array} \right).$$
Hence $x_3$ has $9=\frac{2^3+1}{1}$ terms.
\end{exmp}
\begin{exmp}\label{main_example2}Let $r=3$.
$$\aligned F(z_2)|_{\emptyset}&=\left(\begin{array}{cccccc}\text{ }& 1 & -1 & -1&-1&-1\\0& 1&2 &3&2& \text{ } \end{array} \right)\\
F(z_2)|_{\{1\}}&=\left(\begin{array}{cccccc}\text{ }& 1 & -1 & -1&-1&-1\\0& 1&-1 &3&2& \text{ } \end{array} \right)\\
F(z_2)|_{\{2\}}&=\left(\begin{array}{cccccc}\text{ }& 1 & -1 & -1&-1&-1\\0& 1&2 &0&2& \text{ } \end{array} \right)\\
F(z_2)|_{\{3\}}&=\left(\begin{array}{cccccc}& 1 & -1 & -1&-1&-1\\0& 1&2 &3&-1& \end{array} \right)\endaligned\,\,\,\,\,\,\,\,\,\,\,\,\, \aligned
F(z_2)|_{\{1,2\}}&=\left(\begin{array}{cccccc}& 1 & -1 & -1&-1&-1\\0& 1&-1 &0&2& \end{array} \right)\\
F(z_2)|_{\{1,3\}}&=\left(\begin{array}{cccccc}& 1 & -1 & -1&-1&-1\\0& 1&-1 &3&-1& \end{array} \right)\\
F(z_2)|_{\{2,3\}}&=\left(\begin{array}{cccccc}& 1 & -1 & -1&-1&-1\\0& 1&2 &0&-1& \end{array} \right)\\
F(z_2)|_{\{1,2,3\}}&=\left(\begin{array}{cccccc}& 1 & -1 & -1&-1&-1\\0& 1&-1 &0&-1& \end{array} \right)
\endaligned $$
\tiny{
$$\aligned \tilde{F}(F(z_2)|_{\emptyset})=\sum_{\delta_1=0}^1\cdots\sum_{\delta_8=0}^1&\left(\begin{array}{cccccccccccc}\text{ }& 1 & -1 & -1&-1& -1&-1&-1& -1&-1&-1\\0& 1&2-3\delta_1 &3-3\delta_2&2-3\delta_3& 3-3\delta_4&3-3\delta_5&2-3\delta_6&3-3\delta_7&2-3\delta_8& \text{ } \end{array} \right)\\
\tilde{F}(F(z_2)|_{\{1\}})=\sum_{\delta_4=0}^1\cdots\sum_{\delta_8=0}^1&\left(\begin{array}{cccccccccccc}\text{ }& 1 & -1 & 0&1&-1& -1&-1&-1&-1& -1\\0& 1&2-3\cdot1 &3-3\cdot1&2-3\cdot1& 3-3\delta_4&3-3\delta_5&2-3\delta_6&3-3\delta_7&2-3\delta_8& \text{ } \end{array} \right)\\
\tilde{F}(F(z_2)|_{\{2\}})=\sum_{\delta_1=0}^1\cdots\sum_{\delta_8=0}^1&\left(\begin{array}{cccccccccccc}\text{ }& 1 & -1 & -1&-1&0& 0&0&-1& -1&-1\\0& 1&2-3\delta_1 &3-3\delta_2&2-3\delta_3& 3-3\cdot1&3-3\cdot1&2-3\cdot1&3-3\delta_7&2-3\delta_8& \text{ }\end{array} \right)\\
\tilde{F}(F(z_2)|_{\{3\}})=\sum_{\delta_1=0}^1\cdots\sum_{\delta_5=0}^1&\left(\begin{array}{cccccccccccc}\text{ }& 1 & -1 & -1&-1& -1&-1&-1& 0&1&-1\\0& 1&2-3\delta_1 &3-3\delta_2&2-3\delta_3& 3-3\delta_4&3-3\delta_5&2-3\cdot1&3-3\cdot1&2-3\cdot1& \text{ } \end{array} \right)\\
\tilde{F}(F(z_2)|_{\{1,2\}})=\sum_{\delta_7=0}^1\sum_{\delta_8=0}^1&\left(\begin{array}{cccccccccccc}\text{ }& 1 & -1 & 0&1&0& 0&0&-1& -1&-1\\0& 1&2-3\cdot1 &3-3\cdot1&2-3\cdot1& 3-3\cdot1&3-3\cdot1&2-3\cdot1&3-3\delta_7&2-3\delta_8& \text{ }\end{array} \right)\\
\tilde{F}(F(z_2)|_{\{1,3\}})=\sum_{\delta_4=0}^1\sum_{\delta_5=0}^1&\left(\begin{array}{cccccccccccc}\text{ }& 1 & -1 & 0&1& -1&-1&-1& 0&1&-1\\0& 1&2-3\cdot1 &3-3\cdot1&2-3\cdot1& 3-3\delta_4&3-3\delta_5&2-3\cdot1&3-3\cdot1&2-3\cdot1& \text{ } \end{array} \right)\\
\tilde{F}(F(z_2)|_{\{2,3\}})=\sum_{\delta_1=0}^1\sum_{\delta_2=0}^1&\left(\begin{array}{cccccccccccc}\text{ }& 1 & -1 & -1&-1& 0&0&0& 0&1&-1\\0& 1&2-3\delta_1 &3-3\delta_2&2-3\cdot1& 3-3\cdot1&3-3\cdot1&2-3\cdot1&3-3\cdot1&2-3\cdot1& \text{ } \end{array} \right)\\
\tilde{F}(F(z_2)|_{\{1,2,3\}})=&\left(\begin{array}{cccccccccccc}\text{ }& 1 & -1 & 0&1& 0&0&0& 0&1&-1\\0& 1&2-3\cdot1 &3-3\cdot1&2-3\cdot1& 3-3\cdot1&3-3\cdot1&2-3\cdot1&3-3\cdot1&2-3\cdot1& \text{ } \end{array} \right)
\endaligned$$
}
\normalsize{Hence $x_4$ has $(2^8+3\cdot 2^5+ 3\cdot 2^2+1\cdot 2^0)=\frac{(\frac{2^3+1}{1})^3+1}{2}$ terms.}
\end{exmp}
\section{Proofs}
\begin{lem}\label{mainlemma}
Let $W'$ be any subset of $[c_{n}]$. Then
$$
{F}\left( \sum_{V: f(V)\subset W'} \tilde{F}(F(z_{n-2})|_V)|_{W'\setminus f(V)} \right)
=\tilde{F}({F}(z_{n-1})|_{W'}).
$$
\end{lem}
\begin{lem}
Lemma~\ref{mainlemma} implies Theorem~\ref{mainthm}.
\end{lem}
\begin{proof}
The second equality in $(\ref{maineq})$ follows from Definition~\ref{restrict_to_term}. By induction on $n$, we assume that Theorem~\ref{mainthm} holds for $n$. Then
$$\aligned x_{n+1}
&=F(x_{n})\\
&= {F}\left( \sum_{V\subset [c_{n-1}]} \left(\sum_{W\subset [c_n]\setminus f(V)} \tilde{F}({F}(z_{n-2})|_V)|_W\right) \right)\\
&= {F}\left( \sum_{W\subset [c_n]} \left(\sum_{V: W\cap f(V)=\emptyset} \tilde{F}({F}(z_{n-2})|_V)|_W\right) \right)\\
&\overset{(*)}={F}\left( \sum_{W'\subset [c_n]} \left(\sum_{V:f(V)\subset W'} \tilde{F}({F}(z_{n-2})|_V)|_{W'\setminus f(V)}\right) \right)\\
&= \sum_{W'\subset [c_n]} {F}\left(\sum_{V:f(V)\subset W'} \tilde{F}({F}(z_{n-2})|_V)|_{W'\setminus f(V)}\right) \\
&= \sum_{W'\subset [c_n]}\tilde{F}({F}(z_{n-1})|_{W'}), \endaligned$$where the last equality follows from Lemma~\ref{mainlemma}.
To show $(*)$, we give a bijection from
$$\{(V,W) \,|\, W\subset [c_n], W\cap f(V)=\emptyset \}$$
to
$$
\{(V,W') \,|\, W'\subset [c_n], V:f(V)\subset W' \},
$$
which is defined by $(V,W)\mapsto (V,W\cup f(V))$ and its inverse $(V,W')\mapsto (V, W'\setminus f(V))$.
\end{proof}
\begin{proof}[Sketch of proof of Lemma~\ref{mainlemma}]
If $W'=\emptyset$ then the statement is an easy consequence of Definition~\ref{def_of_tildeF} together with $F(z_{n-2})|_{\emptyset}=z_{n-1}$. If $W'\neq\emptyset$, then there exist integers $j,e_i, l_i>0$ such that $W'=\cup_{i=1}^j\{e_i,e_i+1,\cdots, e_i+l_i-1\}$ with $e_i+l_i<e_{i+1}$ for all $1\leq i\leq j-1$. We use induction on $j$. We prove the base case of $j=1$ in the next Lemma, which has an essential idea. In fact, once we establish the base case, then the case of $j>1$ is straightforward. Here we will sketch how the former case implies the latter case.
Suppose that $j>1$. Then there is an integer $m'$ such that $\min W'< m'< \max W'$ but $m'+1\not\in W'$.
Let $W'_1=\{i\in W' \,|\, i<m'+1\}$ and $W'_2=\{i\in W' \,|\, i>m'+1\}$. Let $m$ be the integer satisfying $f(f([m]))\subset [m']$ but $f(f([m+1]))\not\subset [m']$.
Let $$(z_{n-2})_1=\left(\begin{array}{ccccccc}& 1 & -1 & -1 & \cdots & -1&-1\\ 0&1& b_{n-2,1} & b_{n-2,2} & \cdots & b_{n-2,m}&\end{array} \right),$$ and $$(z_{n-2})_2=\left(\begin{array}{ccccc} & -1 & \cdots & -1& -1\\ b_{n-2,m+1}& b_{n-2,m+2} & \cdots & b_{n-2,c_{n-2}} & \end{array} \right).$$
Let $$(z_{n-1})_1=\left(\begin{array}{ccccccc}& 1 & -1 & -1 & \cdots & -1&-1\\ 0&1& b_{n-1,1} & b_{n-1,2} & \cdots & b_{n-1,1+\sum_{i=1}^m b_{n-2,i}}&\end{array} \right),$$ and $$(z_{n-1})_2=\left(\begin{array}{ccccc} & -1 & \cdots & -1& -1\\ b_{n-1,2+\sum_{i=1}^m b_{n-2,i}}& b_{n-1,3+\sum_{i=1}^m b_{n-2,i}} & \cdots & b_{n-1,c_{n-1}} & \end{array} \right).$$
Then
$$\aligned
&{F}\left( \sum_{V: f(V)\subset W'} \tilde{F}(F(z_{n-2})|_V)|_{W'\setminus f(V)} \right)\\
&={F}\left( \sum_{V: f(V)\subset W'_1\cup W'_2} \tilde{F}(F(z_{n-2})|_V)|_{ (W'_i\cup W'_2)\setminus f(V)} \right)\\
&={F}\left( \sum_{V_1: f(V_1)\subset W'_1} \,\,\,\,\,\sum_{V_2: f(V_2)\subset W'_2}\tilde{F}(F((z_{n-2})_1 (z_{n-2})_2)|_{V_1\cup V_2})|_{(W'_1\setminus f(V_1))\cup (W'_2\setminus f(V_2))} \right)\\
&={F}\left( \sum_{V_1: f(V_1)\subset W'_1} \tilde{F}(F((z_{n-2})_1)|_{V_1})|_{W'_1\setminus f(V_1)} \right){F}\left( \sum_{V_2: f(V_2)\subset W'_2} \tilde{F}(F((z_{n-2})_2)|_{V_2})|_{W'_2\setminus f(V_2)} \right)\\
&=\tilde{F}({F}((z_{n-1})_1)|_{W'_1})\tilde{F}({F}((z_{n-1})_2)|_{W'_2})\,\,\,\,\,\,\,\,\,\,(\text{this is implied by induction on }j \text{ and abuse of }\tilde{F})\\
&=\tilde{F}({F}(z_{n-1})|_{W'}).\\
\endaligned$$
\end{proof}
\begin{lem}\label{sublemma}
Let $W'$ be a subset of $[c_{n}]$, which consists of consecutive numbers. Then
$$
{F}\left( \sum_{V: f(V)\subset W'} \tilde{F}(F(z_{n-2})|_V)|_{W'\setminus f(V)} \right)
=\tilde{F}({F}(z_{n-1})|_{W'}).
$$
\end{lem}
\begin{proof}
The reader is recommended to see Example~\ref{exmp3}.
Let $V^M=\cup_{f(V)\subset W'} V$. Since $W'\subset [c_{n}]$ is a set of consecutive numbers, Definition~\ref{def_of_f} implies that $V^M$ also consists of consecutive numbers. Write $$V^M= \{e,e+1,\cdots,e+l-1\}.$$ Assume that $e>1$ (The case of $e=1$ can be done in a similar manner). For each $V\subset V^M$, we have
$$\aligned F(z_{n-2})|_V=&\left(\begin{array}{ccc} \text{ }&1 & -1 \\ 0&1& r-1 \end{array} \right)\left[\prod_{i=2}^{e-1} \left(\begin{array}{c} -1\\b_{n-1,i} \end{array} \right)\right]\\
&\times \left[\prod_{i=e}^{e+l-1} \left(\begin{array}{c} -1\\b_{n-1,i}-\delta_{i}(V)r \end{array} \right)\right]\left[\prod_{i=e+l}^{c_{n-1}} \left(\begin{array}{c} -1\\b_{n-1,i} \end{array} \right)\right]\left(\begin{array}{c} -1\\ \text{ } \end{array} \right),\endaligned$$
where $\delta_{i}(V)=1$ for $i\in V$ and $\delta_{i}(V)=0$ for $i\not\in V$.
On the other hand, there are non-negative integers $s_1$ and $s_2$ such that $$W'=[s_2+1+\sum_{k=1}^{e+l-1}b_k]\setminus [-s_1+1+\sum_{k=1}^{e-1}b_k].$$ Note that we have $$1+\sum_{k=1}^{e-2}b_k< -s_1+2+\sum_{k=1}^{e-1}b_k$$ and $$s_2+1+\sum_{k=1}^{e+l-1}b_k< 1+\sum_{k=1}^{e+l}b_k,$$ because otherwise $\cup_{f(V)\subset W'} V\neq \{e,e+1,\cdots,e+l-1\}$. It is worth mentioning that $s_1=0$ implies $(b_{n-1,e},b_{n-1,e+1},\cdots,b_{n-1,e+l-1})\in \text{Exc}$.
If $s_1>0$ then the desired statement is obtained by the following computations:
$$\aligned
&{F}\left( \sum_{V: f(V)\subset W'} \tilde{F}(F(z_{n-2})|_V)|_{W'\setminus f(V)} \right)\\
&={F}\left( \sum_{V: f(V)\subset W'} \tilde{F}\left(\left(\begin{array}{ccc} \text{ }&1 & -1 \\ 0&1& r-1\end{array} \right)\left[\prod_{i=2}^{e-1} \left(\begin{array}{c} -1\\b_{n-1,i} \end{array} \right)\right] \right.\right.\\
&\,\,\,\,\,\,\,\,\,\,\,\times \left.\left.\left[\prod_{i=e}^{e+l-1} \left(\begin{array}{c} -1\\b_{n-1,i}-\delta_{i}(V)r \end{array} \right)\right]\left[\prod_{i=e+l}^{c_{n-1}} \left(\begin{array}{c} -1\\b_{n-1,i} \end{array} \right)\right]\left(\begin{array}{c} -1\\ \text{ } \end{array} \right) \right)|_{W'\setminus f(V)} \right)\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\endaligned$$
$$\aligned
&={F}\left( \sum_{V: f(V)\subset W'} \left(\sum_{j=1}^{r}\sum_{\delta^{(j)}=0}^{1}\begin{array}{c}\underbrace{\left(\begin{array}{cccccccc} \text{ }&1& -1&-1 & \cdots & -1& -1\\ 0& 1& (1-\delta^{(1)})r-1 & (1-\delta^{(2)})r& \cdots & (1-\delta^{(r-1)})r & (1-\delta^{(r)})r-1 \end{array}\right)}\\ r+2 \text{ columns} \end{array}\right.\right.\\
&\,\,\,\,\,\,\,\,\,\,\,\times\left[\prod_{i=2}^{e-2}\sum_{j=1}^{b_{n-1,i}}\sum_{\delta^{(j)}=0}^{1}\begin{array}{c} \underbrace{\left( \begin{array}{cccc} -1 & \cdots & -1 & -1 \\ (1-\delta^{(1)})r &\cdots& (1-\delta^{(b_{n-1,i}-1)})r & (1-\delta^{(b_{n-1,i})})r -1\end{array} \right)} \\ b_{n-1,i} \text{ columns }\end{array}\right] \\
&\,\,\,\,\,\,\,\,\,\,\,\times\sum_{j=1}^{b_{n-1,e-1}-1}\sum_{\delta^{(j)}=0}^{1}\sum_{\delta^{(b_{n-1,e-1})}=0}^{\Delta_{V,e-1}}\begin{array}{c} \underbrace{\left( \begin{array}{cccc} -1 & \cdots & -1 & -1 \\ (1-\delta^{(1)})r &\cdots& (1-\delta^{(b_{n-1,e-1}-1)})r & \Delta_{V,e-1}(1-\delta^{(b_{n-1,e-1})})r-1\end{array} \right)} \\ b_{n-1,e-1} \text{ columns }\end{array} \\
&\,\,\,\,\,\,\,\,\,\,\,\times\left[\prod_{i=e}^{e+l-1} \sum_{j=1}^{b_{n-1,i}-1}\sum_{\delta^{(j)}=0}^{1-\delta_{i}}\sum_{\delta^{(b_{n-1,i})}=0}^{\Delta_{V,i}(1-\delta_{i})}\begin{array}{c} \underbrace{\left( \begin{array}{cccc} \delta_i(V)-1 & \cdots & \delta_i(V)-1 & (r-b_{n-1,i}+1)\delta_i(V)-1 \\ (1-\delta^{(1)})r &\cdots& (1-\delta^{(b_{n-1,i}-1)})r & \Delta_{V,i}(1-\delta^{(b_{n-1,i})})r -1\end{array} \right)} \\ b_{n-1,i} \text{ columns }\end{array}\right]\\
&\,\,\,\,\,\,\times \left.\left.\left[\prod_{i=e+l}^{c_{n-1}} \sum_{j=1}^{b_{n-1,i}}\sum_{\delta^{(j)}=0}^{1}\begin{array}{c} \underbrace{\left( \begin{array}{cccc} -1 & \cdots & -1 & -1 \\ (1-\delta^{(1)})r &\cdots& (1-\delta^{(b_{n-1,i}-1)})r & (1-\delta^{(b_{n-1,i})})r -1\end{array} \right)} \\ b_{n-1,i} \text{ columns }\end{array}\right]\left(\begin{array}{c} -1\\ \text{ } \end{array} \right) \right)|_{W'\setminus f(V)} \right)
\endaligned$$\footnote{By abuse of notation, $\sum_{j=1}^{b_{n-1,i}-1}\sum_{\delta^{(j)}=0}^{1-\delta_{i}}$ means $\sum_{\delta^{(1)}=0}^{1-\delta_{i}}\cdots\sum_{\delta^{(b_{n-1,i}-1)}=0}^{1-\delta_{i}}$. These are made due to lack of space.}
$$\aligned
&={F}\left( \sum_{V: V\subset V^M} {\left(\begin{array}{ccc} \text{ }&1& -1\\ 0& 1& r-1 \end{array}\right)\left(\begin{array}{c} -1\\ r \end{array}\right)^{r-2}\left(\begin{array}{c}-1 \\ r-1 \end{array}\right) }\prod_{i=2}^{e-2}\left[{\left( \begin{array}{c} -1 \\ r \end{array} \right)^{b_{n-1,i}-1} \left( \begin{array}{c} -1 \\ r -1\end{array} \right)}\right] \right.\\
&\,\,\,\,\,\,\,\,\,\,\,\times\left(\begin{array}{c} -1 \\ r \end{array}\right)^{b_{n-1,e-1}-s_1} \begin{array}{c} \underbrace{\left( \begin{array}{cccc} -1 & \cdots & -1 & -1 \\ 0 &\cdots& 0 & -1\end{array} \right)}\\ s_1\text{ columns}\end{array} \\
&\,\,\,\,\,\,\,\,\,\,\,\times\prod_{i=e}^{e+l-1}\left[ {\left( \begin{array}{c} \delta_i(V)-1 \\ 0 \end{array} \right)^{b_{n-1,i}-1} \left( \begin{array}{c} (r-b_{n-1,i}+1)\delta_i(V)-1 \\ -1\end{array} \right)} \right]\\
&\,\,\,\,\,\,\,\,\,\,\,\times \left. \left( \begin{array}{c} -1 \\ 0\end{array} \right)^{s_2}\left( \begin{array}{c} -1 \\ r\end{array} \right)^{b_{n-1,e+l}-s_2-1} \left( \begin{array}{c} -1\\ r -1\end{array} \right)\prod_{i=e+l+1}^{c_{n-1}}\left[ {\left( \begin{array}{c} -1 \\ r\end{array} \right)^{b_{n-1,i}-1} \left( \begin{array}{c} -1 \\r -1\end{array} \right)} \right]\left(\begin{array}{c} -1\\ \text{ } \end{array} \right) \right)
\endaligned$$
$$\aligned
&={F}\left( {\left(\begin{array}{ccc} \text{ }&1& -1\\ 0& 1& r-1 \end{array}\right)\left(\begin{array}{c} -1\\ r \end{array}\right)^{r-2}\left(\begin{array}{c}-1 \\ r-1 \end{array}\right) }\prod_{i=2}^{e-2}\left[{\left( \begin{array}{c} -1 \\ r \end{array} \right)^{b_{n-1,i}-1} \left( \begin{array}{c} -1 \\ r -1\end{array} \right)}\right] \right.\\
&\,\,\,\,\,\,\,\,\,\,\,\times\left(\begin{array}{c} -1 \\ r \end{array}\right)^{b_{n-1,e-1}-s_1} \begin{array}{c} \underbrace{\left( \begin{array}{cccc} -1 & \cdots & -1 & -1 \\ 0 &\cdots& 0 & -1\end{array} \right)}\\ s_1\text{ columns}\end{array} \\
&\,\,\,\,\,\,\,\,\,\,\,\times\prod_{i=e}^{e+l-1} \sum_{\delta_i(V)=0}^{1}\left[ {\left( \begin{array}{c} \delta_i(V)-1 \\ 0 \end{array} \right)^{b_{n-1,i}-1} \left( \begin{array}{c} (r-b_{n-1,i}+1)\delta_i(V)-1 \\ -1\end{array} \right)} \right]\\
&\,\,\,\,\,\,\,\,\,\,\,\times \left. \left( \begin{array}{c} -1 \\ 0\end{array} \right)^{s_2}\left( \begin{array}{c} -1 \\ r\end{array} \right)^{b_{n-1,e+l}-s_2-1} \left( \begin{array}{c} -1\\ r -1\end{array} \right)\prod_{i=e+l+1}^{c_{n-1}}\left[ {\left( \begin{array}{c} -1 \\ r\end{array} \right)^{b_{n-1,i}-1} \left( \begin{array}{c} -1 \\r -1\end{array} \right)} \right]\left(\begin{array}{c} -1\\ \text{ } \end{array} \right) \right)
\endaligned$$
$$\aligned
&= {{F} \left[\left(\begin{array}{ccc} \text{ }&1& -1\\ 0& 1& r-1 \end{array}\right)\left(\begin{array}{c} -1\\ r \end{array}\right)^{r-2}\left(\begin{array}{c}-1 \\ r-1 \end{array}\right) \right] }\prod_{i=2}^{e-2}{F} \left[{\left( \begin{array}{c} -1 \\ r \end{array} \right)^{b_{n-1,i}-1} \left( \begin{array}{c} -1 \\ r -1\end{array} \right)}\right]\\
&\,\,\,\,\,\,\,\,\,\,\,\times{F}\left[ \left(\begin{array}{c} -1 \\ r \end{array}\right)^{b_{n-1,e-1}-s_1} \begin{array}{c} \underbrace{\left( \begin{array}{cccc} -1 & \cdots & -1 & -1 \\ 0 &\cdots& 0 & -1\end{array} \right)}\\ s_1\text{ columns}\end{array}\right] \\
&\,\,\,\,\,\,\,\,\,\,\,\times\prod_{i=e}^{e+l-1} \sum_{\delta_i(V)=0}^{1}{F} \left[ {\left( \begin{array}{c} \delta_i(V)-1 \\ 0 \end{array} \right)^{b_{n-1,i}-1} \left( \begin{array}{c} (r-b_{n-1,i}+1)\delta_i(V)-1 \\ -1\end{array} \right)} \right]\\
&\,\,\,\,\,\,\,\,\,\,\,\times {F}\left[ \left( \begin{array}{c} -1 \\ 0\end{array} \right)^{s_2}\left( \begin{array}{c} -1 \\ r\end{array} \right)^{b_{n-1,e+l}-s_2-1} \left( \begin{array}{c} -1\\ r -1\end{array} \right)\right]\prod_{i=e+l+1}^{c_{n-1}}{F} \left[ {\left( \begin{array}{c} -1 \\ r\end{array} \right)^{b_{n-1,i}-1} \left( \begin{array}{c} -1 \\r -1\end{array} \right)} \right]F\left(\begin{array}{c} -1\\ \text{ } \end{array} \right)
\endaligned$$ $(**)$\footnote{The equality under $(**)$ is a key step, where cancellation with $(1+y^r)^{-1}$ occurs.}
$$\aligned
&={{F} \left[\left(\begin{array}{ccc} \text{ }&1& -1\\ 0& 1& r-1 \end{array}\right)\left(\begin{array}{c} -1\\ r \end{array}\right)^{r-2}\left(\begin{array}{c}-1 \\ r-1 \end{array}\right) \right] }\prod_{i=2}^{e-2}{F} \left[{\left( \begin{array}{c} -1 \\ r \end{array} \right)^{b_{n-1,i}-1} \left( \begin{array}{c} -1 \\ r -1\end{array} \right)}\right]\\
&\,\,\,\,\,\,\,\,\,\,\,\times{F} \left[\left(\begin{array}{c} -1 \\ r \end{array}\right)^{b_{n-1,e-1}-s_1} \begin{array}{c} \underbrace{\left( \begin{array}{cccc} -1 & \cdots & -1 & -1 \\ 0 &\cdots& 0 & -1\end{array} \right)}\\ s_1\text{ columns}\end{array} \right] \\
&\,\,\,\,\,\,\,\,\,\,\,\times\prod_{i=e}^{e+l-1} \left(\begin{array}{c} 1 \\ -b_{n-1,i}\end{array}\right) \\
&\,\,\,\,\,\,\,\,\,\,\,\times {F}\left[ \left( \begin{array}{c} -1 \\ 0\end{array} \right)^{s_2}\left( \begin{array}{c} -1 \\ r\end{array} \right)^{b_{n-1,e+l}-s_2-1} \left( \begin{array}{c} -1\\ r -1\end{array} \right) \right] \prod_{i=e+l+1}^{c_{n-1}}{F} \left[ {\left( \begin{array}{c} -1 \\ r\end{array} \right)^{b_{n-1,i}-1} \left( \begin{array}{c} -1 \\r -1\end{array} \right)} \right]F\left(\begin{array}{c} -1\\ \text{ } \end{array} \right)
\endaligned$$
$$\aligned
&= {{F}\left[ \left(\begin{array}{ccc} \text{ }&1& -1\\ 0& 1& r-1 \end{array}\right)\left(\begin{array}{c} -1\\ r \end{array}\right)^{r-2}\left(\begin{array}{c}-1 \\ r-1 \end{array}\right) \right] }\prod_{i=2}^{e-2}{F} \left[{\left( \begin{array}{c} -1 \\ r \end{array} \right)^{b_{n-1,i}-1} \left( \begin{array}{c} -1 \\ r -1\end{array} \right)}\right]\\
&\,\,\,\,\,\,\,\,\,\,\,\times{F} \left[\left(\begin{array}{c} -1 \\ r \end{array}\right)^{b_{n-1,e-1}-s_1} \begin{array}{c} \underbrace{\left( \begin{array}{cccc} -1 & \cdots & -1 & -1 \\ 0 &\cdots& 0 & -1\end{array} \right)}\\ s_1\text{ columns}\end{array} \right] \\
&\,\,\,\,\,\,\,\,\,\,\,\times\prod_{i=e}^{e+l-1} \left[ \left( \begin{array}{c}0 \\ 0 \end{array} \right)^{r-2} \left( \begin{array}{c}1 \\ -1 \end{array} \right)
\left(\left( \begin{array}{c}0 \\ 0 \end{array} \right)^{r-1}\left( \begin{array}{c}0 \\ -1 \end{array} \right) \right)^{b_{n-1,i}-1}\right] \\
&\,\,\,\,\,\,\,\,\,\,\,\times {F}\left[ \left( \begin{array}{c} -1 \\ 0\end{array} \right)^{s_2}\left( \begin{array}{c} -1 \\ r\end{array} \right)^{b_{n-1,e+l}-s_2-1} \left( \begin{array}{c} -1\\ r -1\end{array} \right)\right] \prod_{i=e+l+1}^{c_{n-1}}{F} \left[ {\left( \begin{array}{c} -1 \\ r\end{array} \right)^{b_{n-1,i}-1} \left( \begin{array}{c} -1 \\r -1\end{array} \right)} \right]F\left(\begin{array}{c} -1\\ \text{ } \end{array} \right)
\endaligned$$
$$\aligned
&= {{F}\left[ \left(\begin{array}{ccc} \text{ }&1& -1\\ 0& 1& r-1 \end{array}\right)\left(\begin{array}{c} -1\\ r \end{array}\right)^{r-2}\left(\begin{array}{c}-1 \\ r-1 \end{array}\right) \right] }\prod_{i=2}^{e-2}{F} \left[{\left( \begin{array}{c} -1 \\ r \end{array} \right)^{b_{n-1,i}-1} \left( \begin{array}{c} -1 \\ r -1\end{array} \right)}\right]\\
&\,\,\,\,\,\,\,\,\,\,\,\times{F} \left[\left(\begin{array}{c} -1 \\ r \end{array}\right)^{b_{n-1,e-1}-s_1} \begin{array}{c} \underbrace{\left( \begin{array}{cccc} -1 & \cdots & -1 & -1 \\ 0 &\cdots& 0 & -1\end{array} \right)}\\ s_1\text{ columns}\end{array} \right] \\
&\,\,\,\,\,\,\,\,\,\,\,\times\prod_{i=e}^{e+l-1}\left[ \tilde{F}\left( \begin{array}{c}-1 \\ -1 \end{array} \right)
\left(\tilde{F}\left( \begin{array}{c}-1 \\ 0 \end{array} \right) \right)^{b_{n-1,i}-1} \right] \\
&\,\,\,\,\,\,\,\,\,\,\,\times {F}\left[ \left( \begin{array}{c} -1 \\ 0\end{array} \right)^{s_2}\left( \begin{array}{c} -1 \\ r\end{array} \right)^{b_{n-1,e+l}-s_2-1} \left( \begin{array}{c} -1\\ r -1\end{array} \right)\right] \prod_{i=e+l+1}^{c_{n-1}}{F} \left[ {\left( \begin{array}{c} -1 \\ r\end{array} \right)^{b_{n-1,i}-1} \left( \begin{array}{c} -1 \\r -1\end{array} \right)} \right]F\left(\begin{array}{c} -1\\ \text{ } \end{array} \right) \\
&=\tilde{F}({F}(z_{n-1})|_{W'}).
\endaligned$$
Even if $s_1=0$, we use the same argument with $\left(\begin{array}{c} -1 \\ r \end{array}\right)^{b_{n-1,e-1}-s_1}$ being replaced by $\left(\begin{array}{c} -1 \\ r \end{array}\right)^{b_{n-1,e-1}-1}\left(\begin{array}{c} -1 \\ r-1 \end{array}\right)$. Remark~\ref{Excrem} explains why the same proof works for the case of $s_1=0$.
\end{proof}
\begin{exmp}\label{exmp3}Let $r=3$.
Then $z_2=\left(\begin{array}{cccc} &1 & -1 & -1\\ 0&1&2 & \end{array} \right)$.
So we have
$$F(z_2)|_{\emptyset}=\left(\begin{array}{cccccc}& 1 & -1 & -1& -1& -1\\ 0&1&2 & 3 &2 & \end{array} \right)\text{ and }F(z_2)|_{\{1\}}=\left(\begin{array}{cccccc} &1 & -1 & -1& -1& -1\\ 0&1&-1 & 3 &2 & \end{array} \right).$$
$$\aligned& \tiny{\tilde{F}(F(z_2)|_{\emptyset})=\sum_{\delta_1=0}^1\cdots\sum_{\delta_8=0}^1\left(\begin{array}{ccccccccccc} & 1 & -1 & -1& -1& -1& -1& -1& -1& -1&-1\\ 0&1&2-3\delta_1 & 3-3\delta_2 &2-3\delta_3 & 3-3\delta_4&3-3\delta_5 & 2-3\delta_6&3-3\delta_7 & 2-3\delta_8&\end{array} \right).}
\endaligned$$
$$\aligned&\tiny{ \tilde{F}(F(z_2)|_{\{1\}})=\sum_{\delta_4=0}^1\cdots\sum_{\delta_8=0}^1\left(\begin{array}{cccccccccccc} & 1 & -1 &0&1& -1& -1& -1& -1& -1& -1\\ 0&1&-1 &0&-1 & 3-3\delta_4&3-3\delta_5 & 2-3\delta_6&3-3\delta_7 & 2-3\delta_8\end{array} \right).}
\endaligned$$
$$\aligned& \tilde{F}(F(z_2)|_{\emptyset})|_{\{1,2,3\}}=\left(\begin{array}{ccccccccccc} & 1 & -1 & -1& -1& -1& -1& -1& -1& -1&-1\\0& 1&-1 & 0 &-1 & 3&3 & 2&3 & 2&\end{array} \right).
\endaligned$$
$$\aligned& \tilde{F}(F(z_2)|_{\{1\}})|_{\emptyset}=\left(\begin{array}{cccccccccccc} &1 & -1 &0&1& -1& -1& -1& -1& -1& -1\\ 0&1&-1 &0&-1 & 3&3 & 2&3& 2\end{array} \right).
\endaligned$$
Then $$\aligned &F\left(\tilde{F}(F(z_2)|_{\emptyset})|_{\{1,2,3\}\setminus f(\emptyset)} + \tilde{F}(F(z_2)|_{\{1\}})|_{\{1,2,3\}\setminus f(\{1\})} \right)|_{\emptyset}\\
&=F\left(\tilde{F}(F(z_2)|_{\emptyset})|_{\{1,2,3\}} + \tilde{F}(F(z_2)|_{\{1\}})|_{\emptyset} \right)|_{\emptyset}\\
&=\tiny{\left(\begin{array}{cccccccccc} &1 & -1&0& 1 &0&0 &0&0 &1 \\ 0&1&-1&0 &-1 &0&0&-1 &0 &-1 \end{array} \right)\left(\begin{array}{ccc} -1 &-1 &-1 \\ 3 &3 &2 \end{array} \right)^2\left(\begin{array}{cccccccc} -1 &-1&-1&-1&-1&-1&-1&-1 \\ 3 &2&3 &3 &2 &3 &2 &\end{array} \right)}\\
&=\tilde{F}(F(z_3)|_{\{1,2,3\}})|_{\emptyset}
\endaligned$$
\end{exmp}
\section{Rank 2 Cluster Algebra of affine type}\label{affine}
In this section, we would like to compare our formula with the known formula by Caldero and Zelevinsky \cite{CZ} among others \cite{SZ}, \cite{MP}. Fix $r=2$.
\begin{defn}
Let $\{c_n\}$ be the sequence defined by the recurrence relation $$c_n=2c_{n-1} -c_{n-2},$$ with the initial condition $c_1=0$, $c_2=1$. Then $c_n= n-1.$ \qed
\end{defn}
If we have a map $G : (x, y) \mapsto (xyx^{-1}, y^2x^{-1})$ and define $(z_n, w_n):=G^n(x, y)$, then it is easy to see that $z_n$ is one of the terms in $x_n$. Then we have
$$z_n=\begin{array}{c}\underbrace{\left(\begin{array}{ccccccc} \text{ }&1 & -1 & -1 & \cdots & -1& -1\\ 0&1& 1 & 1 & \cdots & 1 & \text{ }\end{array} \right)}\\ n+2\text{ columns}\end{array}.$$
When $r=2$, no exceptional string occurs, so we do not need $\text{Exc}$. When $r=2$, definition of $f$ can be reduced as follows.
\begin{defn}\label{def_of_f2}
We need a function $f$ from $\{\text{subsets of } [c_{n-1}]\}$ to $\{\text{subsets of } [c_{n}]\}$. For each subset $V\subset [c_{n-1}]$, we define $f(V)$ as follows.
If $V=\emptyset$ then $f(\emptyset)=\emptyset$. If $V\neq\emptyset$ then we write $V=\cup_{i=1}^j\{e_i,e_i+1,\cdots, e_i+l_i-1\}$ with $l_i>0$ $(1\leq i\leq j)$ and $e_i+l_i<e_{i+1}$ $(1\leq i\leq j-1)$. Then
$$
f(V):=\cup_{i=1}^j\{e_i,e_i+1,\cdots, e_i+l_i\}
$$
\qed\end{defn}
\begin{defn}\label{restrict_to_term2}
Let $$T=\sum_{\delta_1\in H_1}\cdots\sum_{\delta_{c_n}\in H_{c_n}} \left(\begin{array}{ccccccc} \text{ }&1 & \alpha_1& \alpha_2 & \cdots & \alpha_{c_n}& -1\\ 0&1& \beta_{1}-2\delta_{1} & \beta_{2}-2\delta_{2} & \cdots & \beta_{c_{n}}-2\delta_{c_n} & \text{ }\end{array} \right),$$ where each of $H_i$ is either $\{0\}$ or $\{0,1\}$. Let $V$ be any subset of $[c_n]$. Then we define
$$T|_V:=\left\{\begin{array}{ll} 0, & \text{ if } H_i=\{0\} \text{ for some } i\in V \\
\text{ } & \text{ }\\
\begin{array}{l}\left(\begin{array}{ccccccc} \text{ }&1 & \alpha_1 & \alpha_2 & \cdots & \alpha_{c_n}& -1\\ 0&1& \beta_{1}-2\delta_{1} & \beta_{2}-2\delta_{2} & \cdots & \beta_{c_{n}}-2\delta_{c_n} & \text{ }\end{array} \right) \\\text{ where }\delta_{i}=1\text{ for }i\in V\text{ and }\delta_{i}=0\text{ for }i\not\in V, \end{array}
& \text{ otherwise.}\end{array}\right.$$
Note that $T=\sum_{V\subset [c_n]} T|_V$.
\qed\end{defn}
It is easy to see that
$$F(z_{n-2})=\sum_{\delta_1=0}^1\cdots\sum_{\delta_{c_{n-1}}=0}^1\left(\begin{array}{ccccccc} \text{ }&1 & -1 & -1 & \cdots & -1& -1\\ 0&1& 1-2\delta_1 & 1-2\delta_2 & \cdots & 1-2\delta_{c_{n-1}} & \text{ }\end{array} \right).$$
\begin{defn}\label{def_of_tildeF2}
Let $$w={F}(z_{n-1})|_V= \left(\begin{array}{ccccccc} \text{ }&1 & -1 & -1 & \cdots & -1& -1\\ 0&1&1-2\delta_{1} & 1-2\delta_{2} & \cdots & 1-2\delta_{c_n} & \text{ }\end{array} \right),$$
where $\delta_{i}=1$ for $i\in V$ and $\delta_{i}=0$ for $i\not\in V$. We will define $\tilde{F}(w)$, which is a modification of taking $F$, where we do not allow $(1+y^2)^{-1}$ to appear. In fact, if none of $1-2\delta_{i}$ is negative, then $\tilde{F}(w)=F(w)$. In general, $\tilde{F}(w)$ can be defined by letting
$$\tilde{F}(w)=\tilde{F}\left(\begin{array}{ccc} \text{ }&1 & -1 \\ 0&1& 1-2\delta_{1} \end{array} \right)\left[\prod_{i=2}^{c_n} \tilde{F}\left(\begin{array}{c} -1\\1-2\delta_{i} \end{array} \right)\right]\left(\begin{array}{c} -1\\ \text{ } \end{array} \right)$$
and using the following rules:
$$\tilde{F}\left(\begin{array}{ccc} \text{ }&1 & -1\\ 0&1 & 1-2\delta_1 \end{array} \right):=\sum_{\delta^{(1)}=0}^{1-\delta_{1}}\sum_{\delta^{(2)}=0}^{\Delta_{V,1}(1-\delta_{1})}\left(\begin{array}{cccc} \text{ }&1& -1& 2\delta_1-1\\ 0& 1& 2(1-\delta^{(1)})-1 & 2\Delta_{V,1}(1-\delta^{(2)})-1 \end{array}\right),$$
$$
\tilde{F}\left(\begin{array}{c} -1\\1-\delta_{i}r \end{array} \right):= \sum_{\delta=0}^{\Delta_{V,i}(1-\delta_{i})}\left( \begin{array}{c} 2\delta_i-1 \\ 2\Delta_{V,i}(1-\delta) -1\end{array} \right) \,\,\,\,\,\,\,\,\, \text{ for }1<i\leq c_n,
$$
where
$\Delta_{V,i}=\left\{\begin{array}{ll} 0, &\text{ if }\delta_i=0 \text{ and }\delta_{i+1}=1\\
1, & \text{ otherwise}. \end{array} \right.$
\end{defn}
\begin{thm}\label{mainthm2}
We have
\begin{equation}\label{maineq2}\aligned x_n&=\sum_{V\subset [n-2]} \left(\sum_{W\subset [n-1]\setminus f(V)} \tilde{F}({F}(z_{n-2})|_V)|_W\right)=\sum_{V\subset [n-2]} \left( \tilde{F}({F}(z_{n-2})|_V)\right)\\&=\sum_{\alpha_i,\beta_i} \left(\begin{array}{ccccccc} \text{ }&1 & \alpha_1 & \alpha_2 & \cdots & \alpha_{n-1}& -1\\ 0&1& \beta_{1} & \beta_{2} & \cdots & \beta_{n-1} & \text{ }\end{array} \right), \endaligned \end{equation}
where the summation runs over $\alpha_i,\beta_i$ satisfying the following:\\
$$\aligned \alpha_i=1\text{ or }&-1, \,\, \beta_i=1\text{ or }-1,\\&\alpha_1=-1,\\ \text{ if }\alpha_i=1&\text{ then }\beta_i=-1,\\ \text{ and if }\alpha_i=-1\text{ and }&\alpha_{i+1}=1\text{ then }\beta_i=-1. \endaligned$$
\end{thm}
Let us compare our formula with the known formula by Caldero and Zelevinsky \cite{CZ}. Let $j(U)$ denote the number of connected components of a subset $U\subset [n-1]\setminus\{1\}$, i.e. $U=\cup_{i=1}^{j(U)}\{e_i,e_i+1,\cdots, e_i+l_i-1\}$ with $l_i>0$ $(1\leq i\leq j(U))$ and $e_i+l_i<e_{i+1}$ $(1\leq i\leq j(U)-1)$. In \cite[Proposition 4.3]{CZ}, it is shown that the number of $p$-element subsets $U\subset [n-1]\setminus\{1\}$ with $j(U)=t$ is equal to ${{p-1}\choose{t-1}}{{n-1-p}\choose{t}}$. For each $U\subset [n-1]\setminus\{1\}$, if we let $\alpha_i=1$ for $i\in U$ and $\alpha_i=-1$ for $i\not\in U$, then $|U|(=p)$ determines the degree of $x$ in the commutative case. The degree of $x$ is equal to $2p-n+1$. Let $t=j(U)$. If $U=\emptyset$ then $\beta_i=1$ or $-1$ for any $i\in[n-1]$. If $U\neq\emptyset$ then only for $i\in [n-1]\setminus (U\cup \{e_i-1\}_{1\leq i\leq t})$, we have choices, namely $\beta_i=1$ or $-1$. If $|\{i\,|\,\beta_i=1\}|=q$ then the degree of $y$ is equal to $2q-n+2$. Since $|[n-1]\setminus (U\cup \{e_i-1\}_{1\leq i\leq t})|= n-1-p-t$ for $U\neq\emptyset$, the specialization of (\ref{maineq2}) at $xy=yx$ is equal to
$$\tiny{\aligned
&\sum_q {{n-1}\choose q}x^{-n+1}y^{2q-n+2} + \sum_{p\geq 1} \sum_{q,t} {{n-1-p-t}\choose {q}}{{p-1}\choose{t-1}}{{n-1-p}\choose{t}}x^{2p-n+1}y^{2q-n+2}\\
&=\sum_q {{n-1}\choose q}x^{-n+1}y^{2q-n+2} +\sum_{p\geq 1} \sum_{q,t}{{n-1-p-t}\choose {q}}{{n-1-p}\choose{n-1-p-t}}{{p-1}\choose{t-1}}x^{2p-n+1}y^{2q-n+2}\\
&=\sum_q {{n-1}\choose q}x^{-n+1}y^{2q-n+2} +\sum_{p\geq 1} \sum_{q,t}{{n-1-p-q}\choose {n-1-p-t-q}}{{n-1-p}\choose{q}}{{p-1}\choose{t-1}}x^{2p-n+1}y^{2q-n+2}\\
&=\sum_q {{n-1}\choose q}x^{-n+1}y^{2q-n+2} +\sum_{p\geq 1} \sum_{q}{{n-1-p}\choose{q}}\sum_t {{n-1-p-q}\choose {n-1-p-t-q}}{{p-1}\choose{t-1}}x^{2p-n+1}y^{2q-n+2}\\
&=\sum_q {{n-1}\choose q}x^{-n+1}y^{2q-n+2} +\sum_{p\geq 1} \sum_{q}{{n-1-p}\choose{q}} {{n-2-q}\choose {n-2-p-q}}x^{2p-n+1}y^{2q-n+2}\\
&=\sum_q {{n-1}\choose q}x^{-n+1}y^{2q-n+2} +\sum_{p\geq 1} \sum_{q} {{n-2-q}\choose {p}}{{n-1-p}\choose{q}}x^{2p-n+1}y^{2q-n+2},
\endaligned}$$
which is precisely the same as \cite[Theorem 4.1 (4.3)]{CZ}.
|
\section{Introduction}
A semi-classical theory considers the back reaction of quantum
fields in a classical geometric background. This approach started
a long time ago with de Witt \cite{DeWitt}, and since then, their
consequences and applications are still under investigation (see,
e.g., Ref. \cite{Hu}).
Differently from the usual Einstein-Hilbert action, the predicted
gravitational action allows differential equations with fourth
order derivatives, which is called the full theory \cite{DeWitt},
see also \cite{liv}.
The full higher order theory was previously studied by Starobinsky
\cite{S}, and more recently, also by Shapiro, Pelinson and others
\cite{Shapiro}. In \cite{Shapiro} only the
homogeneous and isotropic spacetime is studied.
Here, the full theory with four time derivatives is addressed,
which was first investigated for general Bianchi I spacetimes in
Tomita's article \cite{berkin}. They found that the presence of an
anisotropy contributes to the formation of a singularity.
Schmidt has a recent and very interesting review on higher order
gravity theories in connection to cosmology \cite{hjs}.
For general anisotropic Bianchi I homogeneous spacetimes, the full
theory reduces to a system of nonlinear ordinary differential
equations. The numerical solutions for this system with
derivatives of fourth order in time was previously obtained by one
of us and a collaborator \cite{sandro}.
The question of stability of Minkowski space in this quadratic
theory of gravity was also addressed and we have obtained that there
are many initial conditions that result in Minkowski space
asymptotically \cite{daniel_sandro}. The analysis is restricted to
Bianchi I type spacetimes. There is a basin of attraction to the
Minkowski spacetime. Isotropisation can also occur towards de
Sitter spacetime for a non zero cosmological constant. In this
sense the above theory is structurally stable, since for a set of
non zero measure of initial anisotropic geometries, isotropisation
occurs.
It should be mentioned that the isotropisation process for non
zero cosmological constant is not a peculiarity of these higher
derivatives theories. It also occurs in the Einstein's theory of General
Relativity, as proved in a very interesting article
\cite{wald}, where for $\Lambda>0$ all Bianchi models, except a
highly positively curved Bianchi IX, the theory becomes asymptotically a de
Sitter spacetime. The remarkable difference between the theory of General
Relativity and the higher derivatives theories is that the isotropisation depends on the initial conditions. The occurrence of isotropisation is strongly dependent on the initial conditions for quadratic curvature theories such as those studied here.
In the present work our aim is to investigate how isotropisation is approached, and if this very process could naturally generate gravitational waves. For this reason an initial condition is chosen in such a way that it asymptotically generates the Minkowski spacetime, and the Weyl's scalars are also obtained for this solution. It must be emphasized that this solution, although numeric, is an exact solution of this particular quadratic gravity in the sense that it depends only on the precision of the machine. We then consider a perturbative approach to investigate if gravitational waves are in fact generated during the isotropisation process.
The article is organized as follows. In section II a brief discussion on the perturbative approach is given. The field tensor is obtained to first order in the perturbation parameter, resulting in the linearized field equation. In section III we evaluate the Weyl invariants $\Psi_2$ and $\Psi_4$ and check their time dependence, albeit having a decreasing behavior they are non null which indicate the presence of gravitational waves. Then, two distinct methods are applied to extract the GW amplitudes from the exact numerical solution. It is also shown that the metric perturbation can be understood in the TT gauge as a superposition of the Minkowski space, a pure tensorial perturbation and a pure scalar perturbation. Our conclusions and final remarks are presented in section IV.
The following conventions and units are taken
$R_{bcd}^{a}=\Gamma_{bd,c}^{a}-...$, $R_{ab}=R_{acb}^{c}$,
$R=R_{a}^{a}$, metric signature $-+++$, Latin symbols run from
$0-3$, Greek symbols run from $1-3$ and $G=\hbar=c=1$.
\section{Perturbative Analysis}
\subsection{The full equations}
The Lagrangian density we consider here is
\begin{equation}\label{acao}
{\cal L} = \sqrt{-g}\left[\Lambda+R+\alpha\left(R^{ab}R_{ab}
-\frac{1}{3}R^{2}\right)+\beta R^{2}\right]+{\cal L}_{c}\,.
\end{equation}
By varying the action $I = \int {\cal L} d^4x$ with respect to the metric we find the field equations
\begin{equation}\label{tensor E}
E_{ab}=G_{ab}+\left(\beta-\frac{1}{3}\alpha\right)H_{ab}^{(1)}+\alpha
H_{ab}^{(2)}-T_{ab}-\frac{1}{2}\Lambda g_{ab},
\end{equation}
where
\begin{eqnarray}
& & H_{ab}^{(1)}=\frac{1}{2}g_{ab}R^{2}-2RR_{ab}+2R_{;ab}-2\square Rg_{ab},\label{tensor H1}\\
& & H_{ab}^{(2)}=\frac{1}{2}g_{ab}R_{mn}R^{mn}+R_{;ab}-2R^{cn}R_{cbna}-\square R_{ab}-\frac{1}{2}\square Rg_{ab},\label{tensor H2}\\
& & G_{ab}=R_{ab}-\frac{1}{2}g_{ab}R,\label{tensor G}
\end{eqnarray}
and $T_{ab}$ is the energy momentum source, which comes from the classical part of the Lagrangian ${\cal L}_{c}$. Only vacuum solutions $T_{ab}={\cal L}_{c}=0$ will be considered in this article since it seems the most natural condition soon after the Planck era. The covariant divergence of the above tensors are identically zero due to their variational definition. The following Bianchi Type I line element is considered
\begin{equation}
ds^{2}=-dt^{2}+[a_{1}(t)]^{2}dx^{2}
+[a_{2}(t)]^{2}dy^{2}+[a_{3}(t)]^{2}dz^{2},\label{elinha}
\end{equation}
which is a general spatially flat and anisotropic space with
proper time $t$. With this line element all the tensors which
enter the expressions are diagonal. The substitution of Eq. \eqref{elinha} in Eq. \eqref{tensor E} with $T_{ab}=0$ results, for
the spatial part of \eqref{tensor E}, in differential equations of
the type
\begin{eqnarray}
& & \frac{d^{4}}{dt^{4}}a_{1}=f_{1}\left(\frac{d^{3}}{dt^{3}}a_{i},\ddot{a}_{i},\dot{a}_{i},a_{i}\right)\label{edo1}\\
& & \frac{d^{4}}{dt^{4}}a_{2}=f_{2}\left(\frac{d^{3}}{dt^{3}}a_{i},\ddot{a}_{i},\dot{a}_{i},a_{i}\right)\label{edo2}\\
& & \frac{d^{4}}{dt^{4}}a_{3}=f_{3}\left(\frac{d^{3}}{dt^{3}}a_{i},\ddot{a}_{i},\dot{a}_{i},a_{i}\right),\label{edo3}
\end{eqnarray}
where the functions $f_{i}$ involve the $a_{1},\; a_{2},\; a_{3}$ and their derivatives in a polynomial fashion (see the Appendix of the Ref. \cite{daniel_sandro}). The Ref. \cite{Noakes} shows that the theory which follows from Eq. \eqref{acao} has a well posed
initial value problem. This issue has some similarities with
General Relativity theory since it is necessary to solve the
problem of boundary conditions and initial conditions
simultaneously. According to Noakes \cite{Noakes}, once a
consistent boundary condition is chosen initially, the time
evolution of the system is uniquely specified. In \cite{Noakes}
the differential equations for the metric are written in a form
suitable for the application of the theorem of Leray \cite{Leray}.
For homogeneous spaces the differential equations derived from Eq. \eqref{tensor E} reduce to a set of non linear ordinary differential equations. Then, instead of going through the general construction
given in Ref. \cite{Noakes}, in this particular case, the existence and
uniqueness of the solutions of Eqs. \eqref{edo1}-\eqref{edo3} reduce to
the well known problem of existence and uniqueness of solutions of
ordinary differential equations. For a proof on local existence
and uniqueness of solutions of differential equations see, for
instance \cite{reedsimon}.
Besides the Eqs. \eqref{edo1}-\eqref{edo3}, we have the
temporal part of Eq. \eqref{tensor E}. To understand the role of this
equation we have first to study the covariant divergence of the
Eq. \eqref{tensor E},
\begin{equation}
\nabla_{a}E^{ab}=\partial_{a}E^{ab}+\Gamma_{ac}^{a}E^{cb}+\Gamma_{ac}^{b}E^{ac}=0.
\end{equation}
Recall that the coordinates that we are using are those specified by the metric \eqref{elinha} and that $T_{ab}=0$. Since the differential equations
\eqref{edo1}-\eqref{edo3} are solved numerically $E^{\alpha\alpha}\equiv0,$
\begin{eqnarray}
& & \partial_{0}E^{00}+\Gamma_{a0}^{a}E^{00}+\Gamma_{00}^{0}E^{00}=0, \\
& & E^{00}(t)\equiv0,\label{vinculo}
\end{eqnarray}
where $E_{00}$ is the $00$ component of Eq. \eqref{tensor E}. If $E_{00}=0$
initially, it will remain zero at any instant. Therefore the
equation $E_{00}$ acts as a constraint on the initial conditions
and we use it to test the accuracy of our results. The full numerical solution to this problem was worked out in the Ref. \cite{daniel_sandro}.
\subsection{Perturbative Approximation}
Following MTW \cite{MTW} we consider metric perturbations in the form
\begin{eqnarray}
& & g_{ab}=g_{ab}^{(B)}+\varepsilon h_{ab}\\
& & g^{ab}=g^{ab}_{(B)}-\varepsilon h^{ab}+\varepsilon^{2}h^{ac}h_{c}^{\; b}+...
\end{eqnarray}
In a free falling frame we have $g_{ab}^{(B)}=\eta_{ab}$ for the
background metric and locally ${\Gamma^{(B)}}^a_{bc}=0$ so that
the connection compatible with $g_{ab}$ is identical to
$\Gamma_{ab}^{c}=\varepsilon S_{\; ab}^{c}$ and then
\begin{eqnarray}
& & \nabla_{a}T_{b}=T_{a;b}=T_{a|b}-\varepsilon S_{\; ab}^{c}T_{c} \\
& & S_{ab}^{c}=\frac{1}{2}g_{(B)}^{cs}\left(h_{sa|b}+h_{sb|a}-h_{ab|s}\right),\label{conexao}
\end{eqnarray}
where $|$ is the covariant derivative with respect to the
background metric $g_{ab}^{(B)}$ and $\nabla_{a}=;a$ is the
covariant derivative with respect to $g_{ab}$. The tensors which
appear in Eqs. \eqref{tensor E}, \eqref{tensor H1}-\eqref{tensor G} can
be expanded in powers of $\varepsilon$ to first order, namely
\begin{eqnarray}
& & G_{ab}^{1}=R_{ab}^{1}-\frac{1}{2}h_{ab}R^{(B)}-\frac{1}{2}g_{ab}^{(B)}R^{1}\\
& & H_{ab}^{(1)1}=2R_{|ab}^{1}-2\square^{(B)}R^{1}g_{ab}^{(B)}\\
& & H_{ab}^{(2)1}=R_{|ab}^{1}-\square^{(B)}R_{ab}^{1}
-\frac{1}{2}\square^{(B)}R^{1}g_{ab}^{(B)}.
\end{eqnarray}
where
$R^1$, $R^1_{ab}$, are the Ricci scalar and the Ricci tensor to first order in
$\varepsilon$, $R_{abcd}^{1}$ is the Riemann tensor to first order (see \cite{MTW}) and $\square^{B}T^{a}=g^{Bmn}\nabla_{m}^{B}\nabla_{n}^{B}T^{a}=T_{\,|c}^{a\,\,|c}$.
These expressions are combined to give a perturbative sum for
$E_{ab}$, to first order \begin{eqnarray}
& & E_{ab}=E_{ab}^{(B)}+\varepsilon E_{ab}^{1}+..\nonumber \\
& & E_{ab}^{1}=G_{ab}^{1}+\left(\beta-\frac{1}{3}\alpha\right)H_{ab}^{(1)1}+\alpha H_{ab}^{(2)1}-\frac{1}{2}\Lambda h_{ab}\label{tensor E1}
\end{eqnarray}
where $E_{ab}^{(B)}$ correspond to the background equations which
is supposed to be exactly $E_{ab}^{(B)}\equiv0.$ By imposing
$E_{ab}^{1}=0$ it is possible to obtain linearized partial
differential equations for $h_{ab}.$
\section{Numerical Results}
In what follows we apply two distinct methods in
order to obtain a GW solution as a final result of isotropisation.
Our initial ansatz is that after some arbitrary time interval we
can consider the full solution as asymptotically Minkowskian. With
this is mind, it is possible to split the asymptotic solution in a
flat background metric plus a perturbative quantity. Thus, in the first
approach we numerically extract the metric perturbation from the
full solution, while in the second approach we explicitly write
down the governing equations for the perturbations.
\subsection{Decomposition of the asymptotic solution}
First consider the full numerical solution of the Eqs.
\eqref{edo1}, \eqref{edo2} and \eqref{edo3}. In FIG. \ref{fig 1}
we show a particular solution for which isotropisation occurs,
the parameters chosen are $\alpha = 2$, $\beta = -1$, $\Lambda =
0$ with initial conditions
$\mbox{\ensuremath{\dot{a}}}_{1}=0.1,\,\dot{a}_{2}=0.3,\,
a_{1}=1,\, a_{2}=1,\, a_{3}=1$; $\dot{a}_{3}$ is chosen in
accordance with Eq. \eqref{vinculo}, and all the other higher order
derivatives are chosen to be zero.
\begin{figure}
\resizebox{0.85\columnwidth}{!}{\includegraphics{fig1.eps}}
\caption{The numerical solution of Eqs. \eqref{edo1}\eqref{edo2}\eqref{edo3}
for $a_{1}(t),$ $a_{2}(t)$ and $a_{3}(t)$ together with a linear
regression. The parameters are $\alpha=2.0,\,\beta=-1.0,\,\Lambda=0$
and the initial condition is $\mbox{\ensuremath{\dot{a}}}_{1}=0.1,\,\dot{a}_{2}=0.3,\, a_{1}=1,\, a_{2}=1,\, a_{3}=1$.
The numerical value of $\dot{a}_{3}$ is chosen in accordance with
\eqref{vinculo}, and all the other higher derivatives are chosen
to be zero. The linear regression is done for a time interval around $t\simeq8\times10^{7}$ in natural units,
and the resulting values for the coefficients are, see \eqref{regresslinear}
$b_1=0.0002293455,\,b_2=0.0007793157,\,b_3=0.0005822617,\,c_1=9173.663,\,c_2=31171.99,\,
c_3=23289.99$. \label{fig 1}}
\end{figure}
We can check the existence of gravitational waves in that solution by evaluating the resulting $\Psi_4$ Weyl's invariant that has spin 2 (see, e.g., \cite{Stephani1990}). The time dependency of $\Psi_4$ is displayed in FIG. \ref{Psi_4}, where, for completeness, we are also showing the $\Psi_2$ invariant that has spin 0. They were evaluated for the initial condition and parameters considered above and, as can be seen, both of them decay with time albeit being non zero. For the particular line element chosen here we have $\Psi_0\equiv \Psi_4$ and $\Psi_1\equiv \Psi_3\equiv 0$.
\begin{figure}
\begin{center}
\resizebox{0.85\columnwidth}{!}{\includegraphics{fig2.eps} }
\end{center}
\caption{Time decaying behaviour of the Weyl's invariants $\Psi_4$ and $\Psi_2$, respectively of spin $2$ and $0$, for the full solution for $a_1(t)$, $a_2(t)$ and $a_3(t)$ shown in FIG. \ref{fig 1}. \label{Psi_4}}
\end{figure}
Since the existence of tensor gravitational waves is evidenced by the presence of a non-null $\Psi_4$ invariant, and a scalar perturbation is also present as shown by the invariant $\Psi_2$, the next step is to find the time evolution of the amplitudes of the metric perturbations.
In a first approach we consider an arbitrary time interval around $t\simeq 8\times 10^7$ in which we have applied a linear regression (see FIG. \ref{fig 1}) and the following asymptotic expressions were obtained
\begin{eqnarray} \label{regresslinear}
&& a_{1}\simeq c_{1}+b_{1}t\nonumber\\
&& a_{2}\simeq c_{2}+b_{2}t\\
&& a_{3}\simeq c_{3}+b_{3}t,\nonumber
\end{eqnarray}
where
$b_1=0.0002293455,\,b_2=0.0007793157,\,b_3=0.0005822617,\,c_1=9173.66,\,c_2=31171.99
$ and $c_3=23289.99$.
After the subtraction of Eq. (\ref{regresslinear}) from the exact numerical solution of the Eqs. \eqref{edo1}, \eqref{edo2} and \eqref{edo3}, it can be seen that the metric of the complete solution can be rearranged in the following way
\begin{equation}
g_{ab} = {\tilde{g}}^{(B)}_{ab} + {\tilde{h}}_{ab},
\end{equation}
where ${\tilde{g}}^{(B)}_{ab}$ is a constant background given by
${\tilde{g}}^{(B)}_{ab} = {\rm diag}(-1,c_1^2, c_2^2, c_3^2)$ and
${\tilde{h}}_{ab}$ is a superposition of the time oscillating
functions $\tilde{h}_1(t)$, $\tilde{h}_2(t)$ and $\tilde{h}_3(t)$
plus a term that gives a quadratic in time contribution
\begin{eqnarray}\label{h tilde}
\tilde{h}_{ab}(t) & = & \left(\begin{array}{cccc}
0 & 0 & 0 & 0\\
0 & \tilde{h}_{1}(t)+2c_{1}b_{1}t+b_{1}^{2}t^{2} & 0 & 0\\
0 & 0 & \tilde{h}_{2}(t)+2c_{2}b_{2}t+b_{2}^{2}t^{2} & 0\\
0 & 0 & 0 & \tilde{h}_{3}(t)+2c_{3}b_{3}t+b_{3}^{2}t^{2}\end{array}\right).
\end{eqnarray}
\begin{figure}
\resizebox{0.85\columnwidth}{!}{\includegraphics{fig3.eps} }
\caption{The oscilating part
$\tilde{h}_{i}(t)=[a_{i}(t)]^{2}-[c_{i}+b_{i}t]^{2}.$
The values of $a_{i}$ are the numerical solutions of Eqs. \eqref{edo1}-\eqref{edo3}.
The values of $c_{i}$ and $b_{i}$ are shown in FIG. \ref{fig 1}. Note that the time origin was changed to $t=8.0001\times 10^7 +t^*$. \label{h(t)}}
\end{figure}
\begin{figure}
\resizebox{0.85\columnwidth}{!}{\includegraphics{fig4.eps}}
\caption{This is the plot of the spatial trace $\tilde{g}^{B\,ij}\tilde{h}_{ij}$.
It can be seen that the trace of $\tilde{h}$ is almost constant
in comparison with the individual amplitudes of the components of $\tilde{h}_{ab}$.
Note that the scalar component is smaller than the tensor component. Here also a new time origin was chosen to be
$t=8.0\times 10^7+ t^+$.\label{traco h}}
\end{figure}
The oscillating behavior of the functions $\tilde{h}_1(t)$,
$\tilde{h}_2(t)$ and $\tilde{h}_3(t)$ are shown in FIG.
\ref{h(t)}. Therefore, we can say that the sum
$\tilde{g}_{ab}^{(B)} + \tilde{h}_{ab}$ is identical to the metric
$g_{ab} = {\rm diag}(-1, a_1(t)^2, a_2(t)^2, a_3(t)^2)$. Now, we
must ask ourselves if the perturbation $\tilde{h}_{ab}$ indeed
represents gravitational waves. We see that $\tilde{h}_{ab}$ is
spatially divergenceless since $\tilde{h}_{ab}$ depends only on
time. Also, its time components are null as can be seen from the
matrix (\ref{h tilde}), but it is not traceless. Thus, in order to
obtain a gravitational wave signal in TT gauge we need to subtract
the trace of $\tilde{h}_{ab}$ generating the new quantity
\begin{equation}
h_{ab} = \tilde{h}_{ab} - \frac{1}{4}\tilde{g}^{(B)}_{ab}\tilde{h},
\end{equation}
which finally has the properties
\begin{equation}
{h^{ab}}_{|b} = h^{a0} = h = 0,
\end{equation}
and therefore we conclude that $h_{ab}$ represents tensor gravitational waves. The trace $\tilde{h} =
\tilde{g}_{(B)}^{ab}\tilde{h}_{ab}$ represents a scalar
perturbation (related to $\Psi_2$) that evolves smoothly with time and is smaller than
the tensor perturbation $h_{ab}$ as can be seen in FIG. \ref{traco
h}.
\subsection{The linearized solution}
Now, let us formally write the linearized system of equations
which is asymptotically consistent with the full numerical
solution. Considering the metric as $g_{ab} = \hat{g}^{(B)}_{ab} +
\hat{h}_{ab}$, the linear system to be solved is obtained by
setting $E_{ab}^{1}\equiv0,$ resulting in
\begin{eqnarray}\label{equacao de movimento linear}
& & \frac{d^{4}}{dt^{4}}\hat{h}_{1}=f_{1j}^{3}\left(t\right)\frac{d^{3}}{dt^{3}}\hat{h}_{j}+f_{1j}^{2}(t)\ddot{\hat{h}}_{j}+f_{1j}^{1}(t)\dot{\hat{h}}_{j}+f_{1j}^{0}(t)\hat{h}_{j}\nonumber \\
& & \frac{d^{4}}{dt^{4}}\hat{h}_{2}=f_{2j}^{3}\left(t\right)\frac{d^{3}}{dt^{3}}\hat{h}_{j}+f_{2j}^{2}(t)\ddot{\hat{h}}_{j}+f_{2j}^{1}(t)\dot{\hat{h}}_{j}+f_{2j}^{0}(t)\hat{h}_{j}\nonumber \\
& & \frac{d^{4}}{dt^{4}}\hat{h}_{3}=f_{3j}^{3}\left(t\right)\frac{d^{3}}{dt^{3}}\hat{h}_{j}+f_{3j}^{2}(t)\ddot{\hat{h}}_{j}+f_{3j}^{1}(t)\dot{\hat{h}}_{j}+f_{3j}^{0}(t)\hat{h}_{j}
\end{eqnarray}
where $\hat{h}$ is defined as the oscillatory part of the metric
perturbation and the coefficients $f_{ij}^{k}(t),$ also depend on
the parameters $\alpha,\,\beta,$ and $\Lambda$ as can be seen in the
Appendix. Now consider an initial condition for the linearized
differential equations for $\hat{h}_{i}(t)$. Once the numerical
values of the $a_{i}$ and their derivatives up to the third order
are initially known, the numerical values of the $\hat{h}_{i}$ and
their derivatives up to the third order can also be obtained. Then
a numerical solution of the above linearized system is specified
uniquely and is shown in FIG. \ref{h_l}. As can be seen, the solution
of the linearized system is exactly equal to the oscillatory
solution found in our preceding analysis, which corroborate our
initial ansatz that the full solution can be treated as being
asymptotically Minkowski.
Again, $\hat{h}_{ab}$ cannot be regarded as tensor GWs since it
contains a scalar component. Thus we extract its trace as we did
in the preceding section and we can say that the remaining
quantity $h_{ab} = \hat{h}_{ab} - g^{(B)}_{ab}\hat{h}/4$
represents the GW perturbations.
\begin{figure}
\resizebox{0.85\columnwidth}{!}{\includegraphics{fig5.eps}}
\caption{The numerical time evolution of $\hat{h}_{i}$ for the linearized
system given in Eqs. \eqref{equacao de movimento linear} for an initial condition
consistent with the one shown in FIG. \ref{h(t)}.
It can be seen that the solution is indistinguishable from FIG. \ref{h(t)}. Again, note that the time origin was changed to $t=8.0001\times 10^7 +t^*$.\label{h_l}}
\end{figure}
\newpage
\section{Conclusions}
In the present article it is considered general anisotropic
Bianchi I homogeneous spacetimes given by the line element given by Eq.
\eqref{elinha}. For this line element, the theory given in Eq.
\eqref{acao}, reduces to a system of ordinary nonlinear
differential equations with four time derivatives. This system is
numerically investigated. The assumption of a homogeneous Universe is
artificial, and still presents a first generalization for its very
primordial stages.
There is a well known conjecture that dissipative
processes can take place in an infinite dynamical system reducing
the number of degrees of freedom to just a few. In some sense,
this argument reduces the artificiality of the assumption of a
homogeneous Universe.
Only the vacuum energy momentum classical source is considered. It
should be valid soon after the Planck era in which vacuum
classical source seems the most natural condition.
We have found that the Weyl's invariants $\Psi_4$ and $\Psi_2$ are non-null for the full numerical solution, although they have a decaying behavior. This is an indication of the existence of gravitational waves due to an initially anisotropic Universe. Then, we have adopted two distinct approaches in order to obtain a description for the linearized gravitational wave tensor $h_{ab}$. The two methods have coincided in the prediction of the oscillatory pattern of the tensor component of gravitational waves. A scalar component is also present, that evolves smoothly in the time interval considered and is smaller than the tensor component. We conclude that the asymptotic geometry of an initially anisotropic spacetime is a superposition of Minkowski space and gravitational waves with tensor and scalar components.
\begin{acknowledgements}
D. M. wishes to thank the Brazilian projects: {\it Nova F\'\i sica no Espa\c co} and INCT-A. JCNA would like to thank CNPq and FAPESP for the financial support
\end{acknowledgements}
|
\section{Introduction}
The dependence of an atomic transition frequency on the properties of the nucleus is a frequently investigated question in laser spectroscopy. Possible effects include the hyperfine interaction, which relies on the coupling of the angular momenta of electrons and the nucleus, and isotope shifts, which depend on the mass of the nucleus and its internal charge distribution.
In atomic potassium, the isotopic composition makes systematic experiments of hyperfine interactions and isotope shifts of atomic transitions very challenging.
Two highly abundant isotopes ($^{39}$K and $^{41}$K) are accompanied by many isotopes with very low natural abundance. However, the knowledge of the relative frequency shifts of at least {\it two pairs} of isotopes are required to determine the size of the two non-trivial contributions to the isotope shift: the electronic correlation factor and the volume shift due to the charge distribution of the nucleus \cite{King1963}. Therefore, spectroscopy of the third-most abundant isotope $^{40}$K ($0.01\%$) is crucial. Aside from the prospect of exploring the fundamental atomic structure of potassium, demand for precise laser spectroscopic properties of $^{40}$K also stems from its use in the physics of degenerate Fermi gases,
where it is the isotope of choice in many experiments
\cite{Demarco1999c,Roati2002,Aubin2005,Kohl2005b,Ospelkaus2006,Rom2006,Klempt2007,Wille2008,Taglieber2008,Tiecke2010,Frohlich2011}.
The $4s \rightarrow 4p$ transition of potassium near 770\,nm has been investigated thoroughly several decades ago for isotopes
from $^{38}$K through $^{47}$K \cite{Bendali1981,Touchard1982}. The results of laser spectroscopic isotope shift measurements have been compared to nuclear
physics experiments regarding the nuclear charge radius as well as to theoretical calculations for the electronic correlations \cite{Martensson1990}. Meanwhile, frequency-comb based measurements of this transition have significantly improved the accuracy for the isotopes
$^{39}$K through $^{41}$K \cite{Falke2006}.
The {\it blue} $4s \rightarrow 5p$ transition of potassium near $\lambda=405$\,nm is considerably weaker than the $4s \rightarrow 4p$ transition, which
makes experiments on isotopes with small natural abundance even more difficult. Consequently, investigations of this transition are scarce.
However, recently interest in this transition has renewed because advances in semiconductor laser technology make the wavelength easily accessible \cite{Uetake2003}.
Measurements of the isotope shifts of the $4s \rightarrow 5p$ transition have been undertaken in $^{39}$K and $^{41}$K using Doppler-free saturation spectroscopy
\cite{Halloran2009} and in $^{39}$K, $^{40}$K, and $^{41}$K using resonance ionization mass spectrometry (RIMS) \cite{Iwata2010}.
While the sensitivity of the RIMS method is good to detect the low-abundant isotope $^{40}$K, its spectral resolution is limited.
In particular, the hyperfine splitting of the excited state $5s\;^2P_{1/2}$ was not observed for any of the isotopes.
\section{Experimental setup}
We perform Doppler-free saturation spectroscopy on a potassium vapor cell containing isotopically enriched $^{40}$K ($4\%$). Our experimental setup is sketched in Figure \ref{setup}. The cell (length 100\,mm) is heated to approximately $120^\circ$\,C to provide sufficient absorption on the $4s\rightarrow 5p$ transition. The laser source is a diode laser (Mitsubishi ML320G2-11) operating at 405\,nm in an external cavity setup \cite{Ricci1995}. The laser frequency is swept with a frequency of 11\,Hz by angle-tuning the position of the diffraction grating in the laser setup using a piezo transducer.
\begin{figure}[htbp]
\includegraphics[width=0.9\columnwidth,clip=true]{setup.eps}
\caption{Schematic of the Doppler-free saturation spectroscopy setup.}
\label{setup}
\end{figure}
The laser beam is sent through the vapor cell and is retro-reflected from a mirror. After traveling again through the cell, the beam is reflected by a beam splitter to a photodiode. The photodiode signal is amplified and high-pass filtered with a bandwidth of 1\,kHz. We use derivatives of Lorentzian functions to fit the measured data and to determine the zero crossing of each line. Afterwards we use the well known ground state splitting of $^{41}$K of 254.0\,MHz and $^{40}$K of 1285.8\,MHz \cite{Arimondo1977} as normalization of the frequency axis and for linearizing the piezo scan.
The power of the laser when entering the spectroscopy cell is approximately 8\,mW in a beam of 5\,mm diameter, therefore exceeding the saturation intensity $I_{sat}=(\pi h c)/(3 \lambda^3 \tau)\approx$ 3.4\,W/m$^2$ of this transition by two orders of magnitude. Here, $h$ is Planck's constant, $c$ is the speed of light, and $\tau= 930$\,ns is the lifetime of the excited state \cite{NIST2010}. For the $F=7/2 \rightarrow F'$ and $F=9/2 \rightarrow F'$ transitions of $^{40}$K we observe a line width of approximately 13\,MHz, somewhat larger than the expected saturation-broadened line width of 5\,MHz, which possibly could result from residual magnetic fields inside the unshielded vapor cell.
\begin{figure}[htbp]
\includegraphics[width=\columnwidth,clip=true]{spec.eps}
\caption{Doppler-free saturation spectrum of the isotopes $^{39}$K, $^{40}$K, and $^{41}$K on the $4S_{1/2} \rightarrow 5P_{1/2}$ transition. The zero of the frequency axis is chosen at the $F=2 \rightarrow F^\prime$ transition of $^{39}$K.}
\label{spectrum}
\end{figure}
Figure \ref{spectrum} shows the measured spectrum of the three isotopes $^{39}$K, $^{40}$K, and $^{41}$K.
The zero of the frequency axis is chosen at the $F=2 \rightarrow F^\prime=1$ transition of $^{39}$K.
Here $F=I+J$ denotes the total angular momentum of the atom comprising of the nuclear spin $I$ ($I=3/2$ for $^{39}$K
and $^{41}$K, and $I=4$ for $^{40}$K) and electronic angular momentum $J=1/2$. We use the notation $F$ to label the $4s\;^2S_{1/2}$
ground state and $F^\prime$ to label the $5p\;^2P_{1/2}$ excited state. We estimate the uncertainty of the frequency calibration to be $0.6$\,MHz.
\section{Hyperfine splitting of the $5p\;^2P_{1/2}$ state of $^{40}$K}
\begin{figure}[htbp]
\includegraphics[width=\columnwidth,clip=true]{hfs.eps}
\caption{(Color online) Close-up view of the Doppler-free signal of $^{40}$K revealing the hyperfine splitting.
The solid lines show the fit to the data to extract the positions of the lines.}
\label{hfs}
\end{figure}
While for the two most abundant isotopes $^{39}$K and $^{41}$K the hyperfine splitting of the $5p\;^2P_{1/2}$ state has been measured previously \cite{Arimondo1977},
no such data have been available for $^{40}$K. Our Doppler-free saturation spectrum directly reveals the hyperfine splitting of the $5p\;^2P_{1/2}$ state of $^{40}$K,
see Figure \ref{spectrum}, on transitions originating from both the ground states $F=7/2$ and $F=9/2$. Figure \ref{hfs} shows a close-up of the two sections together with our fits. On the $F=9/2 \rightarrow F'$ manifold we observe three equally strong lines, which correspond to the $F=9/2 \rightarrow F'=9/2$ and $F=9/2 \rightarrow F'=7/2$ transitions together with a crossover resonance in between. On the $F=7/2 \rightarrow F'$ manifold we observe two strong lines, which we assign to the $F=7/2 \rightarrow F'=9/2$ transition and a crossover resonance,
and a very weak signal which we assign to the $F=7/2 \rightarrow F'=7/2$ transition. Theoretically, this line has a line strength which is factor of five smaller than the $F=7/2 \rightarrow F'=9/2$ line which is compatible with our data. From the splitting of the lines we determine the hyperfine interaction constant $A_{40}^{5p\;^2P_{1/2}}=(-12.0 \pm 0.9)$\,MHz.
Our measured value is in good agreement with the value obtained by scaling the previously known value $A_{39}^{5p\;^2P_{1/2}}=(9.02 \pm 0.17)$\,MHz \cite{Arimondo1977}
by the known ratio of the nuclear g-factors $g_{I,40}/g_{I,39}=-1.24$ which would result in $A_{40}^{5p\;^2P_{1/2}}=(-11.2\pm 0.2)$\,MHz. The much smaller excited state hyperfine splitting of the isotopes $^{39}$K and $^{41}$K is below our resolution.
\section{Isotope shifts}
The isotope shift $\delta \nu^\alpha_{i,j}$ of an atomic transition $\alpha$ between two different isotopes labeled $i$ and $j$ arises from a sum of contributions
of different physical origins, which can be parameterized by the following expression \cite{Martensson1990}
\begin{equation}
\delta \nu^\alpha_{i,j}= (k^\alpha_{NMS}+ k^\alpha_{SMS}) \frac{(M_j-M_i)}{M_i(M_j+m_e)} + F^\alpha \delta\langle r^2 \rangle_{i,j}.
\label{eqn1}
\end{equation}
Here, $m_e$ is the electron mass, $M_i$ the mass of the lighter nucleus and $M_j$ the mass of the heavier nucleus. This choice defines the convention of the sign of the shift. The first contribution is the normal mass shift (NMS) which results from the change of the reduced mass in the effective two-body problem due to the change of the mass of the nucleus. $k_{NMS}^\alpha= \nu_i^\alpha m_e$, with $\nu_i^\alpha=740.529$\,THz \cite{NIST2010} denoting the resonance frequency of the transition. Usually, the normal mass shift is the dominant contribution to the observed isotope shift.
The remaining shift, often referred to as the residual mass shift $\delta \nu^\alpha_{RMS,(i,j)}=k^\alpha_{SMS} \frac{(M_j-M_i)}{M_i(M_j+m_e)} + F^\alpha \delta\langle r^2 \rangle_{i,j}$, results from more subtle effects. The constant $k^\alpha_{SMS}$ describing the specific mass shift is the expectation value of the operator $\sum_{i>j} \textbf{p}_i \textbf{p}_j$ of the different electronic momenta $\textbf{p}_i$. Generally, it is difficult to evaluate this contribution theoretically for atoms with many electrons. However, since $k^\alpha_{SMS}$ depends only on electron-electron correlations, it is a constant for a chosen transition $\alpha$ and does not depend on the nuclear mass. In contrast, the contribution of the field shift $F^\alpha \delta\langle r^2 \rangle_{i,j}$ is determined by the difference of the electron density at the nucleus for two involved states times the difference of the charge radius of the nucleus for the two isotopes. For determining both constants $k^\alpha_{SMS}$ and $F^\alpha$ at least two pairs of isotopes have to be measured if the nuclear charge radii are known.
Our spectrum (see Figure \ref{spectrum}) contains all three isotopes $^{39}$K, $^{40}$K, and $^{41}$K and we determine the isotope shifts. We find $\Delta \nu^{4s\rightarrow 5p}_{39,41}=(454.2\pm 0.8)$\,MHz, which is in agreement with the results of Halloran et al. \cite{Halloran2009} and Iwata et al. \cite{Iwata2010}. The isotope shift of the low-abundance isotope $^{40}$K with respect to $^{39}$K is $\Delta \nu^{4s \rightarrow 5p}_{39,40}=(235.0\pm 2.0)$\,MHz. Our value for this isotope shift is in disagreement with the very recent result of $(207\pm13)$\,MHz \cite{Iwata2010}. This discrepancy could be due to the fact that the hyperfine structure was not resolved in experiment \cite{Iwata2010} or because of additional systematic errors, which were estimated up to another $10\%$ in their setup. From the measured isotope shift we determine the residual mass shift $\Delta \nu^\alpha_{RMS,(i,j)}$. For the residual mass shift we find $\Delta\nu^{4s \rightarrow 5p}_{RMS,(39,41)}=(-54.4\pm 0.8)$\,MHz and $\Delta \nu^{4s\rightarrow 5p}_{RMS,(39,40)}=(-26.0\pm 2.0)$\,MHz.
Using the nuclear charge radii differences for potassium tabulated in reference \cite{Martensson1990} we determine $k^{4s\rightarrow 5p}_{SMS}=(-39\pm5)$\,GHz amu and $F^{4s\rightarrow 5p}=(-43\pm55)$\,MHz fm$^{-2}$ for the $4s\rightarrow 5p$ transition. The error bars are largely determined by the uncertainty of the $\Delta \nu^{4s\rightarrow 5p}_{RMS,(39,40)}$ frequency shift. Our results should be compared to $k^{4s\rightarrow 4p}_{SMS}=(-15.4\pm 3.8)$\, GHz amu and $F^{4s\rightarrow4p} = (-110\pm 3)$\,MHz fm$^{-2}$ on the $4s\rightarrow 4p$ transition \cite{Martensson1990}. We conclude that for the $4s \rightarrow 5p$ transition the specific isotope shift $k_{SMS}$ is larger, possibly because the higher electronically excited state gives rise to larger electron-electron correlations. In contrast, the field shift reduces as compared to the $4s \rightarrow 4p$ transition which could reflect the smaller overlap of the $5p$ wave function with the nucleus as compared to the $4p$ electronic wave function.
\section{Conclusion}
In conclusion, we have measured the hyperfine splitting of the $5p\;^2P_{1/2}$ state of $^{40}$K and determined the isotope shift of the $4s\;^2S_{1/2} \rightarrow 5p\;^2P_{1/2}$ transition, exceeding the previously achieved accuracy for the low-abundance isotope $^{40}$K by one order of magnitude. Our results contribute to determining the atomic structure of potassium more accurately. From the measurement of two isotope shifts we could determine the specific mass shift constant $k^{4s\rightarrow 5p}_{SMS}=(-39\pm5)$\,GHz amu and field shift constant $F^{4s\rightarrow 5p}=(-43\pm55)$\,MHz fm$^{-2}$ of the $4s\rightarrow 5p$ transition. The hyperfine interaction constant of $(-12\pm 0.9)$\,MHz is in agreement with the theoretically anticipated value.
The work has been supported by {EPSRC} (EP/G029547/1) and {ERC} (Grant number 240335).
|
\section{Introduction}
\label{intro}
This section constitutes a survey of our main results: proofs are omitted for ease of reading, as are technical points which are not critical to the main storyline. The remaining sections, with the exception of the appendix, follow the ordering of the introduction with proofs and the necessary details included.
The main goal of this paper is to prove the following theorem. Let $G$ be a $p$-adic unipotent group and consider the Bernstein center of the category of smooth representations of $G$, which we denote by $\mathscr{Z}(G)$ (all of the relevant definitions are given below). Write $\widehat{G}$ for the space of isomorphism classes of irreducible smooth representations of $G$, equipped with the Fell topology, and $C^{\infty}(\widehat{G}) \subset \Fun(\widehat{G})$ for the $\C$-algebras of locally constant and arbitrary $\C$-valued functions on $\widehat{G}$. A suitable version of Schur's lemma tells us that there is a natural algebra homomorphism $\mathscr{Z}(G) \to \Fun(\widehat{G})$, and the results of section \ref{stalkstatements} on ``stalks" of representations may be used to show that this homomorphism is injective. We shall prove that the image of $\mathscr{Z}(G)$ is precisely $C^{\infty}(\widehat{G})$, which is stated in a slightly more general form as Theorem \ref{bernstein} below.
\subsection{Acknowledgements}
First and foremost, I would like to thank Mitya Boyarchenko for his guidance throughout the writing of this paper. The main results were conjectured by him, and in many cases the technical results and the key ideas of proofs are due to him as well. I also express my gratitude to the NSF for funding my summer REU project which eventually gave rise to this paper, and the University of Michigan Mathematics Department for their support of undergraduate research. I would especially like to thank Stephen DeBacker for his encouragement.
\subsection{Basic notions and results}
We aim to study the representation theory of $p$-adic unipotent groups, but for most of this paper we work with totally disconnected topological groups under a few other hypotheses, introduced below, so as not to limit the applicability of our results. First we introduce some convenient terminology.
\begin{definition}
By an $\ell$-\emph{space} we shall mean a Hausdorff, locally compact, and totally disconnected topological space. An $\ell$-\emph{group} is a topological group whose underlying space is an $\ell$-space.
\end{definition}
Recall the following technically indispensable characterization of $\ell$-groups: a Hausdorff topological group $G$ is an $\ell$-group if and only if the identity element of $G$ has a basis of neighborhoods consisting of compact open subgroups. Clearly discrete groups are $\ell$-groups, and compact $\ell$-groups are just profinite groups.
\begin{definition}
Let $G$ be an $\ell$-group and $(\pi,V)$ a complex representation of $G$, meaning a complex vector space $V$ and a homomorphism $\pi : G \to GL(V)$. We shall say that $(\pi,V)$ is \emph{smooth} provided that the action map $G \times V \to V$ is continuous when $V$ is given the discrete topology. If, in addition, the subspace of $V$ fixed by any compact open subgroup of $G$ is finite-dimensional, we call $(\pi,V)$ \emph{admissible}.
\end{definition}
Note that a complex representation $(\pi,V)$ of $G$ is smooth if and only if the map $g \mapsto \pi(g)(v)$ is locally constant for every $v \in V$, which is to say that any vector in $V$ has an open stabilizer in $G$. In the sequel we may abuse terminology by referring to smooth representations simply as representations, since we are only concerned with smooth representations in this paper. We write $\mathcal{R}(G)$ for the category of smooth representations of an $\ell$-group $G$, a full abelian subcategory of the category $\Rep(G)$ of all complex representations of $G$. Recall that if $K$ is a compact $\ell$-group, then the category $\mathcal{R}(K)$ is semisimple.
Given any complex representation $(\rho,W)$ of $G$ (not necessarily smooth), we write $W^{sm}$ for the smooth subrepresentation of $W$ consisting of vectors whose stabilizer is open in $G$, called the \emph{smooth part of} $W$. It is not hard to see that $W \mapsto W^{sm}$ extends to a functor $\Rep(G) \to \mathcal{R}(G)$ which is right adjoint to the inclusion functor $\mathcal{R}(G) \to \Rep(G)$. Now if $(\pi,V)$ is a smooth representation of $G$, the usual dual representation $V^*$ is not necessarily smooth, so we define the \emph{smooth dual of} $(\pi,V)$ to be $V^{\vee} = (V^*)^{sm}$. The usefulness of the admissibility condition is partially explained by the following easy result.
\begin{proposition}
A smooth representation $(\pi,V)$ is admissible if and only if the natural monomorphism $V \to (V^{\vee})^{\vee}$ is an isomorphism.
\label{admissible}
\end{proposition}
The additive group $(\Q_p,+)$ is an important example of an $\ell$-group. The multiplicative group $(\Q_p \setminus \{ 0 \}, \cdot)$ is also an $\ell$-group, but lacks a useful property: it doesn't have enough compact open subgroups ``at the top."
\begin{definition}
An $\ell_c$-\emph{group} is an $\ell$-group which is a filtered union of its compact open subgroups. This means that every element of an $\ell_c$-group is contained in a compact open subgroup, and that any two compact open subgroups are contained in a third.
\end{definition}
For example, $\Z_p \subset p^{-1}\Z_p \subset p^{-2}\Z_p \subset \cdots$ is a filtration of $(\Q_p,+)$ by compact open subgroups. By taking (topological group) products it is easy to see that the additive group of any finite-dimensional $p$-adic vector space $V$ is an $\ell_c$-group. Recall that the Pontryagin dual of $V$ is noncanonically isomorphic to $V$ as a topological group, and more generally, the Pontryagin dual of an abelian $\ell_c$-group is totally disconnected (therefore an $\ell$-group). The next theorem tells us that smooth representations of such a group can be viewed as sheaves of complex vector spaces on its Pontryagin dual in a natural way.
\begin{theorem}[Rodier]
Let $G$ be an abelian $\ell_c$-group. Then the category $\mathcal{R}(G)$ is equivalent to the category $\Sh(\widehat{G})$ of sheaves of complex vector spaces on the Pontryagin dual of $G$.
\label{rodier}
\end{theorem}
Fran\c cois Rodier proved this remarkable theorem in \cite{Rod}. Notice that the category $\Sh(\widehat{G})$ depends only on the underlying topological space of $\widehat{G}$. The appendix to this paper covers some of the same material and includes an alternate proof of this result (in English).
\subsection{$p$-adic unipotent groups}
A more interesting (in particular, nonabelian) class of $\ell_c$-groups are the $p$-\emph{adic unipotent groups}, which are constructed as follows: let $\mathfrak{g}$ be a finite-dimensional nilpotent Lie algebra over $\Q_p$. The corresponding $p$-adic unipotent group $G$ has the same underlying topological space as $\mathfrak{g}$, with group operation given by the Campbell-Hausdorff formula. More explicitly: for psychological reasons let $\exp : \mathfrak{g} \to G$ and $\log : G \to \mathfrak{g}$ denote the identity map, and then the group operation is \[ \log(\exp(x) \cdot \exp(y)) = x + y + \frac12 [x,y] + \frac{1}{12} [x,[x,y]] - \frac{1}{12} [y,[x,y]] - \cdots \ \ \ \ x,y \in \mathfrak{g}. \] There is no convergence issue because $\mathfrak{g}$ is nilpotent.
More abstractly, a unipotent algebraic group $\mathbb{G}$ over $\Q_p$ gives rise to a $p$-adic unipotent group $G = \mathbb{G}(\Q_p)$, the group of $\Q_p$-points endowed with the $p$-adic topology. Then we can recover $\mathfrak{g}$ in the usual way as the tangent space of $\mathbb{G}$ at the identity under the Lie bracket.
The simplest class of nonabelian $p$-adic unipotent groups are the $p$-\emph{adic Heisenberg groups}. Let $V$ be a finite-dimensional $p$-adic vector space equipped with a symplectic form \[ \omega : V \times V \to \Q_p, \] so that in particular $V$ has even dimension. As a topological space, $G = V \times \Q_p$. The group operation on $G$ is given by the formula \[ (v,a) \cdot (w,b) = (v+w,a+b+\tfrac12 \omega(v,w)). \]
If we let $Z$ denote the center of $G$, which is isomorphic to $\Q_p$ as a topological group, it is not hard to see that $G$ is a central extension of the additive group of $V$ by $Z$: \[ 1 \to Z \to G \to V \to 1. \]
All the irreducible smooth representations of an abelian $\ell_c$-group are one-dimensional, which we prove in the appendix to this paper, so in particular they are admissible. In fact, it is proved in \cite{Rod} that $p$-adic unipotent groups also have this property, which is so useful that we will give it a name for our convenience.
\begin{definition}
We call an $\ell$-group \emph{tame} if all of its irreducible smooth representations are admissible.
\end{definition}
The proof that abelian $\ell_c$-groups are tame is outlined in the appendix to this paper, and Rodier proved that $p$-adic unipotent groups are tame. It should be noted that tameness implies Schur's lemma by an argument in \cite{Pra}.
\begin{proposition}
If $G$ is a tame $\ell$-group and $\pi$ is an irreducible smooth representation of $G$, then the only linear endomorphisms of $\pi$ which commute with the action of $G$ are the scalars.
\label{schur}
\end{proposition}
\subsection{The dual space and its Fell topology}
It is natural to ask if Theorem \ref{rodier} has some sort of generalization to $p$-adic unipotent groups or, more generally, tame $\ell_c$-groups. The first step is to find the appropriate dual space.
\begin{definition}
The \emph{smooth dual space} $\widehat{G}$ of an $\ell$-group $G$ is the space of isomorphism classes of irreducible smooth representations of $G$.
\end{definition}
For abelian $\ell_c$-groups, $\widehat{G}$ coincides with the Pontryagin dual and in particular carries a natural group structure. Recall that the additive group $V$ of a finite-dimensional $p$-adic vector space can be noncanonically identified with $\widehat{V}$ as a topological group. Even for nonabelian groups, there is a canonical topology on $\widehat{G}$ which generalizes the ``compact-open" topology usually employed in Pontryagin duality, called the \emph{Fell topology}. The Fell topology is unpleasant to describe in general, but the interested reader can refer to the beginning of Section \ref{topology}. This brings us to our first original theorem: a more understandable characterization of the Fell topology in the cases we are interested in.
\begin{theorem}
Let $G$ be a tame $\ell_c$-group. For any compact open subgroup $K \subset G$ and $\rho \in \widehat{K}$, put \[ \mathscr{V}(K,\rho) = \{ \pi \in \widehat{G} \ | \ \Hom_K(\pi|_K,\rho) \neq 0 \}. \] Then the $\mathscr{V}$'s are open and form a basis for the Fell topology on $\widehat{G}$, meaning every open set in $\widehat{G}$ can be written as a union of $\mathscr{V}$'s.
\label{Fell}
\end{theorem}
One reason we study the representation theory of a $p$-adic unipotent group $G$ is that then $\widehat{G}$ admits a geometric description via the orbit method. Again we fix a nilpotent $p$-adic Lie algebra $\mathfrak{g}$ and let $G$ denote the corresponding $p$-adic unipotent group. Recall that $G$ acts by Lie algebra automorphisms on $\mathfrak{g}$ via conjugation (the adjoint action), hence also acts linearly on the dual vector space $\mathfrak{g}^*$ (the coadjoint action). The orbit method gives a homeomorphism \[ \mathfrak{g}^*/G \longrightarrow \widehat{G}. \] This is even more helpful when combined with the algebro-geometric result that the orbits in an affine variety of an algebraic action by a unipotent group are closed.
When $G$ is the Heisenberg group, the orbits of $G$ in $\mathfrak{g}^*$ are as follows. Write $\mathfrak{g}^*_0 \subset \mathfrak{g}^*$ for the subspace consisting of those $f \in \mathfrak{g}^*$ which vanish on the center of $\mathfrak{g}$: then every point of $\mathfrak{g}^*_0$ is fixed by $G$, and each nonzero translate (i.e. coset) of $\mathfrak{g}^*_0$ is a $G$-orbit. Under the orbit method homeomorphism $\mathfrak{g}^*_0$ corresponds to those elements of $\widehat{G}$ which are trivial on the center $Z$ of $G$, and we can identify this (topological) subspace with $\widehat{V}$, where $V$ denotes the additive group of the symplectic vector space used to construct $G$. The nonzero translates of $\mathfrak{g}^*_0$ correspond to the representations in $\widehat{G}$ which are nontrivial on $Z$, and these may be identified with $\widehat{Z} \setminus \{ 0 \}$ by a suitable version of the Stone-von Neumann theorem. So as a set $\widehat{G}$ is a disjoint union of $\widehat{V}$ and $\widehat{Z} \setminus \{ 0 \}$, endowed with a rather strange topology. In particular, if we choose a sequence of points in $\widehat{Z} \setminus \{ 0 \}$ which converges to $0$ in $\widehat{Z}$, the limit of this sequence in $\widehat{G}$ is all of $\widehat{V}$. The topology on $\widehat{G}$ is locally compact and totally disconnected, but plainly not Hausdorff.
\subsection{Stalks in the category $\mathcal{R}(G)$}
\label{stalkstatements}
For now, $G$ is a tame $\ell_c$-group, such as a $p$-adic unipotent group. It seems that $\mathcal{R}(G)$ shares many properties with $\Sh(\widehat{G})$, the category of sheaves of complex vector spaces on $\widehat{G}$, including the existence of a ``stalk" functor for each $\pi \in \widehat{G}$ that naturally returns the muliplicity (appropriately defined) of $\pi$ in smooth representations of $G$. But first, there is a meaningful notion of support for smooth representations.
\begin{definition}
The \emph{support} of a smooth representation $M \in \mathcal{R}(G)$ is defined by \[ \Supp M = \{ \pi \in \widehat{G} \ | \ \Hom_G(M,\pi) \neq 0 \}. \]
\end{definition}
For example, if $S \subset \widehat{G}$ is any subset of the dual space, then $\Supp \oplus_{\pi \in S} \pi = S$, but in general irreducible representations may be combined in more complicated ways than direct sum. Let's write $\mathcal{R}(G)_S$ for the full subcategory of smooth representations of $G$ with support contained in $S$. If $S = \{ \pi \}$ is a singleton we'll abuse notation and denote $\mathcal{R}(G)_S$ by $\mathcal{R}(G)_{\pi}$.
The next result gives a simple description of $\mathcal{R}(G)_{\pi}$ via an analogue of the skyscraper sheaf construction: let $\C-\mathbf{vect}$ denote the category of complex vector spaces.
\begin{theorem}
Fix a point $\pi \in \widehat{G}$. The functor $\C-\mathbf{vect} \longrightarrow \mathcal{R}(G)$ which sends $V \mapsto V_G \otimes \pi$, where $V_G$ denotes the trivial representation of $G$ on $V$, is an equivalence onto $\mathcal{R}(G)_{\pi}$.
\label{skyscraper}
\end{theorem}
In view of the analogy with sheaves it is not surprising then that this functor has a left adjoint, which sends a smooth representation to its ``stalk" at $\pi$. This can be seen by factoring the skyscraper functor $\C-\mathbf{vect} \to \mathcal{R}(G)$ through the inclusion $\mathcal{R}(G)_{\pi} \to \mathcal{R}(G)$, which has a left adjoint described as follows: for any $M \in \mathcal{R}(G)$, write $M_{\pi}$ for the largest quotient of $M$ supported at $\pi$. Explicitly, put \[ M_{\pi} = M / \bigcap_{f \in \Hom_G(M,\pi)} \ker f. \] The content of the claim that $M \mapsto M_{\pi}$ is left adjoint to the inclusion $\mathcal{R}(G)_{\pi} \to \mathcal{R}(G)_{\pi}$ is the following.
\begin{lemma}
For any $M \in \mathcal{R}(G)$ and $\pi \in \widehat{G}$ we have $M_{\pi} \in \mathcal{R}(G)_{\pi}$.
\label{point}
\end{lemma}
Now we define $M(\pi) = \Hom_G(\pi,M_{\pi})$, the vector space we want to think of as the stalk of $M$ at $\pi$.
These stalk functors have some useful properties in common with sheaf-theoretic stalk functors. For example, $M = 0$ if and only if $M(\pi) = 0$ for every $\pi \in \widehat{G}$: this follows from the fact that $M = 0$ if and only if $\Supp M = \varnothing$. The nontrivial content of this statement is that every nonzero smooth representation has an irreducible quotient, which is to say a maximal subrepresentation. Much more generally, the following result allows us to adapt the proofs of many statements about sheaves to the setting of smooth representations.
\begin{lemma}
A sequence $M \to N \to P$ in $\mathcal{R}(G)$ is exact if and only if the associated sequence of vector spaces $M(\pi) \to N(\pi) \to P(\pi)$ is exact for every $\pi \in \widehat{G}$.
\label{exact}
\end{lemma}
So we can test exactness on stalks. In particular, a morphism of smooth representations is a monomorphism (respectively epimorphism, isomorphism) if and only if the associated map of stalks is a monomorphism (respectively epimorphism, isomorphism) at every point in the dual space.
\subsection{The second characterization of the Fell topology}
Now we state our next characterization of the Fell topology, which has a categorical flavor, but first let us fix some notation. Consider subsets $Z \subset W \subset \widehat{G}$ and write $i_! : \mathcal{R}(G)_Z \to \mathcal{R}(G)_W$ for the inclusion functor.
This notation suggests an analogy with the pushforward functor for sheaves, and indeed our situation is very similar on a formal level. We might ask for a left adjoint to $i_!$, and our wish is granted as long as $Z$ is closed. But in fact something even better is true.
\begin{theorem}
If $W \subset \widehat{G}$ is locally closed and $Z \subset W$, then the inclusion functor $i_! : \mathcal{R}(G)_Z \to \mathcal{R}(G)_W$ admits a left adjoint $i^* : \mathcal{R}(G)_W \to \mathcal{R}(G)_Z$ if and only if $Z$ is closed in $W$. When $i^*$ exists it is exact.
\label{fell2}
\end{theorem}
Notice that $i^*$ exists when $Z$ is a point and $W = \widehat{G}$ by Lemma \ref{point}, which yields the following corollary.
\begin{corollary}
For any tame $\ell_c$-group $G$, the Fell topology on $\widehat{G}$ is $T_1$.
\label{T1}
\end{corollary}
Theorem \ref{fell2} has another interesting consequence.
\begin{corollary}
An irreducible representation $\pi \in \widehat{G}$ is projective in $\mathcal{R}(G)$ if and only if $\pi$ is an open point of $\widehat{G}$ in the Fell topology.
\label{projective}
\end{corollary}
The proof of Corollary \ref{projective} may not be entirely obvious and is included in Section \ref{adjointproofs}.
\subsection{An interlude on Hecke algebras}
Before we state our last characterization of the Fell topology on $\widehat{G}$ for $G$ a tame $\ell_c$-group, which says that locally the dual space is homeomorphic to the primitive spectrum of a certain unital Hecke algebra equipped with the Jacobson (sometimes called hull-kernel) topology, we need a couple of definitions.
\begin{definition}
Write $\mathcal{H}(G)$ for the \emph{Hecke algebra of} $G$, that is, the convolution algebra of locally constant $\C$-valued functions on $G$ with compact support, which is unital only when $G$ is discrete.
\end{definition}
Recall that the category $\mathcal{R}(G)$ is isomorphic to the full subcategory of $\mathcal{H}(G)$-modules $M$ which are non-degenerate in the sense that $\mathcal{H}(G) \cdot M = M$.
\begin{definition}
Fix a compact open subgroup $K \subset G$ and $\rho \in \widehat{K}$. We write $\mathcal{H}(G,\rho)$ for the convolution algebra of locally constant functions $f : G \to \End_{\C}(W^{\vee})$ with compact support satisfying \[ f(k_1gk_2) = \rho^{\vee}(k_1) \circ f(g) \circ \rho^{\vee}(k_2) \ \ \ \ k_1,k_2 \in K, \ g \in G. \]
\end{definition}
It is well-known that $\mathcal{H}(G,\rho)$ is canonically isomorphic to $\End_G(\cInd_K^G \rho)$ as a $\C$-algebra, so that for any smooth representation $M \in \mathcal{R}(G)$ the space $\Hom_K(\rho,M|_K)$ of $\rho$-invariants has a natural $\mathcal{H}(G,\rho)$-module structure. We point the reader to \cite{BK} for a detailed discussion, with proofs, of these Hecke algebras. The relevant definitions, notations, and facts about induced representations and Frobenius reciprocity may be found in Section \ref{heckeproofs}.
Write $\mathcal{R}_{\rho}(G)$ for the full subcategory of $\mathcal{R}(G)$ consisting of smooth representations $M$ with the property that $M$ is generated by its $\rho$-isotypic component (the sum of all $K$-subrepresentations of $M$ isomorphic to $\rho$). The next theorem is not so hard, but gives some formal motivation for the more interesting Theorem \ref{fell3} and will be quite useful to us.
\begin{theorem}
If $G$ is a tame $\ell_c$-group, $K \subset G$ is any compact open subgroup, and $\rho \in \widehat{K}$, the functor of $\rho$-invariants that sends $M \mapsto \Hom_K(\rho,M|_K)$ restricts to an equivalence of categories \[ \mathcal{R}_{\rho}(G) \to \mathcal{H}(G,\rho)-\mathbf{mod}. \]
\label{type}
\end{theorem}
It is worth noting that if we do not assume that $G$ is an $\ell_c$-group then this theorem is known to be false in general. In the representation theory of $p$-adic reductive groups a pair $(K,\rho)$ for which the theorem holds is called a \emph{type}.
\subsection{The Fell topology revisited}
Let $A$ be any unital associative ring, and recall that a \emph{left primitive ideal} of $A$ is the annihilator of a simple left $A$-module (one may have this discussion for non-unital rings, but we are working locally with unital rings anyway). We will write $\widehat{A}$ for the space of left primitive ideals of $A$. The \emph{Jacobson} or \emph{hull-kernel topology} on $\widehat{A}$ is defined as follows: for an arbitrary subset $S \subset \widehat{A}$ we define the \emph{kernel of} $S$ \[ J_S = \bigcap_{I \in S} I \] to be the intersection of the ideals in $S$. The closure of $S$ is the set of primitive ideals which contains $J_S$, and this closure operator defines the Jacobson topology on $\widehat{A}$.
Now take $A = \mathcal{H}(G,\rho)$. Clearly Theorem \ref{type} implies that the functor of $\rho$-invariants induces a bijection $\mathscr{V}(K,\rho) \to \widehat{A}$, and in fact one can show this directly even when the theorem fails, e.g. when $G$ is reductive and $(K,\rho)$ is not a type. Finally, we state our third characterization of the Fell topology.
\begin{theorem}
Let $G$ be a tame $\ell_c$-group. Then the bijection $\mathscr{V}(K,\rho) \to \widehat{A}$ which takes $\rho$-invariants is a homeomorphism, where $\widehat{A}$ is given the Jacobson topology.
\label{fell3}
\end{theorem}
We point out that in \cite{GK}, the theorem is proved for trivial $\rho$.
\subsection{The Bernstein center of $\mathcal{R}(G)$}
For any abelian category $\mathcal{A}$ we denote by $\mathscr{Z}(\mathcal{A})$ the Bernstein center of $\mathcal{A}$, that is, the commutative $\C$-algebra of endomorphisms of the identity functor on $\mathcal{A}$. For an $\ell$-group $G$ we may consider the Bernstein center of $\mathcal{R}(G)$, which we abusively denote as $\mathscr{Z}(G)$ and call the Bernstein center of $G$, or more generally, if $W \subset \widehat{G}$ we write $\mathscr{Z}_W(G)$ for $\mathscr{Z}(\mathcal{R}(G)_W)$. It is possible to say a few things about $\mathscr{Z}_W(G)$ in our situation, meaning when $G$ is a tame $\ell_c$-group. Notably, it may be considered as an algebra of $\C$-valued functions on $W$, since in light of Proposition \ref{schur} an element of $\mathscr{Z}_W(G)$ acts on each $\pi \in W$ by a scalar. In fact, if we write $\Fun(W)$ for the $\C$-algebra of $\C$-valued functions on $W$ under pointwise multiplication, then the map $\mathscr{Z}_W(G) \longrightarrow \Fun(W)$ thus defined is an algebra homomorphism. The following fact is an easy consequence of Lemma \ref{exact}.
\begin{proposition}
The natural algebra homomorphism $\mathscr{Z}_W(G) \longrightarrow \Fun(W)$ is injective.
\end{proposition}
In case $G$ is abelian and $W = \widehat{G}$, the image of this homomorphism is $C^{\infty}(\widehat{G})$, the algebra of locally constant $\C$-valued functions on $\widehat{G}$. This follows formally from the equivalence of $\mathcal{R}(G)$ with $\Sh(\widehat{G})$. Indeed, if $X$ is any topological space, it is well known that the Bernstein center of $\Sh(X)$ is $C^{\infty}(X)$. Our final theorem is a generalization of this.
\begin{theorem}
If $G$ is a tame $\ell_c$-group and $W \subset \widehat{G}$ is locally closed, then the image of the natural homomorphism $\mathscr{Z}_W(G) \longrightarrow \Fun(W)$ is $C^{\infty}(W)$.
\label{bernstein}
\end{theorem}
The rest of this paper consists of some additional technical exposition and the proofs of the theorems and lemmas stated above.
\section{More on Hecke algebras}
Before proving our first characterization of the Fell topology, we include a general discussion of Hecke algebras. For now $G$ is any $\ell$-group, and we must also fix a left Haar measure on $G$.
Recall that we have written $\mathcal{H}(G)$ for the Hecke algebra of $G$. If $K \subset G$ is a compact open subgroup, the subalgebra of bi-$K$-invariant functions will be denoted by $\mathcal{H}(G,K)$. Explicitly, $\mathcal{H}(G,K)$ consists of those $f \in \mathcal{H}(G)$ such that for any $g \in G$ and $k_1,k_2 \in K$, we have $f(k_1gk_2) = f(g)$. It is easy to show that $\mathcal{H}(G) = \bigcup \mathcal{H}(G,K)$, where the union runs over all compact open subgroups $K \subset G$: every function in the Hecke algebra is just a finite linear combination of indicator functions of double cosets $KgK$ for some sufficiently small compact open subgroup $K \subset G$.
The subalgebra $\mathcal{H}(G,K)$ has a unity element, namely $e_K = \frac{1}{\mu(K)} \cdot \mathbbm{1}_K$, where $\mathbbm{1}_K$ is the indicator function of $K$. In fact, $\mathcal{H}(G,K) = e_K * \mathcal{H}(G) * e_K$. Thus if $(\pi,V)$ is any smooth representation of $G$, the subspace of $K$-invariant vectors $V^K = e_k \cdot V$ is naturally an $\mathcal{H}(G,K)$-module. In particular, if $\pi$ is irreducible then either $V^K = 0$ or $V^K$ is a simple $\mathcal{H}(G,K)$-module (this fact is not specific to our situation, and holds much more generally for modules over an idempotented algebra).
Now if $\rho \in \widehat{K}$ we can define its character $\chi_{\rho} : K \to \C$ in the usual way, as irreducible representations of compact $\ell$-groups are finite-dimensional. Since $\rho$ factors through a representation of a finite quotient of $K$, we can use the orthogonality relations for characters of finite groups to see that $e_{\rho} = \frac{\chi_{\rho}(1)}{\mu(K)} \cdot \chi_{\rho}$ is an idempotent in the Hecke algebra. Moreover, just as in the finite case, if $(\pi,V)$ is any representation of $G$ then $e_{\rho} \cdot V$ is the sum of all $K$-subrepresentations of $\pi$ which are isomorphic to $\rho$. Observe also that $e_{\rho} \cdot V$ is naturally a module for the unital subalgebra $e_{\rho} * \mathcal{H}(G) * e_{\rho}$ of $\mathcal{H}(G)$, and that if $\pi$ is irreducible then $e_{\rho} \cdot V = 0$ or $e_{\rho} \cdot V$ is a simple module for $e_{\rho} * \mathcal{H}(G) * e_{\rho}$. If $\rho$ is trivial then $e_{\rho} = e_K$, so these remarks generalize the previous paragraph.
Recall that in Section 1 we defined a certain Hecke algebra $\mathcal{H}(G,\rho)$ which plays a role in our third characterization of the Fell topology. In fact, there is a canonical isomorphism of $\C$-algebras \[ \mathcal{H}(G,\rho) \otimes_{\C} \End_{\C}(W) \longrightarrow e_{\rho} * \mathcal{H}(G) * e_{\rho}, \] which means the algebras $\mathcal{H}(G,\rho)$ and $e_{\rho} * \mathcal{H}(G) * e_{\rho}$ are Morita equivalent. Note that these two algebras may be identified when $\rho$ is trivial (then they both coincide with $\mathcal{H}(G,K)$), but otherwise $\mathcal{H}(G,\rho)$ is not a subalgebra of $\mathcal{H}(G)$.
\section{The proof of Theorem \ref{Fell}}
\label{topology}
Now we tie up the first loose thread from Section \ref{intro} and prove the characterization of the Fell topology stated in Theorem \ref{Fell}. Throughout this section $G$ will be any tame $\ell_c$-group.
There is a canonical choice of topology on the dual space $\widehat{G}$: if $(\pi,V) \in \widehat{G}$ pick a finite collection of vectors $v_1,\cdots,v_n \in V$, another $\xi_1,\cdots,\xi_n \in \widetilde{V}$, a compact subset $B \subset G$, and $\epsilon > 0$, then let $\mathscr{U}(\pi,v_j,\xi_j,B,\epsilon)$ consist of those $(\rho,W) \in \widehat{G}$ for which we can find $w_1,\cdots,w_n \in W$ and $\zeta_1,\cdots,\zeta_n \in \widetilde{W}$ satisfying \[ \big| \ \langle \xi_j, \pi(g)(v_j) \rangle - \langle \zeta_j, \rho(g)(w_j) \rangle \ \big| < \epsilon \] for all $1 \leq j \leq n$ and $g \in B$. The collection of all such $\mathscr{U}$'s forms a basis of neighborhoods for the Fell topology on $\widehat{G}$, which we have characterized in a way which we hope the reader finds more understandable.
The following technical lemma is critical in our proof of Theorem \ref{Fell}.
\begin{lemma}
Fix $(\pi,V) \in \widehat{G}$ and a nonzero finite-dimensional subspace $W \subset V$. Then there exists a compact open subgroup $K \subset G$ such that the $K$-subrepresentation generated by $W$ is irreducible.
\label{lemma1}
\end{lemma}
\begin{proof}
Choose a compact open subgroup $L \subset G$ such that $W \subset V^L$, so we know that $V^L$ is a simple $\mathcal{H}(G,L)$-module. Moreover, $V^L$ is finite-dimensional since $G$ is tame. Observe that $\mathcal{H}(G,L)$ is a filtered union of finite-dimensional subalgebras $\mathcal{H}(K,L)$, where $K$ ranges over compact open subgroups of $G$ containing $L$, by the $\ell_c$ assumption. Now let \[ \varphi : \mathcal{H}(G,L) \longrightarrow \End_{\C}(V^L) \] be the action map and put $A = \im \varphi$ (in fact, $\varphi$ must be surjective, but this is not necessary for the proof). Then $A$ is a finite-dimensional quotient of $\mathcal{H}(G,L)$, so there exists a compact open subgroup $K \supset L$ such that $\mathcal{H}(K,L)$ surjects onto $A$. But this means $V^L$ is already a simple $\mathcal{H}(K,L)$-module.
Now let $(\rho,X)$ be the $K$-subrepresentation generated by $V^L$ (so if we can show that $\rho$ is irreducible, $W$ generates $X$ also). Suppose $X$ decomposes as a $K$-representation into $X_1 \oplus X_2$. Then \[ V^L = X^L = X_1^L \oplus X_2^L \] implies $X_1^L = 0$ or $X_2^L = 0$ by simplicity of $V^L$, whence $V^L \subset X_1$ or $V^L \subset X_2$. Thus $X_1 = X$ or $X_2 = X$ as desired, since $V^L$ generates $X$.
\end{proof}
It seems natural to mention here one consequence of the lemma, which does not play a role in the proof of Theorem \ref{Fell} but will help us later in this paper.
\begin{lemma}
For any $\pi \in \widehat{G}$, we can find a compact open subgroup $K \subset G$ and $\rho \in \widehat{K}$ such that $| \rho : \pi|_K | = 1$, which is to say the space $\Hom_K(\rho,\pi|_K)$ is one-dimensional.
\end{lemma}
\begin{proof}
First find some compact open subgroup $L \subset G$ such that the space of $L$-invariants $\pi^L$ is nonzero. Then apply Lemma \ref{lemma1} to find a compact open subgroup $K$ such that the $K$-subrepresentation $\rho$ of $\pi$ generated by $\pi^L$ is irreducible. In particular, $\rho^L \neq 0$, and if $\rho \to \pi|_K$ is a $K$-morphism it must map $\rho^L$ into $\pi^L \subset \rho$, whence the space of such morphisms is one-dimensional by Schur's lemma.
\end{proof}
Now we are in a position to prove the theorem.
\begin{proof}[Proof of Theorem \ref{Fell}]
First we show that $\mathscr{V}(K,\rho)$ is open. Resuming our notation in the statement of the theorem, choose $(\pi,V) \in \mathscr{V}(K,\rho)$. By our remarks in section 1.2 there exists $v \in V$ such that $e_{\rho} \cdot v \neq 0$, and certainly there also exists $f \in \widetilde{V}$ such that $f(e_{\rho} \cdot v) \neq 0$. Now put $L = \ker \rho$ and $C = \max_{k \in K/L} |\chi_{\rho}(k)|$, and let $0 < \epsilon < \frac{|f(e_{\rho} \cdot v)|}{C \cdot \chi_{\rho}(1)}$. Then it is not hard to see by a direct calculation that $\mathscr{U}(\pi,v,f,K,\epsilon) \subset \mathscr{V}(K,\rho)$.
Now to prove that the $\mathscr{V}$'s form a basis for the Fell topology, let $(\pi,V) \in \widehat{G}$ be arbitary and choose $v_1,\cdots,v_n \in V$, $f_1,\cdots,f_n \in \widetilde{V}$, compact $B \subset G$, and $\epsilon > 0$. This defines a Fell neighborhood $\mathscr{U}(\pi,v_j,f_j,B,\epsilon)$ of $\pi$. Let $W$ be the $\C$-span of the $v_1,\cdots,v_n$ and apply Lemma \ref{lemma1} to find a compact open subgroup $K \subset G$ such that the $K$-subrepresentation $\rho$ generated by $W$ is irreducible. Note that we can use the fact that $G$ is an $\ell_c$-group to choose $K$ large enough so that $B \subset K$. An easy computation confirms that $\mathscr{V}(K,\rho) \subset \mathscr{U}(\pi,v_j,f_j,B,\epsilon)$.
\end{proof}
\section{Proving the results on stalks in the category $\mathcal{R}(G)$}
\label{stalkfunct}
In this section we prove the assertions made about the functors $\pi \mapsto M(\pi)$ for $\pi \in \widehat{G}$, with the goal of applying some sheaf-theoretic techniques to smooth representations. These results are independent of our previous work but play an important role in the proofs that follow. Recall that we denote the largest quotient of a smooth representation $M$ supported at $\pi$ by $M_{\pi}$, and that $M(\pi)$ is the ``multiplicity space" of $\pi$ in $M_{\pi}$.
First we need some technical consequences of the $\ell_c$-condition. Recall that we can define the so-called Jacquet functor $J_G : \mathcal{R}(G) \to \C-\mathbf{vect}$ by sending $M \in \mathcal{R}(G)$ to the the space of coinvariants, i.e. the quotient of $M$ by the $\C$-span of $\{ g \cdot v - v \ | \ v \in M \}$. This extends to a right exact functor which is left adjoint to the functor $\C-\mathbf{\vect} \to \mathcal{R}(G)$ that sends a vector space $V$ to the trivial representation of $G$ on $V$.
\begin{proposition}
For an $\ell_c$-group $G$, the functor $J_G$ is exact.
\end{proposition}
Jacquet proved this proposition: see, for instance, \cite{BZ}.
\begin{lemma}
If $G$ is an $\ell_c$-group and $M \in \mathcal{R}(G)$ the smooth dual $M^{\vee}$ is an injective object in $\mathcal{R}(G)$.
\label{injective}
\end{lemma}
\begin{proof}
For any $N \in \mathcal{R}(G)$ we have natural isomorphisms
\begin{align*}
\Hom_G(N,M^{\vee}) &= \Hom_G(N,(M^*)^{sm}) \cong \Hom_G(N,\Hom_{\C}(M,\C)) \\
&\cong \Hom_G(N \otimes_{\C} M,\C) \cong \Hom_{\C}(J_G(N \otimes_{\C} M),\C).
\end{align*}
Then we see that $M \mapsto M^{\vee}$ is the composition of three exact functors, since vector spaces are flat and injective and $J_G$ is exact.
\end{proof}
Note that Proposition \ref{admissible} and Lemma \ref{injective} imply the following key fact.
\begin{proposition}
Irreducible smooth representations of tame $\ell_c$-groups are injective objects in $\mathcal{R}(G)$.
\label{inj}
\end{proposition}
At this point we can apply Zorn's lemma to see that our definition of support was a reasonable choice.
\begin{lemma}
If $M \in \mathcal{R}(G)$, then $M = 0$ if and only if $\Supp M = \varnothing$.
\label{emptysupp}
\end{lemma}
\begin{proof}
The ``only if" direction is trivial. For the other implication, it suffices to show that any nonzero $M \in \mathcal{R}(G)$ has an irreducible quotient. For this, choose a nonzero finitely generated subrepresentation $N \subset M$, whence an easy application of Zorn's lemma shows that $N$ has an irreducible quotient $\pi \in \widehat{G}$. Then use Proposition \ref{inj} to see that $\pi$ is in fact a quotient of $M$.
\end{proof}
\begin{corollary}
Let $f : M \to N$ be a morphism in $\mathcal{R}(G)$. Then $f = 0$ if and only if for each $\pi \in \widehat{G}$ and $g \in \Hom_G(N,\pi)$, we have $g \circ f = 0$.
\label{zeromap}
\end{corollary}
\begin{proof}
Again we need only prove the ``if" direction, and by the lemma it is enough to show that $\Supp \im f = \varnothing$. But this is clear, since every morphism in $\Hom_G(\im f,\pi)$ lifts to $N$ by injectivity of $\pi$.
\end{proof}
Here we prove Theorem \ref{skyscraper}, which roughly says that smooth representations are ``pointwise semisimple."
\begin{proof}[Proof of Theorem \ref{skyscraper}]
The functor $V \mapsto V_G \otimes \pi$ is obviously fully faithful, and to see that it is essentially surjective it is enough to show that every $M \in \mathcal{R}(G)_{\pi}$ is a direct sum of copies of $\pi$. For this, observe that the proof of Lemma \ref{emptysupp} can be adapted to show that the natural map \[ M \longrightarrow \prod_{\Hom_G(M,\pi)} \pi \] is a monomorphism: apply Zorn's lemma to prove that the subrepresentation generated by any nonzero $m \in M$ has an irreducible quotient, and use Proposition \ref{inj} to see that this subquotient of $M$ is in fact a quotient, necessarily isomorphic to $\pi$. More concisely, every nonzero element of $M$ survives in an irreducible quotient.
Now to finish the proof, we claim that if $I$ is any set, the smooth part of the product $\prod_I \pi$ is a sum of copies of $\pi$. But in fact the (clearly injective) map given below is an isomorphism.
\begin{align*}
\bigg( \prod_I \C \bigg) \otimes \pi & \longrightarrow \bigg( \prod_I \pi \bigg)^{sm} \\
\bigg( \prod_{i \in I} a_i,v \bigg) & \mapsto \prod_{i \in I} a_iv
\end{align*}
To see why, choose $v_i \in \pi$ for each $i \in I$ so that $\prod v_i$ is smooth, i.e. lies in the smooth part of $\prod_I \pi$. Then $G$ has a compact open subgroup $K$ such that $\prod v_i$ is contained in the subspace $(\prod_I \pi)^K = \prod_I \pi^K$ of $K$-invariants. Now $\dim_{\C} \pi^K < \infty$ by the tameness assumption, so we can choose a finite spanning set $x_1,\cdots,x_m$ for $\pi^K$. Finally, write $v_i = \sum_j a_{ij}x_j$ and observe that $\sum_j (\prod_i a_{ij}) \otimes x_j$ is sent to $\prod v_i$.
\end{proof}
Next we prove Lemma \ref{point}.
\begin{proof}[Proof of Lemma \ref{point}]
Notice that $M_{\pi}$ is defined precisely so that the natural morphism \[ M_{\pi} \to \prod_{f \in \Hom_G(M_{\pi},\pi)} \pi \] is a monomorphism. But the smooth part of $\prod \pi$ is a direct sum of copies of $\pi$ by the proof of Theorem \ref{skyscraper}.
\end{proof}
Recall that exactness of sequences in $\mathcal{R}(G)$ can be tested on stalks: let us prove Lemma \ref{exact}.
\begin{proof}[Proof of Lemma \ref{exact}]
To see that the functor $M \mapsto M(\pi)$ is exact, observe that there is a sequence of natural isomorphisms \[ \Hom_G(M,\pi) \cong \Hom_G(M_{\pi},\pi) \cong \Hom_G(M(\pi) \otimes \pi,\pi) \cong M(\pi)^*, \] where the first isomorphism uses the above-mentioned adjunction and the latter two use Proposition \ref{schur}. But the functor $M \mapsto \Hom_G(M,\pi)$ is exact by injectivity of $\pi$, which means that the the dual sequence \[ P(\pi)^* \longrightarrow N(\pi)^* \longrightarrow M(\pi)^* \] is exact, and consequently so is the sequence \[ M(\pi) \longrightarrow N(\pi) \longrightarrow P(\pi). \]
For the converse, we first prove that $g \circ f = 0$. Fix $\pi \in \widehat{G}$ and $h \in \Hom_G(P,\pi)$, so by Corollary \ref{zeromap} it suffices to show that $h \circ g \circ f = 0$. Well, since $(g \circ f)_{(\pi)} = g_{(\pi)} \circ f_{(\pi)} = 0$ it is easy to see that the induced map $M_{\pi} \to P_{\pi}$ is zero, and then we have what we need because $h$ factors through the natural projection $P \to P_{\pi}$. A similar sort of argument, which we leave to the reader, shows that $\Supp \ker g / \im f = \varnothing$ and consequently Lemma \ref{emptysupp} gives us exactness.
\end{proof}
\section{The proof of Theorem \ref{type} and a useful lemma}
\label{heckeproofs}
Let us review induction functors and their adjunction relations. Fix an $\ell$-group $G$ and a closed subgroup $H \subset G$. Then the functor $\Res_H^G : \mathcal{R}(G) \to \mathcal{R}(H)$ which restricts representations to $H$ has a right adjoint described as follows. Choose $(\pi,V) \in \mathcal{R}(H)$, and let \[ I(V) = \{ f : G \to V \ | \ f(hg) = \pi(h)(f(g)) \ \text{for all} \ h \in H,g \in G \} \] with a $G$-action defined by $(g \cdot f)(x) = f(xg)$. Then the induced representation $\Ind_H^G \pi$ is defined to be the smooth part of $I(V)$, and after extending to morphisms in the obvious way we call the resulting functor $\Ind_H^G : \mathcal{R}(H) \to \mathcal{R}(G)$ \emph{smooth induction}. Also, we define $\cInd_H^G \pi$ to be the subrepresentation of $\Ind_H^G \pi$ consisting of those functions with compact support modulo $H$, and call $\cInd_H^G : \mathcal{R}(H) \to \mathcal{R}(G)$ \emph{smooth induction with compact supports}, a subfunctor of $\Ind_H^G$. In case $H$ is open, $\cInd_H^G$ is left adjoint to $\Res_H^G$.
\begin{proof}[Proof of Theorem \ref{type}]
By Proposition 3.3 in \cite{BK}, it suffices to prove that $\mathcal{R}_{\rho}(G)$ is closed under taking subquotients. But in fact this full subcategory of $\mathcal{R}(G)$ coincides with $\mathcal{R}(G)_{\mathscr{V}(G,\rho)}$, which is clearly closed under taking subquotients.
To see why, suppose $M \in \mathcal{R}_{\rho}(G)$ and consider the inclusion $e_{\rho} \cdot M \to M|_{K}$ as a $K$-morphism. Then there is a corresponding $G$-morphism $\cInd_K^G e_{\rho} \cdot M \to M$, which is surjective because $M$ is generated by $e_{\rho} \cdot M$, so that $\Supp M \subset \Supp (\cInd_K^G e_{\rho} \cdot M)$. But clearly $\cInd_K^G e_{\rho} \cdot M$ is supported on $\mathscr{V}(K,\rho)$, because for any $\pi \in \widehat{G}$ we have \[ \Hom_G(\cInd_K^G e_{\rho} \cdot M,\pi) \cong \Hom_K(e_{\rho} \cdot M,\pi|_K) \] and if $\pi \notin \mathscr{V}(K,\rho)$ then the latter space is zero.
Conversely, let $M \in \mathcal{R}(G)$ be supported on $\mathscr{V}(K,\rho)$, and write $N = \mathcal{H}(G) * e_{\rho} \cdot M$ for the $G$-subrepresentation of $M$ generated by the $\rho$-isotypic component $e_{\rho} \cdot M$. Obviously $\Supp M/N \subset \Supp M \subset \mathscr{V}(K,\rho)$, and also $e_{\rho} \cdot M \subset N$ implies $\Supp(M/N) \cap \mathscr{V}(K,\rho) = \varnothing$. Thus $\Supp M/N = \varnothing$, so by Lemma \ref{emptysupp} we have $M = N$ as desired.
\end{proof}
Here we include a technical lemma, which will in fact be used in all the proofs that follow. But first it is useful to introduce an auxiliary category, which we denote by $\check{G}$, in order to clean up the statement of the lemma. An object of $\check{G}$ is a pair $(K,\rho)$ consisting of a compact open subgroup $K \subset G$ and an irreducible smooth representation $\rho$ of $K$, and there are no morphisms $(L,\sigma) \to (K,\rho)$ unless $K \subset L$. If $K \subset L$, then a morphism $(L,\sigma) \to (K,\rho)$ in $\check{G}$ is a nonzero (hence surjective) $K$-morphism $\sigma|_K \to \rho$. As the notation suggests, the category $\check{G}$ is closely related to the topological space $\widehat{G}$.
Now each $M \in \mathcal{R}(G)$ gives rise to a presheaf $\mathscr{M}$ on $\check{G}$ with values in complex vector spaces. Define \[ \mathscr{M}(K,\rho) = \Hom_K(\rho,M|_K), \] and we obtain the ``restriction" map associated with a morphism $(L,\sigma) \to (K,\rho)$ as follows. There is a canonical isomorphism \[ \mathscr{M}(K,\rho) = \Hom_K(\rho,M|_K) \to \Hom_L(\cInd_K^L \rho,M|_L), \] and the chosen $K$-morphism $\sigma|_K \to \rho$ corresponds to an $L$-morphism \[ \sigma \to \Ind_K^L \rho = \cInd_K^L \rho, \] which induces a pullback map \[ \Hom_L(\cInd_K^L \rho,M|_L) \to \Hom_L(\sigma,M|_L) = \mathscr{M}(L,\sigma). \] It is not so hard to see that these ``restrictions" are surjective. Now we state and prove the aforementioned lemma, the proof of which uses some ideas from \cite{GK}.
\begin{lemma}
Let $M \in \mathcal{R}(G)$ and suppose $\pi \in \widehat{G}$ but $\pi \notin \Supp M$. Then if $(K,\rho) \in \widehat{G}$ with $\pi \in \mathscr{V}(K,\rho)$, for any $\alpha \in \mathscr{M}(K,\rho)$ there exists $(L,\sigma) \in \check{G}$ and a morphism $(L,\sigma) \to (K,\rho)$ such that $\pi \in \mathscr{V}(L,\sigma)$ and $\alpha$ is annihilated by the induced map $\mathscr{M}(K,\rho) \to \mathscr{M}(L,\sigma)$. If $(K,\rho)$ is chosen so that $| \rho : \pi|_K | = 1$, then $(L,\sigma)$ can be chosen so that $| \sigma : \pi|_L | = 1$ and the morphism $(L,\sigma) \to (K,\rho)$ is unique modulo nonzero scalars.
\label{shrink}
\end{lemma}
\begin{proof}
Let us write $A = \mathcal{H}(G,\rho)$, so $\widetilde{M} = \mathscr{M}(K,\rho)$ is naturally an $A$-module, as is $\widetilde{\pi} = \Hom_K(\rho,\pi|_K)$, and the latter is finite-dimensional and simple (in particular, nonzero). In view of Theorem \ref{type} it is clear that $\Hom_A(\widetilde{M},\widetilde{\pi}) = 0$, and in particular $\Hom_A(A \cdot \alpha,\widetilde{\pi}) = 0$ since $\widetilde{\pi}$ is injective in $A - \mathbf{mod}$. If we put $I = \Ann_A(\widetilde{\pi})$ and $J = \Ann_A(\alpha)$, this means that $I + J = A$, and in particular we can find $x \in I$ and $y \in J$ such that $x + y = 1$. Now by the $\ell_c$ condition $A$ is a filtered union of subalgebras $\mathcal{H}(L,\rho)$, where $L$ is a compact open subgroup containing $K$, whence it is possible to find $L \supset K$ large enough that $x,y \in \mathcal{H}(L,\rho)$ and $\widetilde{\pi}$ is a simple $\mathcal{H}(L,\rho)$-module. Since the space \[ \Hom_L(\cInd_K^L \rho,\pi|_L) = \Hom_K(\rho,\pi|_K) \] is nonzero by assumption, there is some $\sigma \in \widehat{L}$ such that $| \sigma : \cInd_K^L \rho |, | \sigma : \pi|_L | > 0$. This means we can also find some nonzero $L$-morphism $\sigma \to \cInd_K^L \rho$, which corresponds to the needed $K$-morphism $\sigma|_K \to \rho$. If we chose $\rho$ so that $| \rho : \pi|_K | = 1$, then now we have a unique choice of $\sigma \in \widehat{L}$ with $| \sigma : \cInd_K^L \rho |, | \sigma : \pi|_L | = 1$, and the morphism $\sigma|_K \to \rho$ is also unique modulo nonzero scalars.
Now we show that the morphism $(L,\sigma) \to (K,\rho)$ we have found has the property that $\alpha$ is annihilated by the ``restriction" $\mathscr{M}(K,\rho) \to \mathscr{M}(L,\sigma)$. Let us write $B = \mathcal{H}(L,\rho)$ and observe that $I \cap B + J \cap B = B$ (because $x+y=1$), which says that $\Hom_B(B \cdot \alpha, \widetilde{\pi}) = 0$. To say that $\alpha$ vanishes in the restriction is to say that the composite $L$-morphism \[ \sigma \to \cInd_K^L \rho \to M|_L \] is zero, so let us assume otherwise and find a contradiction. In that case we can certainly find an $L$-morphism $\varphi : M|_L \to \pi|_L$ such that the composition \[ \sigma \to \cInd_K^L \rho \to M|_L \to \pi|_L \] is nonzero. But then this induces a $B$-morphism $\varphi_* : \widetilde{M} \to \widetilde{\pi}$ with $\varphi_*(\alpha) \neq 0$, and we just saw that $\Hom_B(B \cdot \alpha, \widetilde{\pi}) = 0$.
\end{proof}
\begin{corollary}
In the situation of the lemma, if we assume also that $M$ is admissible, it is possible to find $(L,\sigma) \to (K,\rho)$ so that $\mathscr{M}(L,\sigma) = 0$.
\end{corollary}
\begin{proof}
If $M$ is admissible then $\dim_{\C} \mathscr{M}(K,\rho) < \infty$, so the corollary follows from the lemma in view of the fact that the ``restriction" maps $\mathscr{M}(K,\rho) \to \mathscr{M}(L,\sigma)$ are surjective.
\end{proof}
\section{The proof of Theorem \ref{fell2}}
\label{adjointproofs}
Next we will prove Theorem \ref{fell2}, but first let us recall some notation. Let $Z \subset W \subset \widehat{G}$ be any subsets and $U = W \setminus Z$, and write $i_! : \mathcal{R}(G)_Z \to \mathcal{R}(G)_W$ and $j_! : \mathcal{R}(G)_U \to \mathcal{R}(G)_W$ for the inclusions. Now $j_!$ always has a right adjoint $j^!$ for formal reasons, given by the formula \[ j^!M = \bigcap_{\substack{f \in \Hom_G(M,\pi) \\ \pi \in Z}} \ker f \] on objects $M \in \mathcal{R}(G)_W$. It is easy to see that this construction is functorial. We can define $i^*M = M/j^!M$ or more formally $M/j_!j^!M$, and it is clear enough that if $i_!$ has a left adjoint it must be given by this formula. The subtle issue is whether $i^*M$ is always supported on $Z$, and the theorem says that as long as $W$ is locally closed, this happens precisely when $Z$ is closed in $W$.
\begin{proof}[Proof of Theorem \ref{fell2}]
First suppose that $Z$ is closed in $W$ and $M \in \mathcal{R}(G)_W$. To see that $\Supp i^*M \subset Z$, fix $\pi \in \Supp i^*M$ and $(K,\rho) \in \check{G}$ such that $\pi \in \mathscr{V}(K,\rho)$, so it suffices to prove that $Z \cap \mathscr{V}(K,\rho) \neq \varnothing$ because then $\pi \in \overline{Z} = Z$. These assumptions give us nonzero (hence surjective) morphisms $i^*M \to \pi$ and $\pi|_K \to \rho$, which we can compose to get a nonzero morphism $i^*M|_K \to \pi|_K \to \rho$. Then by semisimplicity of $\mathcal{R}(K)$ we can find a nonzero morphism $\rho \to M|_K$ which does not land in $j^!M$, so by our definition of the latter there exists $\zeta \in Z$ and a morphism $M \to \zeta$ such that the composition $\rho \to M|_K \to \zeta|_K$ is nonzero. That is, $\zeta \in Z \cap \mathscr{V}(K,\rho)$.
The rest is more formal. For all $M \in \mathcal{R}(G)_W$ and $N \in \mathcal{R}(G)_Z$ we have a pullback map \[ \Hom_G(i^*M,N) \to \Hom_G(M,N), \] which we need to show is an isomorphism (the naturality requirement is clear). This map is injective because $M \to i^*M$ is surjective, so for the surjectivity fix a morphism $M \to N$. By Corollary \ref{zeromap}, to see that this morphism descends to the quotient it is enough to check that for any $\pi \in W$, the composition \[ j^!M \to M \to N \to \pi \] is zero. If $\pi \notin Z$, then this is just because $\Supp N \subset Z$, and if $\pi \in Z$ then use the definition of $j^!M$.
For the other implication, we can immediately reduce to the case that $W$ is open in $\widehat{G}$: let us assume for the moment that we have proved the theorem in this situation. Find open $V \supset W$ such that $W$ is closed in $V$, so by the first implication the inclusion $\mathcal{R}(G)_W \to \mathcal{R}(G)_V$ has a left adjoint. But then if $\mathcal{R}(G)_Z \to \mathcal{R}(G)_W$ has a left adjoint we can compose the two to get a left adjoint for $\mathcal{R}(G)_Z \to \mathcal{R}(G)_V$, and the special case implies that $Z$ is closed in $V$, hence closed in $W$.
So assume $W$ is open. If $i_!$ has a left adjoint, it is easy to check that it must be isomorphic to $i^*$, and in particular $i^*M$ is supported on $Z$ for every $M \in \mathcal{R}(G)_W$. Take $M \subset \mathcal{H}(G)$ to be the ``restriction" of the left regular representation to $W$: explicitly, if $k : W \to \widehat{G}$ is the open inclusion then $M = k^!\mathcal{H}(G)$, where $\mathcal{H}(G)$ is given the $G$-action by translations. Clearly $\Supp M = W$, so it makes sense to consider $i^*M$, and with our assumption we get $\Supp i^*M = Z$. In particular, if $\pi \in W \setminus Z$ then $\Hom_G(M,\pi) = 0$.
Find $(K,\rho) \in \check{G}$ such that $\pi \in \mathscr{V}(K,\rho) \subset W$, and as before put $A = \mathcal{H}(K,\rho)$. If we restrict $M$ further to $\mathscr{V}(K,\rho)$, then, under the equivalence of Theorem \ref{type}, $M$ corresponds to $\mathscr{M}(K,\rho) = A$ viewed as an $A$-module. Thus we can apply Lemma \ref{shrink} to find $(L,\sigma) \in \check{G}$ and a morphism $(L,\sigma) \to (K,\rho)$ such that $\pi \in \mathscr{V}(L,\sigma) \subset \mathscr{V}(K,\rho)$ and $1 \in A$ is annihilated by the restriction map $\mathscr{M}(K,\rho) \to \mathscr{M}(L,\sigma)$. But then this $A$-map is zero, and we know it to be surjective also, whence $\mathscr{M}(L,\sigma) = 0$ or equivalently $\mathscr{V}(L,\sigma) \cap Z = \varnothing$ and $Z$ is closed as desired.
Now for the exactness, notice that when $i^*$ exists it is right exact simply because it is a left adjoint, so it suffices to show that $i^*$ preserves monomorphisms. Let $f : M \to N$ be an injective morphism in $\mathcal{R}(G)_W$, write $P \subset i^*M$ for the kernel of $i^*f$, and fix a morphism $g : P \to \pi$ for some $\pi \in W$. Then by injectivity we can lift $g$ to $i^*M \to \pi$, so if $\pi \notin Z$ then already this morphism is zero. Otherwise this naturally induces a morphism $M \to \pi$, and then use injectivity again to lift to $N \to \pi$. Since $\pi \in Z$ this descends to $\widetilde{g} : i^*N \to \pi$, whence $g(P) = \widetilde{g}(i^*f(P)) = 0$ and $P = 0$ by Lemma \ref{emptysupp}.
\end{proof}
We include a proof of Corollary \ref{projective} here.
\begin{proof}[Proof of Corollary \ref{projective}]
To see that an open point $\pi \in \widehat{G}$ is projective, write $i : \{ \pi \} \to \widehat{G}$ for the inclusion and observe that for any $M \in \mathcal{R}(G)$, the composition \[ i^!M \to M \to i^*M \] is an isomorphism because $\pi$ is open and closed. Notice $i^*M = M_{\pi}$ with the notation from section \ref{stalkstatements}, so \[ \Hom_G(\pi,M) = \Hom_G(\pi,i^!M) = \Hom_G(\pi,M_{\pi}) = M(\pi) \] and $M \to \Hom_G(\pi,M)$ is just the functor of taking stalks at $\pi$. But this functor is exact by Lemma \ref{exact}.
Now for the converse, in view of Theorem \ref{fell2} it suffices to prove that if $j : \widehat{G} \setminus \{ \pi \} \to \widehat{G}$, the inclusion $j_! : \mathcal{R}(G)_{\widehat{G} \setminus \{ \pi \}} \to \widehat{G}$ has a left adjoint. Moreover, by the proof of the theorem it is enough to check that $j^*M$ is supported on $\widehat{G} \setminus \{ \pi \}$ for all $M \in \mathcal{R}(G)$. Consider the short exact sequence \[ 0 \to \Hom_G(\pi,i^!M) \to \Hom_G(\pi,M) \to \Hom_G(\pi,j^*M) \to 0 \] (so here we are using the projectivity of $\pi$). But $\Hom_G(\pi,i^!M) \to \Hom_G(\pi,M)$ is obviously an isomorphism, which forces $\Hom_G(\pi,j^*M) = 0$. Fix a morphism $j^*M \to \pi$, which induces a map \[ \Hom_G(\pi,j^*M) \to \Hom_G(\pi,\pi) \cong \C, \] and this is obviously not surjective since the first space is zero. So $j^*M \to \pi$ is not surjective, hence zero, and $\Supp j^*M \subset \widehat{G} \setminus \{ \pi \}$ as desired.
\end{proof}
\section{The proof of Theorem \ref{fell3}}
Now we are ready to prove Theorem \ref{fell3}, which we will deduce from Theorem \ref{fell2}.
\begin{proof}[Proof of Theorem \ref{fell3}]
We start by proving that the bijection $\mathscr{V}(K,\rho) \to \widehat{A}$ is closed. Suppose $Z \subset \mathscr{V}(K,\rho)$ is closed, so to prove that its image in $\widehat{A}$ is closed we must show that if $I \in \widehat{A}$ contains \[ J_Z = \bigcap_{\zeta \in Z} \Ann_{A}(\widetilde{\zeta}) \] (here again we write $\widetilde{\zeta} = \Hom_K(\rho,\zeta|_K)$), then $\pi \in Z$ where $I = \Ann_{A}(\widetilde{\pi})$. If we take $W = \mathscr{V}(K,\rho)$ in the setup of Theorem \ref{fell2} and let $M \in \mathcal{R}(G)_W$ be such that $\widetilde{M} = A$ as an $A$-module, then $\widetilde{i^*M} = A/J_Z$ and in particular $A/J_Z$ is supported on the image of $Z$ in $\widehat{A}$, in the sense that $\Hom_A(A/J_Z,\widetilde{\zeta}) = 0$ for any $\zeta \in W \setminus Z$. The assumption $I \supset J_Z$ says that there is a canonical surjection $A/J_Z \to \widetilde{\pi}$, whence $\pi \in Z$.
Conversely, suppose $Z \subset \widehat{A}$ is closed in the Jacobson topology and let $J_Z$ be as before. We want to use Theorem \ref{fell2} again, so $W = \mathscr{V}(K,\rho)$ still, and it is enough to show that the inclusion $\mathcal{R}(G)_Z \to \mathcal{R}(G)_W$ has a left adjoint. Recall from the proof of Theorem \ref{fell2} that actually it suffices just to check that $i^*M$ is supported on $Z$ for all $M \in \mathcal{R}(G)_W$, so for this observe that $\widetilde{i^*M} = \widetilde{M}/J_Z\widetilde{M}$. But since $Z$ is closed in $\widehat{A}$, we have $\Hom_A(A/J_Z,\widetilde{\zeta}) = 0$ for $\zeta \in W \setminus Z$, which implies that $\widetilde{M}/J_Z\widetilde{M}$ and hence $i^*M$ is supported on $Z$.
\end{proof}
\section{The proof of Theorem \ref{bernstein}}
Finally, we come to Theorem \ref{bernstein}, which describes the Bernstein center of $\mathcal{R}(G)$. We will need the exactness of induction, which is proved, for example, in \cite{BZ}.
\begin{proposition}
For any $\ell$-group $G$ and closed subgroup $H$, the functors $\cInd_H^G,\Ind_H^G : \mathcal{R}(H) \to \mathcal{R}(G)$ are exact.
\end{proposition}
\begin{proof}[Proof of Theorem \ref{bernstein}]
First we show that the image of the homomorphism $\mathscr{Z}_W(G) \to\Fun(W)$ contains $C^{\infty}(W)$. Fix $f \in C^{\infty}(W)$ and decompose $W = \coprod_{i \in I} U_i$ into the level sets of $f$, which are open and closed in $W$. Given $M \in \mathcal{R}(G)_W$, let us write $M_{U_i}$ for the largest quotient of $M$ supported on $U_i$ obtained from Theorem \ref{fell2}, so writing $i_{U_i} : U_i \to W$ for the inclusions this means $M_{U_i} = i^*_{U_i}M$. In fact, since each $U_i$ is also open, the composition \[ i_{U_i}^!M \to M \to i_{U_i}^*M \] is an isomorphism and $M_{U_i}$ may be viewed as a subrepresentation in a canonical way. We claim that the natural morphism \[ \bigoplus_{i \in I} M_{U_i} \longrightarrow M \] is in fact an isomorphism, which we prove by using Lemma \ref{exact} to test the morphism on stalks. Fix $\pi \in W$, and recall from the proof of that lemma that $M(\pi)^* = \Hom_G(M,\pi)$, which is easier to work with. Therefore we need only observe that the induced map \[ \Hom_G(M,\pi) \longrightarrow \Hom_G \big( \bigoplus_{i \in I} M_{U_i},\pi \big) \cong \prod_{i \in I} \Hom_G(M_{U_i},\pi) \] is an isomorphism, so use Theorem \ref{fell2} and the fact that $\pi \in M_{U_i}$ for exactly one $i \in I$.
Now we define $\varphi_M \in \End_G(M)$ by letting $\varphi_M$ scale each summand $M_{U_i}$ by $f(U_i)$. To see that this is natural in $M$, choose a morphism $g : M \to N$ in $\mathcal{R}(G)$ and observe that by the proof of Theorem \ref{fell2}, $g$ sends each $M_{U_i}$ into $N_{U_i}$. Then the square we want to commute decomposes into a direct sum, and each summand commutes by $\C$-linearity of $g|_{M_{U_i}}$. Now since $\varphi_{\pi}$ is the scalar $f(\pi)$ it is easy to see that the homomorphism $\mathscr{Z}_W(G) \to \Fun(W)$ sends $\varphi$ to $f$ as needed.
As for the converse, choose $\theta \in \mathscr{Z}_W(G)$, which is a functorial family of endomorphisms $\theta_M : M \to M$ for all $M \in \mathcal{R}(G)_W$. Fix $\pi \in W$, so we need to find an open neighborhood of $\pi$ in $W$ such that $\theta$ acts on all points of this neighborhood by the same scalar. To this end we can assume without loss of generality that $\theta_{\pi} = 0$, and let us also choose $(K,\rho) \in \check{G}$ such that $| \rho : \pi|_K | = 1$ and $Z = \mathscr{V}(K,\rho) \cap W$ is closed in $\mathscr{V}(K,\rho)$. Let us write $i_! : \mathcal{R}(G)_Z \to \mathcal{R}_{\rho}(G)$ for the inclusion functor, which has an exact left adjoint $i^*$ by Theorem \ref{fell2}, and put $N = \cInd_K^G \rho \in \mathcal{R}_{\rho}(G)$. There is a unique (up to scaling) morphism $N \to \pi$, which factors through $i^*N$ because $\pi \in Z$, so denote by $M$ the kernel of the surjection $i^*N \to \pi$. Notice that $\theta_{\pi} = 0$ implies that $\theta_{i^*N}$ lands in $M$, and clearly $\Hom_G(M,\pi) = 0$. Thus we obtain a nonzero morphism $\alpha : \rho \to M|_K$ corresponding to the composition $N \to i^*N \to \pi$, so we can apply Lemma \ref{shrink} to find $(L,\sigma) \in \check{G}$ and a morphism $(L,\sigma) \to (K,\rho)$ such that $| \sigma : \pi|_K | = 1$, $\mathscr{V}(L,\sigma) \subset \mathscr{V}(K,\rho)$, and $\alpha$ is annihilated in the restriction $\mathscr{M}(K,\rho) \to \mathscr{M}(L,\sigma)$.
We claim that $\theta_{\zeta} = 0$ for any $\zeta \in \mathscr{V}(L,\sigma) \cap W$, so clearly it suffices to prove that for $P = \cInd_L^G \sigma$ we have $\theta_{i^*P} = 0$. To see the latter, consider the chosen injection $\sigma \to \cInd_K^L \rho$, which induces an injective $G$-morphism $\cInd_L^G \sigma \to \cInd_K^G \rho$ by exactness of $\cInd_L^G$, then look at the commutative square \[
\begin{diagram}
i^*P & \rTo^{\theta} & i^*P \\
\dTo & & \dTo \\
i^*N & \rTo^{\theta} & i^*N
\end{diagram}. \]
Then recall that $(L,\sigma) \to (K,\rho)$ was chosen to make the composition $i^*P \to i^*N \stackrel{\theta}{\to} i^*N$ zero, so from the injectivity of $i^*P \to i^*N$ the claim follows.
\end{proof}
|
\section{Introduction}
Event-by-event fluctuations of chemical (particle-type) composition of hadronic final states
of relativistic heavy ion collisions are expected to be sensitive to properties of strongly
interacting matter produced in the collisions~\cite{Koch:2008ia}.
Specific fluctuations can signal the onset of deconfinement when the collision
energy becomes sufficiently high to create droplets of
quark-gluon plasma~\cite{Gorenstein:2003hk}.
At higher collision energies, where the quark-gluon phase
is abundantly produced at the early collision stage, large chemical fluctuations can
occur as the system hits the critical point
of strongly interacting matter in the course
of its temporal evolution~\cite{Koch:2005vg,Koch:2005pk}.
It is thus certainly of interest to study event-by-event
chemical fluctuations experimentally. First data coming from the
CERN SPS~\cite{Afanasev:2000fu,Alt:2008ca,Kresan:2009qs} and BNL RHIC~\cite{Abelev:2009if}
were already published. The results are not very conclusive yet and more systematic
measurements are needed. In addition the question arises whether data analysis methods
can be improved.
In real experiments it is impossible to determine uniquely the type of every detected
particle. The identification requires measurements of particle electric charge and
mass. Precise mass measurements are experimentally difficult and expensive.
For this reason analyzes of chemical fluctuations are usually performed in a limited
acceptance where particle identification is relatively reliable.
However, sensitivity to fluctuations
of range larger than the acceptance window is then lost
and signals from fluctuations of shorter range are usually diluted. Furthermore
it should be noted
that results on fluctuations, unlike those on single particle spectra, cannot be corrected
for the limited acceptance. Often it is possible to enlarge the acceptance, but
at the expense of a significant contamination of the sample by misidentified particles.
The effect of particle misidentification can distort measured fluctuations. Thus,
incomplete particle identification is a serious obstacle to the precise
measurement of chemical fluctuations.
Although it is usually impossible to identify every detected particle, one can in general
determine with high accuracy the percentage (averaged over many interactions) of,
say, kaons among produced hadrons. This information will be shown to be sufficient to fully
eliminate the effect of incomplete identification.
In this paper we propose a new experimental
technique called the {\em identity method} which achieves the goal independently
of the specific properties of the chemical fluctuations under study.
In the study of chemical fluctuations
the NA49 Collaboration~\cite{Afanasev:2000fu,Alt:2008ca,Kresan:2009qs}
used the measure $\sigma_{\rm dyn}$,
which is defined as the difference between fluctuations measured in real
and {\it mixed} events.
The effect of particle misidentification is accounted for
by including it in the mixed events.
The STAR Collaboration~\cite{Abelev:2009if} used, in addition to the $\sigma_{\rm dyn}$
measure, the quantity $\nu_{\rm dyn}$. The latter one assumes that particles are
uniquely identified.
It was suggested long ago~\cite{Gazdzicki:1997gm,Mrowczynski:1999sf} to quantify
chemical fluctuations by the measure $\Phi$~\cite{Gazdzicki:1992ri} which proved
to be efficient in experimental studies of event-by-event fluctuations of particle transverse
momentum~\cite{Anticic:2003fd,Anticic:2008vb}, electric charge~\cite{Alt:2004ir}, and
quite recently of azimuthal angle~\cite{Cetner:2010vz}. The $\Phi$ measure, unlike
$\sigma_{\rm dyn}$ and $\nu_{\rm dyn}$, is a strongly intensive measure of fluctuations.
Namely, its magnitude is independent of the number and of the distribution (fluctuation) of the number
of particle sources, if the sources are identical and independent from each other. This
feature, which is discussed in detail in \cite{Gorenstein:2011vq}, is important in experimental
studies of relativistic heavy-ion collisions where the collision centrality is never fully
controllable. However, up to now it was unclear how to correct measurements of
chemical $\Phi$ for the effect of particle misidentification.
The identity method, which is developed here, uses the fluctuation measure $\Psi$,
a simple modification of $\Phi$. The measure $\Psi$, similarly to $\Phi$, is
strongly intensive but the modification allows to correct the measurements for
the effect of particle misidentification. Below we show that the measure $\Psi$
can be factorized into a coefficient, which represents the effect of misidentification,
and the quantity $\Psi_{\rm CI}$ which corresponds to the value $\Psi$
would have for complete identification.
The misidentification coefficient can be determined from the data in a model
independent way.
Therefore, the identity method provides the value of the fluctuation measure as
it would be obtained in an experiment in which
every particle is uniquely identified.
Before the identity method is presented we introduce and discuss in
Sec.~\ref{sec-measures} the fluctuation measures $\sigma_{\rm dyn}$, $\nu_{\rm dyn}$,
$\Phi$, and $\Psi$. In Sec.~\ref{sec-misidentification} we demonstrate
by a Monte Carlo simulation how the effect of misidentification distorts the chemical
fluctuations as quantified by $\nu_{\rm dyn}$, $\Phi$ and $\Psi$. The identity method
is formulated in Sec.~\ref{sec-id-method}. Instead of conclusions we present in the
last section the steps required to apply the identity method to experimental data.
In order to simplify the presentation, we consider chemical fluctuations of events
composed only of kaons and pions. Clearly, kaons and pions can be replaced
by particles of any other sort.
\section{Measures of fluctuations}
\label{sec-measures}
As mentioned in the Introduction, fluctuations of chemical composition of final
states of relativistic heavy-ion collisions can be studied in several ways.
The NA49 Collaboration~\cite{Afanasev:2000fu,Alt:2008ca,Kresan:2009qs}
measured event-by-event fluctuations of the particle ratios $K/\pi, \; K/p, \; p/\pi$
and determined the quantity $\sigma_{\rm dyn}$ defined as
\be
\sigma_{\rm dyn} = {\rm sgn}(\sigma^2_{\rm data} - \sigma^2_{\rm mixed})
\sqrt{|\sigma^2_{\rm data} - \sigma^2_{\rm mixed}|}~,
\ee
where $\sigma_{\rm data}$ and $\sigma_{\rm mixed}$ is the relative width
(the width divided by the mean) of the event-by-event particle ratio distribution
in, respectively, the
data and artificially generated mixed events where every particle comes from
a different real event. The fluctuations present in mixed events are due to
the effect of particle misidentification and the
statistical noise caused by the finite number of particles.
The STAR Collaboration used~\cite{Abelev:2009if} the quantity
$\nu_{\rm dyn}$ to measure chemical fluctuations. For the
case of a two-component system of pions and kaons $\nu_{\rm dyn}$ is defined as
\be
\label{nu-dyn}
\nu_{\rm dyn} = \frac {\langle N_K (N_K-1) \rangle}
{\langle N_K \rangle ^2} + \frac {\langle N_\pi (N_\pi-1) \rangle}
{\langle N_\pi \rangle ^2} - 2 \frac{\langle N_K N_\pi \rangle}
{\langle N_K \rangle \langle N_\pi \rangle}~,
\ee
where $N_K$ and $N_\pi$ are the numbers of kaons and pions in a given event
and $\langle ... \rangle $ denotes averaging over events.
The quantity $\nu_{\rm dyn}$
is defined in such a way that, in particular,
it vanishes when the multiplicity distributions of pions
and kaons are both poissonian ($\langle N_i (N_i -1) \rangle =
\langle N_i \rangle^2$, $i=\pi, \, K$) and independent from each other
($\langle N_K N_\pi \rangle = \langle N_K \rangle \langle N_\pi \rangle$). Thus,
it is constructed to quantify the deviations of the fluctuations from the poissonian noise.
For large enough particle multiplicities one finds the approximate relation
$\nu_{\rm dyn} \approx \sigma^2_{\rm data} - \sigma^2_{\rm mixed}$
which gives $\nu_{\rm dyn} \approx {\rm sgn} (\sigma_{\rm dyn}) \: \sigma^2_{\rm dyn} $
\cite{Abelev:2009if}. We note that the quantity $\nu_{\rm dyn}$ implicitly assumes
unique identification of all particles.
As already noted, it was advocated long ago~\cite{Gazdzicki:1997gm,Mrowczynski:1999sf}
to employ the measure $\Phi$~\cite{Gazdzicki:1992ri} to study chemical fluctuations. The
measure is defined in the following way. One introduces the variable $z \equiv x - \overline{x}$,
where $x$ is a single particle characteristic such as the transverse momentum
or azimuthal angle. The over-line denotes averaging over the single particle inclusive
distribution. The event variable $Z$, which is a multi-particle analog of $z$, is defined as
$Z \equiv \sum_{i=1}^{N}(x_i - \overline{x})$, where the sum runs over the $N$ particles in
a given event. By construction, $\langle Z \rangle = 0$. The measure $\Phi$ is finally
defined as
\be
\label{phi-def}
\Phi \equiv \sqrt{\langle Z^2 \rangle \over \langle N \rangle} -
\sqrt{\overline{z^2}}~.
\ee
The measure
$\Phi$ vanishes in the absence of inter-particle correlations. This situation
is discussed in some detail below for the case of chemical fluctuations.
Here we note that the measure also possesses another important
property - it is strongly intensive which means that it is independent of the
number and of the distribution (fluctuation) of the number of particle sources, if the sources
are identical and independent from each other. In particular, if a nucleus-nucleus
collision is a simple superposition of nucleon-nucleon interactions, then
$\Phi_{AA} = \Phi_{NN}$. The strongly intensive property is a very valuable feature of
$\Phi$, as centrality selection in relativistic heavy-ion collisions is never
perfect and events of different numbers of particle sources are always
mixed up. The strongly intensive property is also desirable when different centralities
or different colliding systems are compared. For a discussion of strongly
intensive quantities see~\cite{Gorenstein:2011vq}.
The analysis of chemical fluctuations
can be performed with the help of
$\Phi$ in two different but fully equivalent ways.
In the first method~\cite{Gazdzicki:1997gm}, using the identity variable,
chemical fluctuations are treated in analogy to fluctuations
of transverse momentum. In the second method $\Phi$ is calculated from
the moments of the multiplicity distributions~\cite{Mrowczynski:1999sf}.
We next describe the first method for the example
of a two-component system of pions and kaons.
One defines the single particle variable $x$ as $x = w_K$, where $w_K$ is called
the {\it kaon identity} and
$w_K^i=1$ if the $i-$th particle is a kaon and $w_K^i =0$ if the $i-$th
particle is a pion. This implies unique particle identification.
One then directly uses the definition (\ref{phi-def}) to evaluate $\Phi$.
Let us now discuss the most important case for
which the measure $\Phi$ of chemical fluctuations
vanishes. Since the inclusive distribution of $w_K$ equals
\be
P(w_K) = \left\{ \begin{array}{ccl}
\frac{\langle N_\pi \rangle}{\langle N \rangle} & {\rm for} & w_K=0~, \\[2mm]
\frac{\langle N_K \rangle}{\langle N \rangle} & {\rm for} & w_K=1~,
\end{array} \right.
\ee
one finds
\be
\overline{z^2} \equiv \overline{w_K^2} - \overline{w_K}^2
= \frac{\langle N_K \rangle}{\langle N \rangle}
\bigg(1 - \frac{\langle N_K \rangle}{\langle N \rangle} \bigg)~.
\ee
When inter-particle correlations are absent, the distribution of particle
identities in events of multiplicity $N$ reads
\be
\label{dist-N-w}
P_N (w_K^1,w_K^2, \dots, w_K^N) = {\cal P}_N
P(w_K^1) \, P(w_K^2) \cdots P(w_K^N)~,
\ee
where ${\cal P}_N$ is an arbitrary multiplicity distribution of particles of any
type. One shows that $\langle Z^2 \rangle$ computed with the event distribution
(\ref{dist-N-w}) equals $\langle Z^2 \rangle = \langle N \rangle \overline{z^2}$
and consequently, $\Phi = 0$.
In the second method the measure $\Phi$ of chemical
fluctuations is obtained from the moments of the experimentally measured
multiplicity distributions of kaons and pions.
As shown in Ref.~\cite{Mrowczynski:1999sf}, one has
\ba
\label{z2bar}
\overline{z^2}
&=& { \langle N_K \rangle \langle N_\pi \rangle \over
\langle N \rangle^2 }~,
\\ \nonumber
{\langle Z^2 \rangle \over \langle N \rangle}
&=&
{\langle N_\pi \rangle^2 \over \langle N \rangle^3}
\big(\langle N_K^2 \rangle - \langle N_K \rangle^2 \big)
+{\langle N_K \rangle^2 \over \langle N \rangle^3}
\big(\langle N_\pi^2 \rangle - \langle N_\pi \rangle^2 \big)
\\ \label{Z2/N}
&&- 2 \,{\langle N_K \rangle \langle N_\pi \rangle \over \langle N \rangle^3}
\big(\langle N_K N_\pi \rangle - \langle N_K \rangle \langle N_\pi \rangle\big)~,
\ea
which substituted in Eq.~(\ref{phi-def}) give the measure $\Phi$.
The formulas (\ref{z2bar}, \ref{Z2/N}) clearly show that $\Phi$, like $\nu_{\rm dyn}$,
vanishes when the multiplicity distributions of pions and kaons are both poissonian
and independent from each other. However, more generally, $\Phi$ vanishes for any multiplicity distribution
provided it satisfies (\ref{dist-N-w}). The distribution (\ref{dist-N-w}) leads to the
multiplicity distribution of the form
\ba
{\cal P}_{N_K N_\pi} &=& {\cal P}_{N_K + N_\pi}
\\ [2mm] \nonumber
&\times& {N_K + N_\pi \choose N_K}
\bigg(\frac{\langle N_K \rangle}{\langle N \rangle} \bigg)^{N_K}
\bigg(1 - \frac{\langle N_K \rangle}{\langle N \rangle} \bigg)^{N_\pi}~,
\ea
with the moments
\ba
\langle N_K (N_K -1) \rangle = \frac{\langle N_K \rangle^2}{\langle N \rangle^2}
\langle N (N -1) \rangle~,
\\ [2mm]
\langle N_\pi (N_\pi -1) \rangle = \frac{\langle N_\pi \rangle^2}{\langle N \rangle^2}
\langle N (N -1) \rangle~,
\\ [2mm]
\langle N_K N_\pi \rangle
= \frac{\langle N_K \rangle \langle N_\pi \rangle}{\langle N \rangle^2}
\langle N (N -1) \rangle~.
\ea
One checks that $\Phi$ and $\nu_{\rm dyn}$ both vanish when these moments
are substituted into Eq.~(\ref{Z2/N}) and Eq.~(\ref{nu-dyn}), respectively.
\begin{figure*}[t]
\centering
\vspace{0.4cm}
\includegraphics[width=0.49\textwidth]{./figures/mass_res_0-006.eps}
\hspace{0.1cm}
\includegraphics[width=0.49\textwidth]{./figures/mass_res_0-06.eps}
\vspace{-0.5cm}
\caption{(Color online) The distribution of $dE/dx$ with non-overlapping peaks
of pions and kaons (a), which allows unique particle identification, and the
distribution with overlapping peaks (b), which does not allow unique identification.}
\label{fig-dEdx_res}
\end{figure*}
It appears convenient for our further considerations to modify $\Phi$ to
the form
\be
\label{Psi-def}
\Psi = \frac{\langle Z^2 \rangle}{\langle N \rangle}
-\overline{z^2}~,
\ee
which preserves the properties of $\Phi$ - it vanishes in the absence of
inter-particle correlations and it is strongly intensive. The measure $\Psi$
will be used to formulate the identity method for the study of chemical fluctuations.
When expressed through moments of the multiplicity distribution, it equals
\ba
\label{Psi-moments}
\Psi
&= &
\frac{1}{\langle N \rangle^3}
\Bigg[\langle N_\pi^2 \rangle \langle N_K \rangle^2 +
\langle N_\pi \rangle^2 \langle N_K^2 \rangle
\\ \nonumber
&-& 2 \langle N_\pi \rangle \langle N_K \rangle
\langle N_\pi N_K \rangle
- \langle N_\pi \rangle^2 \langle N_K \rangle
- \langle N_\pi \rangle \langle N_K \rangle^2
\bigg]~.
\ea
Comparing Eq.~(\ref{nu-dyn}) to Eq.~(\ref{Psi-moments}) one finds
that $\Psi$ and $\nu_{\rm dyn}$ are proportional to each other:
\be
\label{Phi-nu-dyn}
\Psi = \frac {\langle N_\pi \rangle^2 \langle N_K \rangle^2}
{\langle N \rangle ^3} \; \nu_{\rm dyn}~.
\ee
We note here that $\nu_{\rm dyn}$ is not intensive but it becomes
even strongly intensive when multiplied by $\langle N \rangle$,
$\langle N_K \rangle$ or $\langle N_{\pi} \rangle$.
\section{Effect of misidentification}
\label{sec-misidentification}
As mentioned in the Introduction, complete identification of every particle is
impossible. In this section we show how the incomplete particle identification
influences the magnitudes of fluctuation measures. For this purpose we
considered a simple model of chemical fluctuations where the multiplicity of
pions is poissonian with a mean value of 100 and the number of kaons
is 20\% of the number of pions (strict correlation of the numbers of kaons
and pions). Actually, $N_K$ is taken as the integer number closest
to $N_\pi/5$ which is smaller or equal to $N_\pi/5$. The fluctuation measures
can be easily computed analytically for the model but our aim here is to simulate
the effect of incomplete particle identification.
There are many experimental techniques to measure particle mass. We discuss
here the effect of misidentification referring to measurements of energy loss,
$dE/dx$, in a detector material. This method is applied by the experiments NA49, NA61 and
STAR. The detectors are equipped with Time Projection Chambers in which $dE/dx$
is measured. The value of $dE/dx$ can be used to identify particles
because it depends both on particle mass $m$ and momentum $p$
in the combination of velocity ($\beta = p/\sqrt{m^2 + p^2}$).
In the case of large separation
of the energy loss distributions of pions and kaons, as schematically shown
in Fig.~\ref{fig-dEdx_res}a, almost unique particle identification is possible.
This is, however, not possible when the measured pion and kaon
$dE/dx$ distributions overlap, as illustrated in Fig.~\ref{fig-dEdx_res}b.
Performing the Monte Carlo simulations we assumed that the $dE/dx$ distributions
of pions and kaons are gaussians centered, respectively, at 1.4 and 1.2 in arbitrary units.
They are normalized to the mean multiplicity of pions and of kaons, respectively.
In order to quantify the bias caused by particle misidentification a simple particle
identification scheme is used, namely, a particle is identified as pion if $dE/dx > 1.3$
and as kaon if $dE/dx \le 1.3$. The width $\sigma$ of both gaussians
is chosen to be the same but its value is varied from 0 to 0.08. With growing width
of the peaks of the $dE/dx$ distribution, the fraction of misidentified particles
obviously increases. The results of the simulation are illustrated in
Fig.~\ref{fig-dEdx_res_phi_psi_nu} where the fluctuation measures $\Phi$, $\Psi$
and $\nu_{\rm dyn}$ are shown as a function of $\sigma$. As seen, the magnitudes
of $\Phi$, $\Psi$ and $\nu_{\rm dyn}$ decrease as the fraction of
misidentified particles grows and the measures vanish when particle identification
becomes totally random. This sizable and experimentally unavoidable effect was
the main motivation to develop the identity method which fully eliminates the problem.
\begin{figure*}[t]
\centering
\vspace{-0.3cm}
\includegraphics[width=0.49\textwidth]{./figures/fichem_psichem_sigmadEdx.eps}
\includegraphics[width=0.49\textwidth]{./figures/nudyn_sigmadEdx.eps}
\vspace{-1.0cm}
\caption{(Color online) The measures of chemical fluctuations $\Phi$, $\Psi$ (a) and
$\nu_{\rm dyn}$ (b) as functions of the width of the energy-loss distribution.}
\label{fig-dEdx_res_phi_psi_nu}
\end{figure*}
\section{Identity Method}
\label{sec-id-method}
The identity method, which is described here for a two-component system of pions and
kaons, utilizes the measure $\Psi$ defined by Eq.~(\ref{Psi-def}). However, the kaon
identity $w_K$ is not limited to either 1 or 0 any more, but can take
any value from the interval $[0,1]$.
In the previous section we assumed that particles are identified according to
the energy-loss distribution. To make the presentation of the identity method
more general we assume here that particle identification is achieved by
measurement of particle mass not specifying the particular experimental technique
which is used for this purpose. Since any measurement is of finite resolution,
we deal with continuous distributions of observed masses of pions and kaons
which are denoted as $\rho_\pi (m)$ and $\rho_K (m)$, respectively. They
are normalized as
\be
\label{norm-rho-K}
\int dm \,\rho_\pi (m) = \langle N_\pi \rangle~,
\;\;\;\;\;\;
\int dm \,\rho_K (m) = \langle N_K \rangle~.
\ee
The kaon identity is defined as:
\be
w_K (m) \equiv \frac{\rho_K (m)}{\rho (m)}~,
\ee
where $\rho (m) \equiv \rho_\pi (m) + \rho_K (m)$, and is
normalized as
\be
\label{norm-rho}
\int dm \,\rho (m) = \langle N \rangle
\equiv \langle N_\pi \rangle + \langle N_K \rangle~.
\ee
If the distributions $\rho_\pi (m)$ and $\rho_K (m)$ do not overlap,
the particles can be uniquely identified and $w_K = 0$ for a pion and $w_K = 1$
for a kaon. When the distributions $\rho_\pi (m)$ and $\rho_K (m)$
overlap, $w_K$ can take the value of any real number from $[0,1]$. Figure~\ref{rho_real_identif}
illustrates the latter case. The mass distributions are shown in
Fig.~\ref{rho_real_identif}a and the distribution of kaon identity in
Fig.~\ref{rho_real_identif}b.
The peaks close to~0 and~1 in Fig.~\ref{rho_real_identif}b correspond to the
mass regions in which pions, respectively kaons, are well identified.
The $w_K$ values around 0.5 correspond to particles for which
the measured mass is in the transition region between the kaon and pion
peaks ($m \approx 280$~MeV) in the distribution $\rho(m)$
shown in Fig.~\ref{rho_real_identif}a.
Let us now explain how the fluctuation measure $\Psi$ is calculated once the mass
distributions were experimentally obtained. The single particle variable entering
Eq.~(\ref{Psi-def}) is defined as in Sec.~\ref{sec-measures}:
$z \equiv w_K - \overline{w_K}$. The bar denotes the inclusive average
which is computed as follows:
\ba
\overline{w_K} &\equiv& \frac{1}{\langle N \rangle }
\int dm \,\rho (m) \, w_K(m)
\\ \nonumber
&=& \frac{1}{\langle N \rangle } \int dm \,\rho_K (m)
= \frac{\langle N_K \rangle }{\langle N \rangle }~.
\ea
Analogously one finds $\overline{w_K^2}$ and
$\overline{z^2} \equiv \overline{w_K^2} - \overline{w_K}^2 $.
The quantity $\langle Z^2 \rangle$ is obtained as
\be
\langle Z^2 \rangle =
\frac{1}{N_{\rm ev}} \sum_{n=1}^{N_{\rm ev}}
\bigg( \sum_{i=1}^{N_n} w_K^i - N_n \, \overline{w_K}\bigg)^2~,
\ee
where $N_{\rm ev}$ is the number of events and $N_n$
is the multiplicity of the $n-$th event. Substituting $\langle Z^2 \rangle$,
$\overline{z^2}$ and $\langle N \rangle$ into Eq.~(\ref{Psi-def}),
one finds the measure $\Psi$, the magnitude of which, however,
is biased by the effect of particle misidentification.
Next we discuss the correction procedure.
As shown in the Appendix, the measure $\Psi$ can be
expressed through the moments of the multiplicity distributions of pions
and kaons as
\be
\label{Psi-final}
\Psi = A
\bigg(\frac{\langle N_K \rangle}{\langle N \rangle}
- \overline{u_K} \bigg)^2~,
\ee
where
\ba
\nonumber
A &\equiv&
\frac{1}{\langle N \rangle}
\bigg[\langle N_\pi^2 \rangle
\frac{\langle N_K \rangle^2 }{\langle N_\pi\rangle^2}
+ \langle N_K^2 \rangle - \langle N_K \rangle
- \frac{\langle N_K \rangle^2 }{\langle N_\pi\rangle}
\\ \label{A}
&& \;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;
-2 \langle N_\pi N_K \rangle \frac{\langle N_K \rangle}{\langle N_\pi\rangle}
\bigg]~,
\ea
and
\be
\overline{u_K}
\equiv \frac{1}{\langle N_K \rangle }
\int dm \,\rho_K (m) \, w_K(m)~.
\ee
\begin{figure*}[t]
\vspace{0.3cm}
\centering
\includegraphics[width=0.49\textwidth]{./figures/mass_distr.eps}
\hspace{0.1cm}
\includegraphics[width=0.49\textwidth]{./figures/w_mis.eps}
\vspace{-0.5cm}
\caption{(Color online) The distributions of observed masses of pions $\rho_\pi (m)$,
of kaons $\rho_K (m) $, and their sum $\rho (m)$ (a) and the corresponding
distribution of kaon identity $w_{K}$ (b).}
\label{rho_real_identif}
\end{figure*}
In the case of complete particle identification (CI)
the distributions $\rho_\pi (m)$ and $\rho_K (m)$
do not overlap and thus $\overline{u_K} = 1$. Then, the result for $\Psi$ is:
\be
\label{Psi-CI}
\Psi_{\rm CI} =
A \bigg(\frac{\langle N_K \rangle}{\langle N \rangle}
- 1 \bigg)^2~,
\ee
which is equivalent to the expression (\ref{Psi-moments}).
Although particle-by-particle identification is usually difficult, statistical
identification is reliable. In the latter case, we do not know whether a given particle is a kaon,
but we know the average numbers of kaons and of pions. We introduce the concept
of {\em random identification} which assumes that for every particle the
probability of being a kaon equals $\langle N_K \rangle/\langle N \rangle$. Such a
situation is described by mass distributions of the form
\begin{displaymath}
\rho_i(m) = \left\{
\begin{array}{ccc}
0 & {\rm for}& m < m_{\rm min}~,
\\[2mm]
\frac{\langle N_i \rangle}{ m_{\rm max} - m_{\rm min}} & {\rm for}&
m_{\rm min} \le m \le m_{\rm max}~,
\\ [2mm]
0 & {\rm for}& m_{\rm max} < m~,
\end{array}
\right.
\end{displaymath}
where $i=\pi, K$ and $m_{\rm min}$ and $m_{\rm max}$ denote lower and upper limits
of the measured mass range, respectively. With this distribution
\be
\label{wk-randnom}
\overline{u_K} = \overline{w_K} =
\frac{\langle N_K \rangle}{\langle N \rangle}~.
\ee
When $\overline{u_K} = \langle N_K \rangle /\langle N \rangle$
is substituted into Eq.~(\ref{Psi-final}), $\Psi = 0$.
Thus, the measure $\Psi$ vanishes when
particle identification is random.
Finally, we arrive at the crucial point of the considerations.
It appears that the
measure $\Psi$ can be expressed as
\be
\label{conjecture}
\Psi = \Psi_{\rm CI} \Big(1 - \frac{V_I}{V_R} \Big)^2~,
\ee
where $\Psi_{\rm CI}$ is the measure $\Psi$ for the complete identification,
as given by Eq.~(\ref{Psi-CI}). The quantities $V_I$ and $V_R$ are the
values of the integral
\be
\label{V}
V \equiv \int dm \, \rho(m) \, w_K(m) \, \big(1 - w_K(m) \big)~,
\ee
evaluated for the cases of imperfect and random identification, respectively.
One proves the equality (\ref{conjecture}) by observing that
\be
\label{reso}
V_I = \langle N_K \rangle ( 1 - \overline{u_K})~,
\;\;\;\;\;\;
V_R = \frac{ \langle N_K \rangle \langle N_\pi \rangle}
{\langle N \rangle}~.
\ee
Substituting Eqs.~(\ref{Psi-CI}, \ref{reso}) into the equality (\ref{conjecture}),
one obtains the formula (\ref{Psi-final}). Actually, the relation (\ref{conjecture})
was first discovered by performing various numerical simulations and only then it was
proven analytically. Equation~(\ref{conjecture}) allows one to experimentally obtain
$\Psi_{\rm CI}$ from $\Psi$. Thus, the effect of misidentification is fully corrected by
the factor $(1- V_I/V_R)^2$ which measures the quality of the applied procedure of
particle identification. The factor is independent of the correlations under study and
can be determined from experimental data.
\section{Experimental procedure}
The application of the identity method to experimental data is not difficult. We present
here a step-by-step procedure to obtain the measure $\Psi_{\rm CI}$ of fluctuations
of kaons (or any other selected particle type) with
respect to all particles
(represented by the sum of kaons and pions in the previous sections).
It is important to note that if misidentification occurs between more
than two particle types, the identity method does not allow to study relative
fluctuations of two of them, e.g. of kaons and pions in the presence of protons.
We come back to the specific, but typical, example
considered in Sec.~III of
particle identification via measurements of
particle energy loss in the detector material. The energy loss
$dE/dx$ is denoted by $X$. The energy loss of particles of a given type depends
on the particle mass and momentum (via the velocity)
and detector characteristics. The
distribution of $X$ is usually a multi-dimensional function, which can be determined
experimentally by averaging over particles from many interactions. The energy
loss distribution, which for a given particle momentum is typically fitted by a sum
of four gaussians corresponding to electrons, pions, kaons and (anti-)protons, allows
one to determine the average multiplicities of kaons and of all particles. Having
obtained this information, one should proceed as follows.
\begin{enumerate}[(i)]
\setlength{\itemsep}{1pt}
\item Extract the energy-loss distribution of kaons $\rho_K(X)$ from the inclusive
distribution $\rho(X)$. The distributions should be normalized as
$$
\int dX \, \rho(X) = \langle N \rangle~, \;\;\;\;\;\;
\int dX \, \rho_K(X) = \langle N_K \rangle~.
$$
\item
Determine the kaon identity
$$w_K(X)=\frac {\rho_K(X)} {\rho(X)}
$$
for every registered particle.
\item
Compute the fluctuation measure $\Psi$ from the definition (\ref{Psi-def}).
This is the {\em raw} value of $\Psi$ which is not corrected yet for the effect
of misidentification.
\item
Knowing the mean multiplicities, compute the quantity
$$
V_R = \frac{\langle N_K \rangle ( \langle N \rangle-\langle N_K \rangle)}
{\langle N \rangle}~.
$$
\item
Using all particles calculate the integral
$$
V_I= \int dX \rho(X) \, w_K(X) \,\big(1-w_K(X)\big)~.
$$
\item
Determine the fluctuation measure $\Psi_{\rm CI}$, which is free of
the effect of misidentification, as
$$
\Psi_{\rm CI} = \frac{\Psi}{(1-V_I/V_R)^2}~.
$$
\end{enumerate}
The experimental data accumulated by the NA49 Collaboration are
currently under analysis using the identity method proposed
in this paper~\cite{Mackowiak:2011sc}.
\acknowledgments
We are very grateful to Peter Seyboth for strong encouragement and
numerous critical discussions. We are also indebted to Mark Gorenstein
for critical reading of the manuscript. This work was partially supported by
the Polish Ministry of Science and Higher Education under grants N N202 204638
and 667/N-CERN/2010/0 and by the German Research Foundation under
grant GA 1480/2-1.
\begin{widetext}
|
\section{introduction}\label{intro}
A dynamical system on a measurable space $\mathbb{S}$ is understood
as a triple $(\mathbb{S};\mu;f)$, where $\mathbb{S}$ is a set
endowed with a measure $\mu
$, an
\[
f:\mathbb{S\rightarrow S
\]
is a measurable function, that is, the $f$-preimage of any measurable
subset is a measurable subset.
A trajectory of the dynamical system is a sequenc
\begin{equation}\label{e:cycle}
x_{0}, f(x_{0}), f^{(2)}(x_{0}), f^{(3)}(x_{0}), \cdots
\end{equation}
of points of the space $\mathbb{S}$, $x_{0}$ is called an initial
point of the trajectory. If ${\mathbb S}$ is a finite set of $k$
elements and every element of ${\mathbb S}$ appears in the trajectory (\ref{e:cycle}),
then the finite sequence
\[
x_{0}, f(x_{0}), f^{(2)}(x_{0}), f^{(3)}(x_{0}), \cdots, f^{(k-1)}(x_0)
\]
is a single cycle (also called single orbit) of ${\mathbb S}$ given by the function $f$.
\begin{defn}
A mapping $F:\mathbb{S}\rightarrow\mathbb{Y}$ of a measurable space
$\mathbb{S}$ into a measurable space $\mathbb{Y}$ endowed with
probabilistic measure $\mu$ and $\nu$, respectively, is said to be
measure-preserving whenever $\mu(F^{-1}(S))=\nu(S)$ for each
measurable subset $S\subset$ $\mathbb{Y}$. In case ${\mathbb S}=$
$\mathbb{Y}$ and $\mu=\nu$, a measure preserving mapping $F$ is said
to be ergodic if $F^{-1}(S)=S$ for a measurable set $S$ implies
either $\mu(S)=1$ or $\mu(S)=0$.
\end{defn}
In non-Archimedean dynamics, the measurable space ${\mathbb S}$
would be taken to be a complete discrete valuation ring. There are
two kinds of such rings as follows,
\begin{itemize}
\item[(1)] characteristic $0$ case: finite extensions ${\mathcal O}$ of rings ${\Z}_p$ of $p$-adic integers for
prime numbers $p$\,;
\item[(2)] positive characteristic case: rings ${\F}_r[[T]]$ of formal power series over finite fields
${\F}_r$, where ${\F}_r$ is the field of $r$ elements and $r$ is a
power of some prime number $p$.
\end{itemize}
Both of these two kinds of complete rings are commutative and
compact, so these rings are certainly equipped with Haar measures with respect to
addition, which we always
assume to be normalized so that the measure of the total space in all cases is
equal to $1$.
In the characteristic $0$ case, much has been studied when the space
is the ring ${\Z}_p$ of $p$-adic integers, especially when the prime
number $p=2$. On ${\Z}_p$, the normalized measure $\mu$ and $p$-adic
absolute value $|\cdot|_p$ are related by the following: for any
$0\ne a \in {\Z}_p$, we can express it as $a=p^{{\rm ord}(a)}\cdot
u$ with ${\rm ord}(a)$ a non-negative integer and
$u=\sum_{i=0}^\infty u_i\, p^i$, where $u_i\in \{0, 1, 2, \cdots, p-1\}$
for each $i$ and $u_0\ne 0$, then $\mu(a{\Z}_p)=(1/p)^{{\rm
ord}(a)}=|a|_p$. To study measure preserving functions on ${\Z}_p$,
we at first consider $1$-Lipschitz functions. A function
$f:{\Z}_p\to {\Z}_p$ is said to satisfy the $1$-Lipschitz condition
if
\[
|f(x+y)-f(x)|_p\le |y|_p, \text{ for any } x,y\in {\Z}_p.
\]
Such a condition is also called ``compatibility" condition. It is clear
that a $1$-Lipschitz function on ${\Z}_p$ is continuous.
A continuous function $f: {\Z}_p\to {\Z}_p$ has its Mahler
expansion:
\[
f(x)=\sum_{i=0}^{\infty}a_{i}\dbinom{x}{i}, \quad\text{ with
$a_{i}\in\mathbb{Z}_{p}$\,, and $a_i\to 0$ as $i \to \infty$},
\]
where $\binom{x}{i}=x(x-1)\cdots(x-i+1)/i$ for an integer $i\ge 1$
and $\binom{x}{0}=1$. In the work of
\cite{An1}\cite{An1998}\cite{An2}\cite{An3}\cite{An4}, some
sufficient and necessary conditions on the Mahler coefficients for
$f$ to be $1$-Lipschitz or measure-preserving are given. When $p$ is
odd (respectively, $p=2$), the sufficient (respectively, sufficient
and necessary) conditions on the Mahler coefficients and the Van der
Put coefficients are also given for $f$ to be ergodic in these
papers.
Ergodic functions enjoy many applications in cryptography theory. At
the beginning of 1990-th, V. Anashin \cite{An1} (1994) studied
differentiable functions on ${\Z}_p$ and gave criteria for
measure-preservation and ergodicity of $1$-Lipschitz transformations
on ${\Z}_2$. He did further work in the theory of
measure-preservation and ergodicity of $1$-Lipschitz functions on
${\Z}_p$ in \cite{An2} (2002). The ergodic theory for $1$-Lipschitz
transformations on the ring of $2$-adic integers is of
cryptographical significance in application. A. Klimov and A. Shamir
\cite{KS2002}\cite{KS2003}\cite{KS2004} proposed the concept of
$T$-function and used $T$-functions to produce long period
pseudo-random sequences, some of which are used as primitive
components in stream ciphers. Mathematically, these
results are derivable from V. Anashin's work \cite{An1}. Most
recently, V. Anashin et al \cite{An4} have used the Van der Put
basis to describe the ergodic functions on ${\Z}_2$.
In the positive characteristic case, we need to consider the ring ${\F}_r[[T]]$ of formal power series
over ${\F}_r$. The absolute value $|\cdot|_T$ on ${\F}_r[[T]]$ is normalized by $|T|_T=1/r$.
Carlitz polynomial functions $G_i(x)$ ($i\ge 0$) (Carlitz polynomials are reviewed in section \ref{carlitz})
are analogues of the binomial functions $\binom{x}{i}$, and every continuous function
$f: {\F}_r[[T]]\to {\F}_r[[T]]$ can be expressed as
\[
f(x)\sum\limits_{i=0}^\infty a_i G_i(x), \quad\text{ where $a_i\in {\F}_r[[T]]$, and $a_i\to 0$ as $i\to \infty$}.
\]
There are also analogous results on the theory of differentiability
of functions over ${\F}_r[[T]]$, for example, in \cite{Wa}\cite{Ya}.
Hence some natural questions would be asked: do we also have a
measure-preservation and ergodicity theory over ${\F}_r[[T]]$ or
${\F}_2[[T]]$? and if such a theory exists, how does it compare to
the theory over ${\Z}_p$? The purpose of this paper tries to answer
the first question, that is, to establish a theory of
measure-preservation and ergodicity over ${\F}_2[[T]]$.
In this paper, section \ref{prelim} will recall some of the existing
results on the theory of measure-preservation and ergodicity over
${\Z}_p$. In section \ref{ergodic}, we will give a description of
measure-preservation and ergodicity of $1$-Lipschitz functions over
${\F}_2[[T]]$ in terms of the Van der Put basis. In section
\ref{carlitz}, the conditions of a continuous function on
${\F}_r[[T]]$ to satisfy the $1$-Lipschitz condition are given in
terms of the Carlitz basis. Section \ref{ecarlitz} gives necessary
and sufficient conditions for a $1$-Lipschitz function on
${\F}_2[[T]]$ to be ergodic in terms of Carlitz basis. And finally, section \ref{discussion}
gives some discussion on the topics related to this paper.
The work of this paper is inspired by the ergodic theory established
by V. Anashin et al. The first author and second author want to
thank V. Anashin for many helpful talks about the topics of $p$-adic
dynamical systems and related problems. The authors also appreciate
the referees for many suggestions on the paper to make the
exposition better.
\section{Preliminary}\label{prelim}
In this section, we will list some of the existing results on the theory of
measure-preservation and ergodicity of $1$-Lipschitz
transformations of ${\Z}_p$.
For a real number $x$, let $\lfloor x \rfloor$ denote the maximum
integer which is smaller than or equal to $x$.
\begin{prop}[\cite{An1}]\label{prop:1}
A $1$-Lipschitz function $f:\mathbb{Z}_{p}\rightarrow\mathbb{Z}_{p}$
is measure-preserving (respectively, ergodic) if and only if it is bijective
(respectively, transitive) modulo $p^{k}$ for all
integers $k\ge 0$.
\end{prop}
\begin{theorem}[\cite{An1}, \cite{An2}]\label{thm:An1}
(measure-preservation property) The function
$f:\mathbb{Z}_{p}\rightarrow\mathbb{Z}_{p}\,,$
\[
f(x)=\sum_{i=0}^{\infty}a_{i}\dbinom{x}{i}
\]
defines a measure-preserving $1$-Lipschitz transformation on
$\mathbb{Z}_{p}$ whenever the following conditions hold
simultaneously:
$$
\begin{array}{ll}
&a_{1}\not \equiv 0 \quad (\operatorname{mod} p),\\
&a_{i} \equiv0 \quad (\operatorname{mod} p^{\lfloor\log_{p}(i)\rfloor +1}), \quad
i=2,3,\cdots.
\end{array}
$$
The function f defines an ergodic $1$-Lipschitz transformation on
$\mathbb{Z}_{p}$ whenever the following conditions hold
simultaneously:
$$
\begin{array}{ll}
&a_{0}\not \equiv0 \quad (\operatorname{mod}p),\\
&a_{1}\equiv1 \quad (\operatorname{mod}p), \quad for \ p \ odd, \\
&a_{1}\equiv1 \quad (\operatorname{mod}4), \quad for \ p=2, \\
&a_{i}\equiv0 \quad (\operatorname{mod}p^{\lfloor \log_{p}(i+1)\rfloor +1}), \quad
i=2,3,\cdots.
\end{array}
$$
Moreover, in the case $p = 2$ these conditions are necessary.
\end{theorem}
For any non-negative integer $m$, the Van der Put function
$\chi(m,x)$ on ${\mathbb Z}_p$ is the characteristic function of the ball $B_{p^{-\lfloor \log_p(m)\rfloor -1}}(m)$
which is centered at $m$ and is of radius $p^{-\lfloor \log_p(m)\rfloor -1}$:
\[
\chi(m,x)=
\begin{cases}
1, & \quad \text{ if } |x-m|_p\le p^{-\lfloor \log_p (m)\rfloor -1},\\
0, & \quad \text{ otherwise, }
\end{cases}
\]
for $m\ne 0$, and is the characteristic function of the ball
$B_{1/p}(0)$ which is centered at $0$ and is of radius $1/p$:
\[
\chi(0,x)=
\begin{cases}
1, & \quad \text{ if } |x|_p\le 1/p,\\
0, & \quad \text{ otherwise .}
\end{cases}
\]
The Van der Put functions $\chi(m,x)$ constitute an orthonormal
basis of the space $C({\Z}_p, {\Q}_p)$ of the continuous functions
from ${\Z}_p$ to ${\Q}_p$ (see Theorem 1 to Theorem 4, Chapter 16, \cite{Ma}). In terms of the Van der
Put basis $\{\chi(m,x)\}_{m \ge 0}$, we have
\begin{theorem}[\cite{An4}]\label{thm:An4}
A function $f:\mathbb{Z}_{2}\rightarrow\mathbb{Z}_{2}$ is compatible
and preserves the Haar measure if and only if it can be represented
as
$$
f(x)=b_{0}\chi(0, x)+ b_{1}\chi(1, x)+
\sum\limits_{m=2}^{\infty}2^{\lfloor \log_{2}(m)\rfloor }b_{m}\chi(m, x),
$$
where $b_{m}\in \mathbb{Z}_{2}$ for $m=0,1,2,\cdots$, and
\begin{itemize}
\item $b_{0}+b_{1}\equiv1(\operatorname{mod}2)$
\item $\left\vert b_{m}\right\vert _{2}=1,$ for $m\geq2$.
\end{itemize}
\end{theorem}
\begin{theorem}[\cite{An4}]\label{thm:An4.2}
A function $f:\mathbb{Z}_{2}\rightarrow\mathbb{Z}_{2}$ is compatible
and ergodic if and only if it can be represented as
$$
f(x)=b_{0}\chi(0, x)+ b_{1}\chi(1, x)+
\sum\limits_{m=2}^{\infty}2^{\lfloor \log_{2}(m)\rfloor}b_{m}\chi(m, x),
$$
where $b_{m}\in \mathbb{Z}_{2}$ for $m=0,1,2,\cdots$, and
\begin{itemize}
\item $b_{0}\equiv1 \ (\operatorname{mod}2),\ b_{0}+b_{1}\equiv 3 \ (\operatorname{mod}4),
\ b_{2}+b_{3}\equiv2 \ (\operatorname{mod}4)$,
\item $\left\vert b_{m}\right\vert _{2}=1,$ for $m\geq2$,
\item
$\sum_{m=2^{n-1}}^{2^{n}-1}b_{m}\equiv0(\operatorname{mod}4)$, for
$n\geq3$.
\end{itemize}
\end{theorem}
The two complete discrete valuation rings $\mathbb{F}_{p}[[T]]$ and
$\mathbb{Z}_{p}$ are homeomorphic. Therefore in the same way we
define the Van der Put basis $\{\chi(\a, x)\}_{\a\in {\F}_p[T]}$ on
the space of functions over $\mathbb{F}_{p}[[T]]$:
\[
\chi(\alpha,x)=
\begin{cases}
1, & \quad \text{ if }x\in B_{p^{-\deg_T(\alpha)-1}}(\a),\\
0, & \quad \text{ otherwise, }
\end{cases}
\]
is the characteristic function of the ball
\[
B_{p^{-\deg_T(\alpha)-1}
}(\alpha) =\{x\in {\F}_p[[T]]: \,\,\left\vert x-\alpha\right\vert
_{T}\leq p^{-\deg_T(\alpha )-1}\}
\]
which is centered at $\a$ and is of radius
$|T|_T^{\, \deg_T(\a)+1}$ if ${\a}\ne 0$, and
\[
\chi(0,x)=
\begin{cases}
1, & \quad \text{ if }x\in B_{p^{-1}}(0),\\
0, & \quad \text{ otherwise, }
\end{cases}
\]
is the characteristic function of the ball $B_{p^{-1}}(0)$ centered at $0$ and of radius $p^{-1}$.
In a similar way to the one used in the proof
of the Theorem 1 to Theorem 4 of Chapter 16 in Mahler's book \cite{Ma}, we can see that
every continuous function $f:\mathbb{F}_{p}[[T]]\rightarrow
\mathbb{F}_{p}[[T]]$ can be expressed as
\[
f(x)=\sum_{\alpha\in\mathbb{F}_{p}[T]}B_{\alpha}\chi(\alpha,x),\quad
B_{\alpha}\in\mathbb{F}_{p}[[T]], \,\,\, \text{ with } B_{\a}\to 0
\text{ as } \deg_T{\a}\to \infty.
\]
\section{Ergodic Functions over ${\F}_2[[T]]$ and Van der Put Expansions}\label{ergodic}
We will discuss the ergodic functions over ${\F}_2[[T]]$ and which conditions of
the expansion coefficients under the Van der Put
basis should be satisfied.
The absolute value on the discrete valuation ring ${\F}_2[[T]]$ is
normalized so that $|T|_T=1/2$.
Suppose $f: {\F}_2[T]\to {\F}_2[T]$ is a map. Then $f$ can be
expressed in terms of Van der Put basis:
\begin{equation}\label{e:2.1}
f(x)=\sum_{\a\in {\F}_2[T]} B_{\a}\chi(\a, x).
\end{equation}
If the function $f$ is continuous under the $T$-adic topology, then
$B_{\a}\to 0$ as $\deg_T(\a)\to \infty$, and $f$ can be extended to a
continuous function from ${\F}_2[[T]]$ to itself, which is still
denoted by $f$. The coefficients $B_{\a}$ of the expansion
(\ref{e:2.1}) can be calculated as follows (see Chapter 16 of Mahler's book \cite{Ma} for the detail):
\begin{itemize}
\item $B_0=f(0), B_1=f(1)$;
\item $B_{\a}=f(\a)-f({\a}-\a_n T^n)$, if ${\a}=\a_0+\a_1T+\cdots +\a_n T^n \in {\F}_2[T]$ with $\a_n\ne
0$ ( that is, $\a_n=1$ ) is of degree greater than or equal to $1$.
\end{itemize}
Also for ${\a}=\a_0+\a_1T+\cdots +\a_n T^n \in {\F}_2[T]$, we denote
by
\[
{\a}_{[k]}=\a_0 +\a_1 T + \cdots + \a_k T^k
\]
for any $k$ between $0$ and $n$.
\begin{theorem}
A continuous function $f:
\mathbb{F}_{2}[[T]]\rightarrow\mathbb{F}_{2}[[T]]$ is $1$-Lipschitz
if and only if it can be expressed as
\[
f(x)=b_0\chi(0,x)+\sum_{\alpha\in\mathbb{F}_{2}[T]\backslash\{0\}}T^{\deg_{T}(\alpha)
}b_{\alpha}\chi(\alpha,x), \quad \text{ with } b_{\a}\in
{\F}_2[[T]].
\]
\end{theorem}
\begin{proof}
Suppose $f$ is $1$-Lipschitz. Then it is clear that $b_0=B_0=f(0)\in
{\F}_2[[T]]$ and $b_1=B_1=f(1)\in {\F}_2[[T]]$. For
${\a}=\a_0+\a_1T+\cdots +\a_n T^n \in {\F}_2[T]$ of degree $n\ge 1$,
we have
\[
\left\vert B_{\alpha}\right\vert _{T}=\left\vert
f(\alpha)-f(\alpha-\alpha_{n} T^{n})\right\vert _{T}\leq\left\vert
\alpha_{n}T^{n}\right\vert _{T} =2^{-n}.
\]
Therefore $B_{\alpha}$ can be written as $B_{\alpha}=T^{\deg_{T}
(\alpha)}b_{\alpha}$ with $b_{\a}\in {\F}_2[[T]]$.
Conversely, suppose $f(x)=b_0\chi(0,x)+\sum_{\alpha\in\mathbb{F}_{2}
[T]\backslash \{0\}}T^{\deg_{T}(\alpha)}b_{\alpha}\chi(\alpha,x)$ with
$b_{\a}\in {\F}_2[[T]]$. If $X,Y$ both belong to
$\mathbb{F}_{2}[[T]]$ and $X\equiv Y(\operatorname{mod}T^{n})$, then
\[
\chi(\a,X)=\chi(\a,Y) \text{ for any $\a\in {\F}_2[[T]]$ with $\a=0$
or $\deg_T(\a)< n$.}
\]
Therefore we have $f(X)\equiv f(Y)(\operatorname{mod}T^{n})$, hence
the function $f$ is $1$-Lipschitz.
\end{proof}
\begin{prop}
A $1$-Lipschitz function
$f:\mathbb{F}_{p}[[T]]\rightarrow\mathbb{F}_{p}[[T]]$ is
measure-preserving (respectively, ergodic) if and only if $f$ is
bijective modulo $T^{k}$(respectively, transitive modulo $T^{k}$)
for all integers $k> 0$.
\end{prop}
\begin{proof}
There is a topological bijection
\[
\begin{array}{rrcl}
\sigma: &{\Z}_p&\to &{\F}_p[[T]]\\
& \sum_{i=0}^\infty c_i p^i &\mapsto &\sum_{i=0}^\infty c_i T^i,
\end{array}
\]
where $c_i\in\{0, 1, \cdots, p-1\}$ for each $i$. The map $\sigma$ preserves
the Haar measures, that is, if we denote the normalize Haar measure on ${\Z}_p$
by $\mu$ and that on ${\F}_p[[T]]$ by $\nu$, then $\mu(S)=\nu(\sigma(S))$ for
any measurable subset $S\subset {\Z}_p$. Therefore that $f: \mathbb{F}_{p}[[T]]\rightarrow\mathbb{F}_{p}[[T]]$
is measure-preserving is equivalent to saying that the function
$\sigma^{-1}\circ f\circ \sigma: {\Z}_p\to {\Z}_p$ is measure-preserving, which
is equivalent to saying that $\sigma^{-1}\circ f\circ \sigma$ is bijective modulo $p^k$ for
all positive integers $k$ by Proposition \ref{prop:1}, hence is also
equivalent to saying that $f:\mathbb{F}_{p}[[T]]\rightarrow\mathbb{F}_{p}[[T]]$
is bijective modulo $T^k$ for all positive integers $k$ since $\sigma$ is measure-preserving.
The proof about the ergodicity property is similar.
\end{proof}
\begin{theorem}\label{theorem:mp}\bigskip(measure-preservation property)
A $1$-Lipschitz function
$f:\mathbb{F}_{2}[[T]]\rightarrow\mathbb{F}_{2}[[T]]$,
\[
f(x)=b_0\chi(0,x)+\sum_{\alpha\in\mathbb{F}_{2}[T]\backslash\{0\}}T^{\deg_{T}(\alpha)
}b_{\alpha}\chi(\alpha,x), \quad \text{ with } b_{\a}\in {\F}_2[[T]]
\]
is measure preserving if and only if the following conditions hold
simultaneously:
\begin{itemize}
\item[(1)] $b_{0}+b_{1}\equiv1(\operatorname{mod}T)$;
\item[(2)] $\left\vert b_{\alpha}\right\vert_T =1,$ for $\deg_T(\alpha)\geq 1$.
\end{itemize}
\end{theorem}
\begin{proof}
Suppose $f$ is bijective $\operatorname{mod}T^{n}$ for all
$n\in\mathbb{N}$, we need to show that the two conditions for the
Van der Put coefficients are satisfied. At first, from the
bijectivity of $\operatorname{mod}T$, we get
\begin{itemize}
\item[] $f(0)=b_{0}\equiv 1\,(\operatorname{mod}T)$; $f(1)=b_{1}\equiv
0\,(\operatorname{mod}T)$; or
\item[] $f(0)=b_{0}\equiv 0\,(\operatorname{mod}T)$; $f(1)=b_{1}\equiv
1\,(\operatorname{mod}T)$.
\end{itemize}
Thus we get $b_{0}+b_{1}\equiv 1\,(\operatorname{mod}T)$. Secondly,
we consider the bijectivity of the function $f$ $\mod T^{2}$. As
$f(T)-f(0) \not\equiv 0(\operatorname{mod}T^{2})$, we get
$b_{0}\chi(0,T)+Tb_{T}\chi(T,T)-b_{0} \chi(0,0)=Tb_{T}\not\equiv
0(\operatorname{mod}T^{2})$, therefore $\left\vert b_{T}\right\vert_T
=1$; Also $f(1+T)-f(1)\not\equiv 0(\operatorname{mod}T^{2})$ implies
$b_{1}\chi(1,1+T)+Tb_{1+T}\chi(1+T,1+T)-b_{1}\chi(1,1)=Tb_{1+T}\not\equiv
0(\operatorname{mod}T^{2})$, therefore $\left\vert
b_{1+T}\right\vert_T=1$. In the same way for the general case when
$\deg_T(\alpha)=n$, we use the bijectivity of the function
$\operatorname{mod}T^{n+1}$, thus $f(\alpha)-f(\alpha-\alpha_{n}T^{n
)=T^{\deg_{T}(\alpha)}b_{\alpha}=T^{n}b_{\alpha}\not\equiv
0(\operatorname{mod}T^{n+1})$, therefore $\left\vert
b_{\alpha}\right\vert_T=1.$
Conversely, suppose the two conditions for Van der Put coefficients
are satisfied. As $f(0)=b_{0}, f(1)=b_{1}$, we see that the first
condition implies the bijectivity $\operatorname{mod}T$. To derive
the bijectivity $\operatorname{mod}T^{n}$, $n\ge 2$, we choose
\[
X=X_0 + X_1 T + \cdots + X_{n-1} T^{n-1}, \quad Y=Y_0 + Y_1 T +
\cdots + Y_{n-1}T^{n-1}
\]
such that $f(X)-f(Y)\equiv 0(\operatorname{mod}T^{n})$. If $X\not\equiv
Y(\operatorname{mod}T^{n})$, then we denote the first integer $m$
between $0$ and $n-1$ such that $X_{m}\neq Y_{m}$ and consider the
equation $f(X)-f(Y)\equiv 0(\operatorname{mod}T^{m+1})$. Then we can
assume $X_{[m]}=X_{[m-1]}+T^m$ and $Y_{[m]}=X_{[m-1]}$ for convenience.
Therefore the coefficient $B_{\a}$ of the expansion of $f(x)$ at ${\a}=X_{[m]}$ is
\[
B_{X_{[m]}}=f(X_{[m]})-f(X_{[m-1]})=f(X_{[m]})-f(Y_{[m]})\equiv 0 (\operatorname{mod}T^{m+1}),
\]
which contradicts to $B_{X_{[m]}}=T^m\cdot b_{X_{[m]}}\not\equiv 0 (\operatorname{mod}T^{m+1})$.
Therefore $X\equiv
Y(\operatorname{mod}T^{n})$, and so $f$ is injective. Hence $f$ is
bijective $\operatorname{mod}T^{n}$, as $\mathbb{F}_{2}[[T]]/(T^{n}\cdot {\F}_2[[T]])$
is a finite set.
\end{proof}
For any $x\in {\F}_2[[T]]$, we write
\begin{equation}\label{e:exp}
x=\sum_{i=0}^\infty x_i T^i.
\end{equation}
\begin{lemma}\label{lem:2.1}
Suppose a measure-preserving $1$-Lipschitz function $f:\mathbb{F}_{2}[[T]]\rightarrow\mathbb{F
_{2}[[T]]$ is transitive (single orbit) over
$\mathbb{F}_{2}[[T]]/(T^{n}\cdot {\F}_2[[T]]),$ $n\ge 1$. Then $f$
is transitive over $\mathbb{F}_{2}[[T]]/(T^{n+1}\cdot {\F}_2[[T]])$
if and only if $\,\,\,\#\{$$x$$\in
$${\mathbb F}_{2}[T]$$\,\,:\,\,{\deg}_T(x)$$<$$n$ and $
f(x)_n=1\}$ is an odd integer, where the notation $f(x)_n$ is as (\ref{e:exp}) since
$f(x)\in {\F}_2[[T]]$.
\end{lemma}
\begin{rem}\label{r:1.1}
We are going to give descriptions of ergodic functions on ${\F}_2[[T]]$ in terms
of Van der Put basis (Theorem \ref{theorem:put}) and in terms of Carlitz basis
(Theorem \ref{theorem:car}). But in applications to computer programming of
cryptography, the proof of this lemma could also provide a method in creating
single cycles modulo $T^k$ for a given positive integer $k$, see section \ref{discussion}
for the discussion.
\end{rem}
\begin{proof} Let $A={\F}_2[T]$ be the polynomial ring over ${\F}_2[T]$, and $A_{<n}=\{x\in
A:\ \deg_T(x)<n\}$ for any non-negative integer $n$.
``Necessity". As $f$ is $1$-Lipschitz, when we consider the
trajectory of $f$ modulo $T^{k}$, we need only consider
$\{x_{0}\operatorname{mod}T^{k},f(x_{0}\operatorname{mod}
T^{k}),\cdots,f^{(i)}(x_{0}\operatorname{mod}T^{k}),\cdots\}$ with
representatives of image elements chosen in $A_{<k}$. If $f$ is
transitive over $\mathbb{F}_{2}[[T]]/(T^{n+1}\cdot {\F}_2[[T]])$,
then there exist $x_{0},x_{1}\in A_{<n}$ such that $f(x_0)=x_1+T^n$.
We consider the trajectory of $f$ modulo $T^{n+1}$ starting with
$x_0$:
\begin{equation}\label{e:2.2}
\begin{array}{cccccccc}
x_{0} & \rightarrow & f(x_0) & \rightarrow &\cdots & \rightarrow
&f^{(2^n-1)}(x_{0})& \rightarrow\\
\rightarrow f^{(2^n)}(x_0) & \rightarrow &f^{(2^n+1)}(x_0)
&\rightarrow &\cdots &\rightarrow &f^{(2^{n+1}-1)}(x_{0})
&\rightarrow\\
\rightarrow x_{0}+T^{n+1} (1+\ast)
&&&&(\mod T^{n+1})&&&
\end{array}
\end{equation}
where $``\ast" \in T\mathbb{F}_{2}[[T]]$, and an element of the
second row is equal to the corresponding element of the first row in
the the column plus a $T^n$, since the map $f$ is measure preserving and transitive
$(\operatorname{mod}T^{n+1})$: $f^{(2^n+i)}(x_0) \equiv f^{(i)}(x_0)+T^n \mod
T^{n+1}$ for $0\le i \le 2^n-1$. We look at the elements from the
left to the right in the first row, if there is an element in $A_{<n}$
other than $x_0$ mapped by $f$ to an element in the set $A_{<n}+T^n$,
then there would be some other element in the set $A_{<n}+T^n$ mapped
to $A_{<n}$, and hence in the second row there would be an element in
$A_{<n}$ mapped by $f$ to an element in $A_{<n}+T^n$. This implies that
the total number of elements in $A_{<n}$ in the trajectory mapped by
$f$ to $A_{<n}+T^n$ is an odd integer, that is, $\#\{$$x$$\in
$${\mathbb F}_{2}[T]$$\,\,:\,\,{\deg}_T(x)$$<$$n$ and $
f(x)_n=1\}$ is an odd integer.
``Sufficiency". By the condition, there must exist $x_0, x_1\in A_{<n}$
such that $f(x_0)=x_1+T^n+{\rm higher\,\,\, terms}$. We consider diagram (\ref{e:2.2}) again.
Since $f$ is transitive modulo $T^n$, the elements of the first row
are distinct and so are the elements of the second row. It also
implies that $f^{(2^n)}(x_0)$ is either equal to $x_0$ or $x_0+T^n$ $(\operatorname{mod}T^{n+1})$.
But if $f^{(2^n)}(x_0)=x_0$, then $\#\{$$x$$\in
$${\mathbb F}_{2}[T]$$\,\,:\,\,{\deg}_T(x)$$<$$n$ and $
f(x)_n=1\}$ would be an even integer. Therefore we must have
$f^{(2^n)}(x_0)=x_0+T^n$ $(\operatorname{mod}T^{n+1})$. And we get $f^{(2^n+i)}(x_0) \equiv
f^{(i)}(x_0)+T^n \mod T^{n+1}$ for $0\le i\le 2^n-1$ by the $1$-Lipschitz
measure-preserving assumption on $f$. Hence all the
elements in the first row and the second row of diagram (\ref{e:2.2})
are distinct, that is, $f$ is transitive modulo $T^{n+1}$.
\end{proof}
\begin{theorem}\label{theorem:put}
\bigskip(ergodicity property) A $1$-Lipschitz function $f:\mathbb{F}_{2}[[T]]\rightarrow
\mathbb{F}_{2}[[T]]$
\begin{equation}\label{e:put}
f(x)=b_0\chi(0,x)+\sum_{\alpha\in\mathbb{F}_{2}[T]\backslash\{0\}}T^{\deg_{T}(\alpha)
}b_{\alpha}\chi(\alpha,x), \quad \text{ with } b_{\a}\in {\F}_2[[T]]
\end{equation}
is ergodic if and only if the following conditions hold
simultaneously:
\begin{itemize}
\item[(1)] $b_{0}\equiv1\ (\operatorname{mod}T)$, $b_{0}+b_{1}\equiv
1+T\ (\operatorname{mod}T^{2}),$ $b_{T}+b_{1+T}\equiv T\ (\operatorname{mod
T^{2})$;
\item[(2)] $\left\vert b_{\alpha}\right\vert_T =1,$ for $\deg_T(\alpha)\geq1$;
\item[(3)] $\sum_{\deg_T(\alpha)=n-1}b_{\alpha}\equiv
T(\operatorname{mod}T^{2}), n\ge 2$.
\end{itemize}
\end{theorem}
\begin{proof}
Since $f$ is a $1$-Lipschitz function, we have
\[
f(x)=B_0\chi(0,x) +\sum_{\a\in{\F}_2\backslash\{0\}}
B_{\a}\chi(\a,x)=b_0\chi(0,x) +\sum_{\a\in{\F}_2[T]\backslash\{0\}}
T^{\deg_T({\a})} b_{\a}\chi(\a,x)
\]
with $b_{\a}\in {\F}_2[[T]]$.
``Necessity". Suppose $f$ is
ergodic. By transitivity modulo $T$, we get
\[
f(0)\equiv 1 \mod T, \quad f(1) \equiv 0 \mod T.
\]
Therefore $b_0=B_0\equiv 1 \ (\mod T)$, and also $f(0)+f(1)\equiv 1
\ (\mod T)$. But $f(0)+f(1)\neq 1(\operatorname{mod}T^{2})$,
otherwise we have $f(0)\equiv 1(\operatorname{mod}T^{2})$,
$f(1)\equiv0(\operatorname{mod}T^{2})$; or
$f(0)\equiv1+T(\operatorname{mod}T^{2})$, $f(1)\equiv
T(\operatorname{mod} T^{2})$, but by the transitivity
$\operatorname{mod}T^{2}$, these two cases can not appear. So we get
\begin{equation}\label{e:2.3}
f(0)+f(1) \equiv1+T(\operatorname{mod}T^{2}), \quad \text{ that
is,}\quad b_{0}+b_{1} \equiv1+T(\operatorname{mod}T^{2}).
\end{equation}
By the transitivity $\operatorname{mod}T^{2}$ and Lemma
\ref{lem:2.1}, we know that
\[
f(0)+f(1)+f(T)+f(1+T) \equiv T^{2}(\operatorname{mod}T^{3}),
\]
which gives us
\[
B_{T}+B_{1+T}\equiv T^{2} (\operatorname{mod}T^{3}).
\]
As $B_{T}= Tb_{T}$, $B_{1+T}= Tb_{1+T}$, we get
\begin{equation*}
b_{T}+b_{1+T}\equiv T (\operatorname{mod}T^{2}).
\end{equation*}
Now consider
\begin{equation*}
\sum_{x\in A_{<n}}f(x) = \sum_{\beta\in A_{<n-1}
}f(\beta+T^{n-1})+\sum_{\beta\in A_{<n-1}}f(\beta)
= \sum_{\deg_T(\alpha)=n-1}B_{\alpha}.
\end{equation*}
Lemma \ref{lem:2.1} gives us
\[
\sum_{x\in A_{<n}}f(x)\equiv T^{n}\operatorname{mod}T^{n+1},
\]
so we put these equations together to get
\[
\sum_{\deg_T(\alpha)=n-1}b_{\alpha} \equiv
T\operatorname{mod}T^{2}.
\]
``Sufficiency". Suppose the three conditions are satisfied, we want
to prove $f$ is transitive on every ${\F}_{2}[[T]]/(T^{n}\cdot{\F}_2[[T]])$ for all
$n\in\mathbb{N}$. But this is just to apply Lemma \ref{lem:2.1} on
the induction process for $n$, the first condition gives the first
step of the induction.
\end{proof}
\section{$1$-Lipschitz Functions over ${\F}_r[[T]]$ and Carlitz
Expansions}\label{carlitz}
We first recall some useful formulas in function field
arithmetic, with all the details and expositions in Chapter 3 of \cite{Go}. Let
$A=\mathbb{F}_{r}[T]$ ($r$ is a power of the prime number $p$) with the normalized absolute value
$|\cdot|_T$ such that $|T|_T= 1/r$. The completion of $A$ with respect to this absolute value is
$\hat{A}={\F}_r[[T]]$.
\begin{defn}
We set the following notations:
\begin{itemize}
\item
$[i]=T^{r^{i}}-T$, where $i$ is a positive integer;
\item $L_i=1$ if
$i=0$; and $L_i=[i]\cdot [i-1]\cdots[1]$ \ if $i$ is
a positive integer;
\item $D_i=1$ if $i=0$; and $D_i=[i]\cdot [i-1]^r\cdots [1]^{r^{i-1}}$
\ if $i$ is a positive integer;
\item for any non-negative integer
$n=n_0+n_1\, r+\cdots+n_s\, r^s$, $0\le n_j\le r-1$ for $0\le j\le s$, the $n$-th Carlitz factorial
$\Pi(n)$ is defined by
\[
\Pi(n)=\pd_{j=0}^s D_j^{n_j};
\]
\item
$ e_d(x)=
\begin{cases}
x, &\quad\text{ if } d=0,\\
\prod\limits_{\a\in A,\,\deg_{T}(\alpha )<d}(x-\alpha), &\quad\text{
if $d$ is a positive integer;}
\end{cases}
$
\item $E_{i}(x)=e_{i}(x)/{D_{i}}$, for any non-negative integer $i$;
\item $G_{n}(x)=\prod\limits_{i=0}^{s}(E_{i}(x))^{n_{i}}$, $n=n_{0}+n_{1}r+\cdots
+n_{s}r^{s}$ non-negative integers, $0\leq n_{i}<r$;
\item
$G_{n}^{\prime}(x)=\prod\limits_{i=0}^{s}G_{n_{i}r^{i}}^{\prime}(x)$,
where $G^{\prime}_{n_i r^i}=
\begin{cases}
\left(E_i(x)\right)^{n_i}, &\quad\text{ if } 0\le n_i<r-1,\\
\left(E_i(x)\right)^{n_i}-1, &\quad\text{ if } n_i=r-1.
\end{cases}
$
\end{itemize}
\end{defn}
The polynomials $G_n(x)$ and $G'_n(x)$ are called Carlitz
polynomials.
\begin{prop}[\cite{Ca}]\label{prop:3.1}
The following formulas hold for the Carlitz polynomials
\begin{itemize}
\item $G_{m}(t+x)=\sum\limits_{\substack{k+l=m\\k,\,l\geq0}}\binom{m}{k
G_{k}(t)G_{l}(x),$ $t,x\in\mathbb{F}_{2}[[T]]$.
\item $G_{m}^{\prime}(t+x)=\sum\limits_{\substack{k+l=m\\k,\,l\geq0}}\binom
{m}{k}G_{k}(t)G_{l}^{\prime}(x)$.
\end{itemize}
Orthogonality property of $\{G_{n}(x)\}_{n\geq0}$ and
$\{G_{n}^{\prime }(x)\}_{n\geq0}$:
\begin{itemize}
\item For any $s<r^{m}$, $l$ an arbitrary non-negative integer$,$
\begin{equation}\label{e:3.0}
\sum_{\deg_T(\alpha)<m}G_{l}(\alpha)G_{s}^{\prime}(\alpha)=
\begin{cases}
0, &\quad\text{ if } l+s\ne r^m-1;\\
(-1)^m, &\quad\text{ if } l+s = r^m-1.
\end{cases}
\end{equation}
\end{itemize}
\end{prop}
The polynomials $G_{n}(x)$ and $G_{n}^{\prime}(x)$ map $A$ to $A$.
And it is well known that $\{G_{n}(x)\}_{n\ge 0}$ is an orthonormal
basis of the space $C(\mathbb{F}_{r}[[T]], \mathbb{F}_{r}((T)))$ of
continuous functions from ${\F}_r[[T]]$ to ${\F}_r((T))$, that is,
every $T$-adic continuous function can be written as:
\begin{equation*}
f(x)=\sum_{n=0}^{\infty}a_{n}G_{n}(x),\quad\text{ where $a_n\in
{\F}_r((T))$, \,\, and } a_{n} \rightarrow 0,\text{ as }
n\rightarrow\infty,
\end{equation*}
with the sup-norm $||f||=\max\limits_{n}\{|a_n|_T\}$. Moreover, the
expansion coefficient $a_n$ can be calculated as:
\begin{equation}\label{e:car}
a_{n} =(-1)^{m}\sum_{\deg_T(\alpha)<m}G_{r^{m}-1-n}^{\prime
}(\alpha)f(\alpha),\text{ for any integer such that }r^{m}>n.
\end{equation}
Following Wagner~\cite{Wa}, we define a new sequence of polynomials
$\{H_n(x)\}_{n\geq 0}$ by
\begin{equation*}
\begin{split}
&\ H_0(x)=1,\qquad \text{ and } \\
&\ H_n(x)=\frac{\Pi(n+1)G_{n+1}(x)}{\Pi(n)\, x} \qquad\text{ for
$n\geq 1$. }
\end{split}
\end{equation*}
Then we get
\begin{lemma}[\cite{Wa}]\label{lem:3.1}
$\{H_n(x)\}_{n\geq 0}$ is an orthonormal basis of $C({\F}_r[[T]], {\F}_r((T)))$.
\end{lemma}
To study the $1$-Lipschitz functions over $\hat{A}={\F}_r[[T]]$, we recall the interpolation polynomials
introduced by Amice \cite{Am}. These polynomials $\{Q_n(x)\}_{n\ge 0}$ are constructed from
which is called by Amice the ``very well distributed sequence" $\{u_n\}_{n \ge 0}$ with $ u_n\in A$:
\begin{equation}\label{e:3.1}
Q_n(x)=
\begin{cases}
1, & \text{ if } n=0,\\
\dfrac{(x-u_0)(x-u_1)\cdots (x-u_{n-1})}{(u_n-u_0)(u_n-u_1)\cdots (u_n-u_{n-1})}, &\text{ if } n\ge 1.
\end{cases}
\end{equation}
We choose the sequence $\{u_n\}_{n\ge 0}$ in the following way
such that $u_n\ne 0$ for any $n\ge 0$. Let
$S=\{\a_0,\a_1,\cdots,\a_{r-1}\}$ be a system of representatives of
$A/(T\cdot A)$, and assume that $\a_0=T$ (thus $0\not\in S$). Then
any element $x\in {\F}_r((T))$ can be uniquely written as
\[
x=\sum_{k \ge k_0}^\infty \b_k\,T^k
\]
where $\b_k\in S$ and $k_0\in \Z$, with $x$ in $\hat{A}$ if and only if
$\b_k=0$ for all $k<0$. To each non-negative integer
$n=n_0+n_1q+\cdots+n_s r^s$ in $r$-digit expansion, we
assign the element
\[
u_n=\a_{n_0}+\a_{n_1}T+\cdots+\a_{n_s}T^s.
\]
We have
\begin{theorem}[\cite{Am}]\label{thm:Am}
The interpolation polynomials $\{Q_n(x)\}_{n\ge 0}$ defined above is
an orthonormal basis of $C({\F}_r[[T]],{\F}_r((T)))$. That is, any
continuous function $f(x)$ from ${\F}_r[[T]]$ to ${\F}_r((T))$ can
be written as
\begin{equation}\label{e:3.2}
f(x)=\sum\limits_{n=0}^\infty a_n Q_n(x),
\end{equation}
where $a_n\to 0$ as $n\to \infty$, and the sup-norm of $f$ is given
by $||f||=\max\limits_n\{|a_n|_T\}$.
\end{theorem}
Since $Q_n(u_n)=1$ for all $n$, and $Q_m(u_n)=0$ for $m>n$ by the
equation (\ref{e:3.1}), we see that the expansion coefficients $a_n$
can be deduced by the following induction formula
\begin{equation}\label{e:3.3}
\begin{array}{ll}
&a_0=f(0)\\
&a_n=f(u_n)-\sum\limits_{j=0}^{n-1} a_j Q_j(u_n) \quad\text{ for $n\ge 1$ }.
\end{array}
\end{equation}
Equation (\ref{e:3.2}) is valid not only for continuous functions on
$\hat{A}={\F}_r[[T]]$, but also for any function $f$ from
$A\backslash \{0\}$ to ${\F}_r((T))$. Elements of the sequence
$\{u_n\}$ constitute the set $A\backslash\{0\}$, so the summation of equation (\ref{e:3.2}) is a
finite sum for $x\in A\backslash\{0\}$. The
element $0$ is excluded because of the way the sequence $\{u_n\}$ is
chosen. If the function $f$ is
continuous on $\hat{A}\backslash\{0\}$, then $f$ is certainly
determined by the values of $f$ at the points of $A\backslash\{0\}$.
Suppose $\{R_n(x)\}_{n \ge 0}$ is any orthonormal basis of $C(\hat{A},{\F}_r((T)))$
consisting of polynomials in the variable $x$ with $\deg_T(R_n)=n$. Then we have
\begin{equation}\label{e:3.4}
Q_n(x)=\sum\limits_{j=0}^n \gamma_{n,j} R_j(x),
\end{equation}
where $\gamma_{n,j}\in\hat{A}$ for all $n,j$, and
$\gamma_{n,n}\in {\hat{A}}^{\times}$.
For any positive integer $n$, write $n=n_0+n_1r+\cdots+n_wr^w$ in
$r$-digit expansion, with $n_w\neq 0$,
\begin{itemize}
\item denote $\nu(n)$ the largest integer such
that $r^{\nu(n)}|n$;
\item $l(n)=l_r(n)=n_wr^w$.
\end{itemize}
\begin{lemma}\label{lem:3.2}
Let $n$ be a positive integer, then
\[
\frac{\Pi(n-1)}{\Pi(n)}=\frac{1}{L_{\nu(n)}}
\]
\end{lemma}
\begin{proof}
Straightforward computation.
\end{proof}
\begin{lemma}[\cite{Ya}]\label{lem:3.3}
If a function $f:{\F}_r[[T]]\backslash\{0\} \to {\F}_r((T))$ can be
expressed as $f(x)=\sum\limits_{n=0}^\infty a_n H_n(x)$ for all
$x\in {\F}_r[[T]], x\neq 0$, that is, the summation converges to
$f(x)$ for $x\neq 0$, then the sequence $\{a_n\}_{n\geq 0}$ is
determined by the values $f(x)$ for all $x\in
{\F}_r[T]\backslash\{0\}$. More precisely, for any non-negative
integer $n$, we choose an integer $w$ such that $n<r^w-1$ and set
$S=\{\alpha\in {\F}_r[T]:\,\, \deg_T(\a)<w, \a\neq 0\}$, then $a_n$ is
determined by the values of $f$ at the points of $S$:
\begin{equation}\label{e:coefficient}
a_n=\dfrac{(-1)^w}{L_{\nu(n+1)}}\sum\limits_{\a\in S}\a f(\a)G'_{r^w-2-n}(\a).
\end{equation}
\end{lemma}
\begin{proof}
This is a refined statement of Lemma 5.5 of \cite{Ya}. By definition and Lemma \ref{lem:3.2},
\[
H_n(x)=\dfrac{\Pi(n+1) G_{n+1}(x)}{\Pi(n) x}= L_{\nu(n+1)}\dfrac{G_{n+1}(x)}{x}
\]
for $n\ge 0$, thus
\[
xf(x)=\sum\limits_{n=0}^\infty a_n L_{\nu(n+1)}G_{n+1}(x)
\]
for any $x\neq 0$ in $\hat{A}$. Since $G_{n+1}(0)=0$, we sum up all elements $\a\in S$ in
the above equation, and apply the equation (\ref{e:3.0}) of orthogonality property to get
for any non-negative integer $m<r^w-1$
\[
\begin{array}{ll}
&\sum\limits_{\a\in S}\a f(\a)G'_{m}(\a)=\sum\limits_{\a\in S}\sum\limits_{n\ge 0}
a_n L_{\nu(n+1)} G_{n+1}(\a)G'_m(\a)\\
=&\sum\limits_{\deg_T(\a)<w}\sum\limits_{n\ge 0} a_n L_{\nu(n+1)}G_{n+1}(\a)G'_m(\a)\\
=&\sum\limits_{n\ge 0}a_n L_{\nu(n+1)}\sum\limits_{\deg_T(\a)<w}G_{n+1}(\a)G'_m(\a)\\
=&(-1)^w a_{r^w-2-m} L_{\nu(r^w-1-m)}.
\end{array}
\]
Notice that the summations on $n$ are actually finite sums, thus we can change the order of
summations on $\a$ and on $n$. Therefore we get the formula (\ref{e:coefficient}), and the
conclusion.
\end{proof}
\begin{lemma}\label{lem:3.4}
We have
\begin{itemize}
\item[(1)]
$
|L_n|_T=r^{-n}=|T|_T^{n}
$
for any non-negative integer $n$;
\item[(2)] For any non-negative integer $n$, $\nu(n)\le \lfloor \log_r (n)\rfloor$.
\end{itemize}
\end{lemma}
\begin{proof}
Immediate from definition.
\end{proof}
Denote
\[
((i_1,i_2,\cdots,i_s))=\frac{(i_1+i_2+\cdots+i_s)!}{i_1!\:
i_2!\:\cdots\:i_s!}
\]
for any integers $i_1,i_2,\cdots,i_s\geq 0$. We
have the following assertion about the multinomial
numbers by Lucas \cite{Lu}:
\begin{lemma}[Lucas]\label{lucaslemma}
For non-negative integers $n_0,n_1,\cdots,n_s$,
\begin{equation}\label{e:lucaslemma}
((n_0,n_1,\cdots,n_s)) \equiv \pd_{j\geq 0}
((n_{0,j},n_{1,j},\cdots,n_{s,j})) \mod p
\end{equation}
where $n_i=\sum\limits_{j\geq 0}n_{i,j}\, r^j$ is the
$r$-digit expansion for $i=0,1,\cdots,s$.
\end{lemma}
\begin{rem}\label{r:3.1}
Lemma~\ref{lucaslemma} is useful when $s=1$.
In this case formula (\ref{e:lucaslemma}) is
expressed in the form: let $n=\sum\limits_j n_j\, r^j$ and
$k=\sum\limits_j k_j\, r^j$ be $r$-digit expansion for
non-negative integers $n$ and $k$, then
\[
\binom nk \equiv \pd_{j\geq 0} \binom{n_j}{k_j} \mod p.
\]
\end{rem}
\begin{lemma}\label{lem:3.5}
Let $f(x)=\sum\limits_{n=0}^\infty a_n H_n(x)$ be a continuous
function from $\hat{A}\backslash\{0\}$ to ${\F}_r((T))$ ( this
implies that the series converges for any $x\in
\hat{A}\backslash\{0\}$ ). Suppose that $|f(x)|_T\le 1$ for any $x\in
\hat{A}\backslash\{0\}$. Then $|a_n|_T\le 1$ for $n\ge 0$.
\end{lemma}
\begin{proof}
Since $f$ is continuous, it is determined by the values of $f$ at
the points in $A\backslash\{0\}$. From the explanation in the
paragraph after Theorem \ref{thm:Am}, we see that $f$ can be written
as
\begin{equation}\label{e:3.5}
f(x)=\sum\limits_{n=0}^\infty b_n Q_n(x).
\end{equation}
Induction formula (\ref{e:3.3}) and the condition that $|f(x)|_T\le 1$
for any $x\in \hat{A}\backslash\{0\}$ imply $|b_n|_T\le 1$ for all
$n$.
Now we fix a non-negative integer $w$ and let $N\ge r^w-1$ be an integer.
Then for any $\a\neq 0$ with $\deg_T(\a)<w$, we can write the right hand of equation (\ref{e:3.5})
as a finite sum:
\begin{equation*}\label{e:3.6}
f(\a)=\sum\limits_{n=0}^N b_n Q_n(\a).
\end{equation*}
In the above equation, substitute $Q_n(\a)$ by the equation (\ref{e:3.4}) with $R_n(x)=H_n(x)$ for
any non-negative integer $n$, we get
\[
f(\a)=\sum\limits_{j=0}^N\left(\sum\limits_{n=j}^N b_n\gamma_{n,j}\right) H_j(\a)
=\sum\limits_{j=0}^N a_j H_j(\a),
\]
for any $\a\neq 0$ with $\deg_T(\a)<w$. Therefore for any non-negative integer
$j<r^w-1$, Lemma \ref{lem:3.3} implies that
\[
a_j=\sum\limits_{n=j}^N b_n \gamma_{n,j}.
\]
Hence we have $|a_j|_T\le 1$ for $j<r^w-1$. Since $w$ is arbitrary, we get the conclusion.
\end{proof}
\begin{theorem}\label{thm:3.2}
A continuous function $f(x)=\sum\limits_{n=0}^{\infty} a_n G_n(x)$
from ${\F}_r[[T]]$ to ${\F}_r((T))$ is $1$-Lipschitz if and only if
$|a_n|_T\le |T|_T^{\lfloor \log_r (n)\rfloor}$ for $n\ge 1$ and $|a_0|_T\le 1$.
\end{theorem}
\begin{proof}
The proof is very similar to that on the $C^n$ functions
over positive characteristic local rings \cite{Ya}. We can calculate
for $y_1\neq 0$ by using the equation of Proposition
\ref{prop:3.1},
\begin{align}
\begin{split}
\frac 1{y_1}(f(y_1+x)-f(x))
&=\sum_{n_0=0}^\infty a_{n_0}\frac 1{y_1}
(G_{n_0}(y_1+x)-G_{n_0}(x)) \\
&=\sum_{n_0=0}^\infty\sum_{j_1=0}^\infty\dbinom{n_0+j_1+1}
{j_1+1}\frac{a_{n_0+j_1+1}}{L_{\nu(j_1+1)}} H_{j_1}(y_1)
G_{n_0}(x),
\end{split}\label{e:3.6}
\end{align}
The order of summations can be exchanged since the sequence of the
terms in the summation tends to $0$ as $j_1+n_0 \to \infty$ for any
$y_1\neq 0$, and any $x$ in ${\F}_r[[T]]$.
``Sufficiency". The absolute values of $G_{n_0}(x)$, $H_{j_1}(y_1)$,
and the binomial numbers of equation (\ref{e:3.6}) are all less than
or equal to $1$. By Lemma \ref{lem:3.4} and the condition on
$a_n$, we can estimate that $|a_{n_0+j_1+1}/L_{\nu(j_1+1)}|_T\le 1$.
Therefore the function $f$ is $1$-Lipschitz.
``Necessity". Suppose $f$ is $1$-Lipschitz. The function
$\Psi_1f(x,y_1) =\frac 1{y_1}(f(y_1+x)-f(x))$ is continuous on
${\F}_r[[T]]\times ({\F}_r[[T]]\backslash\{0\})$. Since $\Psi_1f(x,y_1)$ is
continuous with respect to
$x\in {\F}_r[[T]]$, we get a function
\[
F_{n_0}(y_1)=
\sum_{j_1=0}^\infty\dbinom{n_0+j_1+1}{j_1+1}\frac{a_{n_0+j_1+1}}{L_{\nu(j_1+1)}}
H_{j_1}(y_1)
\]
for every $n_0\ge 0$. We have $|F_{n_0}(y_1)|_T\le 1$ for any
$y_1\in {\F}_r[[T]]\backslash\{0\}$, since $f$ is $1$-Lipschitz. And $F_{n_0}(y_1)$
is continuous on $\hat{A}\backslash\{0\}$. Then Lemma \ref{lem:3.5}
implies that
\begin{equation}\label{e:3.7}
\left|
\dbinom{n_0+j_1+1}{j_1+1}\frac{a_{n_0+j_1+1}}{L_{\nu(j_1+1)}}\right|_T\le 1,
\end{equation}
for any $n_0\ge 0, j_1\ge 0$.
It is clear that $|a_0|_T\le 1$. For any integer $n\ge 1$, we write
$n=m_0+m_1 r+\cdots + m_w r^w$ in $r$-digit expansion, where $m_w\neq 0$.
We choose non-negative integers $n_0, j_1$ by
\[
j_1+1=l(n)=m_w r^w, \quad \text{ and } \quad n_0 = n-j_1-1.
\]
Then from Lucas formula (\ref{e:lucaslemma}) and Lemma \ref{lem:3.4}, we get
\[
\dbinom{n_0+j_1+1}{j_1+1} =1, \quad\text{ and }\quad |L_{\nu(j_1+1)}|_T=r^{-w}=|T|_T^{\lfloor \log_r (n)\rfloor}.
\]
And from equation (\ref{e:3.7}), we see that
\[
|a_n|_T\le |T|_T^{\lfloor \log_r (n)\rfloor} \quad\text{ for } n\ge 1.
\]
\end{proof}
\section{Ergodic Functions over ${\F}_2[[T]]$ and Carlitz
Expansions}\label{ecarlitz}
In this section, we take $r=2$. And the Carlitz polynomials $G_n(x)$
and $G'_n(x)$ are defined for $x\in {\F}_2[[T]]$ with coefficients
in ${\F}_2((T))$. We will prove the ergodicity property of functions
over ${\F}_2[[T]]$ by translating the conditions of ergodicity under
Van der Put basis to Carlitz basis. At first we notice that the
polynomials $G_{n}(x)$ and $G_{n}^{\prime}(x)$ have the following
special values:
\begin{itemize}
\item $G_{0}(x)=1$ for any $x$, $G_{n}(0)=0$, if $n\geq 1$;
\item $G_{1}(x)=x$, $G_{n}(1)=0$, if $n\geq 2 $;
\item $G_2(T)=G_2(1+T)=1$, $G_3(T)=T$, $G_3(1+T)=1+T$, and $G_n(T)$ $=$ $G_n(1+T)$ $=0$ if $n\ge
4$;
\item $G_{0}^{\prime}(\alpha)=1$, for any
$\alpha\in\mathbb{F}_{r}[[T]]$.
\end{itemize}
We also recall that $A={\F}_2[T]$, $\hat{A}={\F}_2[[T]]$, $A_n=\{\a\in A:\,\, \deg_T(\a)=n\}$, and
$A_{\le n}=\{\a\in A:\,\, \deg_T(\a)\le n\}$
for any non-negative integer $n$. Moreover, we notice that a function $f\in C(\hat{A},{\F}_2((T)))$
is measure-preserving if and only if
\begin{equation}\label{e:4.1}
\left|\frac{1}{y}\left(f(x+y)-f(x)\right)\right|_T=1
\text{ for any $y\in \hat{A}\backslash\{0\}$ and any $x\in\hat{A}$}.
\end{equation}
\begin{theorem}\label{theorem:car}
\bigskip(ergodicity property) A $1$-Lipschitz function $f:\mathbb{F}_{2}[[T]]\rightarrow
\mathbb{F}_{2}[[T]]$
\[
f(x)=\sum_{n=0}^{\infty}a_{n}G_{n}(x)
\]
is ergodic if and only if the following conditions are satisfied
\begin{itemize}
\item[(1)] $a_{0}\equiv1(\operatorname{mod}T)$, $a_{1}\equiv1+T(\operatorname{mod}
T^{2}),$ $a_{3}\equiv T^{2}(\operatorname{mod}T^{3})$;
\item[(2)] $\left\vert a_{n}\right\vert_T <|T|_T^{\lfloor \log_2(n)\rfloor}=2^{-\lfloor \log_2 (n)\rfloor}$, for $n\geq
2$;
\item[(3)] $a_{2^{n}-1}\equiv T^{n}(\operatorname{mod}T^{n+1})$ for $n\ge 2$.
\end{itemize}
\end{theorem}
\begin{proof}
We have
$f(x)=\sum\limits_{n=0}^{\infty}a_{n}G_{n}(x)=\sum\limits_{\alpha
\in\mathbb{F}_{2}[T]}B_{\alpha}\chi(\alpha,x)$. At first we translate
the conditions (1) and (3) of Theorem \ref{theorem:put} to those on the coefficients of the
Carlitz basis. In the expansion (\ref{e:put}) of Theorem \ref{theorem:put}, we also
use the notation $B_{\a}=T^{\deg_T(\a)} b_{\a}$ for $\a\in {\F}_2[T]\backslash\{0\}$.
\begin{itemize}
\item[(1)] $B_{0}=b_0\equiv1(\operatorname{mod}T)$:
as $B_{0}=f(0)=\sum_{n=0}^{\infty}a_{n}G_{n}(0)$, this condition is equivalent to
\[
a_0=\sum_{n=0}^{\infty}a_{n}G_{n}(0)=f(0)=B_0 \equiv1(\operatorname{mod}T).
\]
\item[] $B_{0}+B_{1}=b_0 + b_1\equiv1+T(\operatorname{mod}T^{2})$:
from $B_{0}+B_{1}=f(0)+f(1)=\sum_{n=0}^{\infty}a_{n}G_{n}(0)+\sum_{n=0}^{\infty
}a_{n}G_{n}(1)$, this condition is equivalent to
\begin{align*}
a_1&\equiv a_0+a_0+a_1\equiv f(0)+f(1)\\
&\equiv B_0+B_1\equiv 1+T (\operatorname{mod} T^2).
\end{align*}
\item[] $b_{T}+b_{1+T}\equiv T(\operatorname{mod}T^{2})$:
this is the same as $B_{T}+B_{1+T}\equiv T^2(\operatorname{mod}T^{3})$. From the explicit calculation
\begin{align*}
B_{T}+B_{1+T} & =(f(T)-f(0))+(f(1+T)-f(1))\\
& =\sum_{n=0}^{\infty}a_{n}(G_{n}(T)-G_{n}(0))+\sum_{n=0}^{\infty}a_{n
(G_{n}(1+T)-G_{n}(1))\\
& =a_{3},
\end{align*}
we see that this condition is equivalent to $a_3\equiv T^2 (\operatorname{mod} T^3)$.
\item[(3)] The third condition of Theorem \ref{theorem:put} (ergodicity property under Van der Put basis)
is $\sum_{\a\in A_{n-1}}b_{\a}\equiv
T(\operatorname{mod}T^{2})$, which is equivalent to
$\sum_{\a\in A_{n-1}}B_{\a}\equiv
T^{n}(\operatorname{mod}T^{n+1})$. We can calculate
\[
\begin{array}{ll}
&\sum\limits_{\a\in A_{n-1}}B_{\a}=\sum\limits_{\b\in A_{\le n-2}}
(f(\b + T^{n-1})-f(\b))\\
=&\sum\limits_{\b\in A_{\le n-2}}\sum\limits_{m=0}^\infty\left(a_m G_m(\b+T^{n-1})-a_m G_m(\b)\right)\\
=&\sum\limits_{\b\in A_{\le n-2}}\sum\limits_{m=1}^\infty a_m \sum\limits_{j=0}^{m-1}
\binom{m}{j} G_j(\b) G_{m-j}(T^{n-1})\\
=&\sum\limits_{m=1}^\infty a_m \sum\limits_{j=0}^{m-1} \binom{m}{j} G_{m-j}(T^{n-1})
\sum\limits_{\b\in A_{\le n-2}} G_j(\b) G'_0(\b)\\
=&\sum\limits_{m=1}^\infty a_m \sum\limits_{j=0}^{m-1} \binom{m}{j} G_{m-j}(T^{n-1})
(-1)^{n-1}\delta_{j,\,2^{n-1}-1}\\
=&a_{2^n-1}.
\end{array}
\]
The last equality of the above equations holds because
$\binom{m}{2^{n-1}-1}\neq 0$ only when $m=(2^{n-1}-1)+l\cdot
2^{n-1}$ for some integer $l\ge 0$ (due to the Lucas formula
(\ref{e:lucaslemma})) and $m-1\ge 2^{n-1}-1$, and we have the
special values of Carlitz polynomials:
\[
G_{l\,\cdot\, 2^{n-1}}(T^{n-1})=
\begin{cases}
1, &\quad\text{if } l=1,\\
0, &\quad\text{if } l>1.
\end{cases}
\]
The order of summation can be exchanged because the function $f$ is
assumed to be $1$-Lipschitz, thus $a_n\to 0$ as $n\to \infty$. Therefore
the third condition of Theorem \ref{theorem:put} is equivalent to
$a_{2^n-1}\equiv T^n \,(\operatorname{mod} T^{n+1})$ for $n\ge 2$.
\end{itemize}
Now we scrutinize equation (\ref{e:3.6}):
\begin{align}
\begin{split}
&\frac 1{y_1}(f(y_1+x)-f(x))\\
=&\sum_{n_0=0}^\infty\sum_{j_1=0}^\infty\dbinom{n_0+j_1+1}
{j_1+1}\frac{a_{n_0+j_1+1}}{L_{\nu(j_1+1)}} H_{j_1}(y_1)
G_{n_0}(x) \\
= &\, a_1+\sum_{j_1=1}^\infty \frac{a_{j_1+1}}{L_{\nu(j_1+1)}} H_{j_1}(y_1)
G_{0}(x)\\
& + \sum_{n_0=1}^\infty\sum_{j_1=0}^\infty\dbinom{n_0+j_1+1}
{j_1+1}\frac{a_{n_0+j_1+1}}{L_{\nu(j_1+1)}} H_{j_1}(y_1)
G_{n_0}(x) \label{e:4.2}\\
\end{split}
\end{align}
for $x\in \hat{A}$ and $y_1\in \hat{A}\backslash\{0\}$.
``Sufficiency". Assume the three conditions of the theorem are
satisfied. Then we know that $a_1\equiv 1 (\operatorname{mod} T)$ and we can deduce from
equation (\ref{e:4.2}) that
\begin{equation}\label{e:4.3}
\frac{1}{y_1}\left(f(y_1+x)-f(x)\right)=1+Th(y_1,x),
\end{equation}
where $h(y_1,x)$ is a continuous function from $\left(\hat{A}\backslash\{0\}\right)
\times \hat{A}$ to $\hat{A}$. Therefore $| \frac 1{y_1}$ $(f(y_1+x)-f(x))|_T = 1$
for any $y_1\in \hat{A}\slash\{0\}$, and any $x\in\hat{A}$. Hence the function $f$ is
measure preserving and Theorem \ref{theorem:mp} implies that the condition (2) of Theorem
\ref{theorem:put}, as well as conditions (1) and (3), is satisfied. Therefore the function
$f$ is ergodic.
``Necessity". Assume that the $1$-Lipschitz function $f$ is ergodic. Then by Theorem \ref{theorem:put}
and the discussion at the beginning of this proof, we see that conditions (1) and (3) are satisfied.
We do the same calculation to get equation (\ref{e:4.3}), where
\[
\begin{split}
h(y_1,x)= &\, \frac{1}{T}\Big(a_1-1+\sum_{j_1=1}^\infty \frac{a_{j_1+1}}{L_{\nu(j_1+1)}} H_{j_1}(y_1)
G_{0}(x)\\
&+ \sum_{n_0=1}^\infty\sum_{j_1=0}^\infty\dbinom{n_0+j_1+1}
{j_1+1}\frac{a_{n_0+j_1+1}}{L_{\nu(j_1+1)}} H_{j_1}(y_1)
G_{n_0}(x)\Big).
\end{split}
\]
As the function $f$ is assumed to be ergodic, $h(y_1,x)$ is a continuous function
from $\left(\hat{A}\backslash\{0\}\right)
\times \hat{A}$ to $\hat{A}$. Therefore we can apply the same proof as Theorem \ref{thm:3.2}
to get the condition (2): $\left\vert a_{n}\right\vert_T <|T|_T^{\lfloor \log_2(n)\rfloor}=
2^{-\lfloor \log_2 (n)\rfloor}$, for $n\geq
2$.
\end{proof}
\begin{example}
In terms of Carlitz basis, the simplest ergodic function on ${\F}_2[[T]]$ would be
\[
f(x)=1 + (1+T)x + \sum\limits_{n=2}^\infty T^n G_{2^n-1}(x).
\]
By Theorem \ref{theorem:car}, we see that an ergodic $1$-Lipschitz function
$f:\mathbb{F}_{2}[[T]]\rightarrow
\mathbb{F}_{2}[[T]]$ can not be a polynomial function.
\end{example}
\section{Discussion}\label{discussion}
Let $U=\{0,1\}$ and let $k$ be a positive integer. Then $U^k$ is the
set of $k$-tuples consisting of $0$ and $1$. In applications, we
need to study the transformations of $U^k$ for large integers $k$.
For convenience, we can make any element in $U^k$ into an infinite
sequence of $0$ and $1$ by adding all $0$'s after the $k$-th
component of a $k$-tuple:
\[
\begin{array}{rcl}
U^k &\hookrightarrow &U^{\infty}\\
(a_0, a_1, \cdots, a_{k-1}) &\mapsto &(a_0, a_1, \cdots, a_{k-1}, 0,
0, \cdots)
\end{array}
\]
where $U^{\infty}$ is the set of infinite sequences of $0$ and $1$.
So we can study transitive transformations of $U^k$ by analyzing transformations of $U^{\infty}$
and interpreting the results on $U^k$.
Now there are two ways of interpreting the elements of $U^{\infty}$.
One of them is to view an element of $U^{\infty}$ as a $2$-adic
integer, which leads to the work of V. Anashin et al
\cite{An1}\cite{An4} on the theory of measure-preservation and
ergodicity over ${\Z}_2$. The other is to view an element of
$U^{\infty}$ as a power series over the field ${\F}_2$, which leads
to the study of the theory of measure-preservation and ergodicity
over ${\F}_2[[T]]$ in this paper. Therefore we have two different ways to
construct transformations of $U^{\infty}$ which can give rise to
transitive maps from $U^k$ to itself for every positive integer $k$.
Thus it is natural to make comparisons between them, especially when
algorithms are implemented in computer programming, which we will do in the future.
In Theorem \ref{thm:An1}, Theorem \ref{thm:An4}, and Theorem
\ref{thm:An4.2} about the theory over ${\Z}_p$, or Theorem
\ref{theorem:put} and Theorem \ref{theorem:car} about the theory
over ${\F}_2[[T]]$, the formulae of ergodic functions are explicitly
given in respective cases, hence we can write algorithms to
calculate the values of these ergodic functions at any given
elements. While it is certainly important to analyze the efficiency
of these algorithms, it is also worth noticing that the method used
in the proof of Lemma \ref{lem:2.1} could also provide an algorithm
of constructing single cycles of ${\F}_2[[T]]$ modulo $T^{n+1}$ for
any given integer $n\ge 0$. Let $S$ be the set of data
$\{a(k,j)\}_{k\ge 1,\, 0\le j\le 2^k-1}$, where $a(k,j)\in \{0, 1\}$
for each $k$ and each $j$. Given a datum $\{a(k,j)\}_{k,j}$ in $S$
and an integer $n\ge 1$ (in fact the integer $n$ should be large),
we can produce a sequence $\{x_j\}_{0\le j\le 2^{n+1}-1}$ by the
following pseudo codes:
\begin{itemize}
\item[]$x_0=0, x_1=1;$
\item[]{\bf for} $k=1$ {\bf to} $n$ {\bf do}
\item[]$\quad$ {\bf for} $j=0$ {\bf to} $2^{k}-1$ {\bf do}
\item[]$\quad\quad$ {\bf if} $a(k,j)=1$ {\bf then} $x_j=x_j+T^{k}$;
\item[]$\quad$ {\bf end for};
\item[]$\quad$ {\bf for} $j=2^{k}$ {\bf to} $2^{k+1}-1$ {\bf do}
\item[]$\quad\quad$ $x_j=x_{j-2^{k}} + T^{k}$;
\item[]$\quad$ {\bf end for};
\item[]{\bf end for};
\end{itemize}
By the process of the proof of Lemma \ref{lem:2.1}, the sequence $\{x_j\}_{0\le j\le 2^{n+1}-1}$ gives a
transitive transformation $f$ of ${\F}_2[[T]]/(T^{n+1}\cdot {\F}_2[[T]])$ such that
$f \mod T^k$ is transitive for every $k=1, 2, \cdots, n+1$,
by defining $f(x_j)=x_{j+1}$ for $0\le j\le 2^{n+1}-2$ and $f(x_{2^{n+1}-1})=x_0$.
From the proof of Lemma \ref{lem:2.1}, we can also see that the above construction gives
all ergodic $1$-Lipschitz functions from ${\F}_2[[T]]$ to itself,
hence the data $\{a(k,j)\}_{1\le k\le n,\, 0\le j\le 2^k-1}$
give rise to all transitive transformations of ${\F}_2[[T]]/(T^{n+1}\cdot {\F}_2[[T]])$ which are
also transitive modulo $T^k$ for every $k=1, 2, \cdots, n+1$.
It might be useful to generalize the results of this paper to the ergodic theory over
${\F}_r[[T]]$ for $r$ being a power of a prime number. To do this, the main difficulty
is that we don't have a description of ergodicity as simple as Lemma \ref{lem:2.1}.
So this problem will have to be left for studies in the future.
|
\section{Introduction}\label{sec:intro}
In \cite{logsin1} the classical log-sine integrals and their extensions were
used to develop a variety of results relating to higher and multiple Mahler
measures \cite{boyd,klo}. The utility of this approach was such that we
continue the work herein. Among other things, it allows us to tap into a rich
analytic literature \cite{lewin2}. In \cite{logsin-evaluations} the
computational underpinnings of such studies are illuminated. The use of related
integrals is currently being exploited for multi-zeta value studies
\cite{ona11}. Such evaluations are also useful for physics \cite{lsjk}:
log-sine integrals appeared for instance in the calculation of
higher terms in the $\varepsilon$-expansion of various Feynman diagrams
\cite{davydychev-eps, kalmykov-eps}. Of particular importance are the log-sine
integrals at the special values $\pi/3$, $\pi/2$, $2\pi/3$, and $\pi$. The
log-sine integrals also appear in many settings in number theory and analysis:
classes of inverse binomial sums can be expressed in terms of generalized
log-sine integrals \cite{mcv, davydychev-bin}.
The structure of this article is as follows. In Section \ref{sec:prelim} our
basic tools are described. After providing necessary results on log-sine
integrals in Section \ref{sec:logsin}, we turn to relationships between random
walks and Mahler measures in Section \ref{sec:walks}. In particular, we will be
interested in the multiple Mahler measure $\mu_n(1+x+y)$ which has a fine
hypergeometric form \eqref{eq:hyper} and a natural trigonometric
representation \eqref{eq:mu-trigintegral} as a double integral.
In Section \ref{sec:epsilon} we directly expand \eqref{eq:hyper} and use
known results from the $\varepsilon$-expansion of hypergeometric functions
\cite{dk1-eps, davydychev-bin} to obtain $\mu_n(1+x+y)$ in terms of multiple
inverse binomial sums. In the cases $n=1,2,3$ this leads to explicit
polylogarithmic evaluations.
An alternative approach based of the double integral representation
\eqref{eq:mu-trigintegral} is taken up in Section \ref{sec:mun} which considers
the evaluation of the inner integral in \eqref{eq:mu-trigintegral}. Aided by
combinatorics, we show in Theorems \ref{thm:mun} and \ref{thm:rhon} that these
can always be expressed in terms of multiple harmonic polylogarithms of weight
$k$. Accordingly, we show in Section \ref{sec:lireductions} how these
polylogarithms can be reduced explicitly for low weights. In Section
\ref{sec:mu2} we reprise from \cite{logsin1} the evaluation of $\mu_2(1+x+y)$.
Then in Section \ref{sec:mu3} we apply our general results from Section
\ref{sec:mun} to a conjectural evaluation of $\mu_3(1+x+y)$.
In Section \ref{sec:boyd} we finish with an elementary proof of two recently
established 1998 conjectures of Boyd and use these tools to obtain a new Mahler
measure.
\section{Preliminaries}\label{sec:prelim}
For $k$ functions (typically Laurent polynomials) in $n$ variables the
\emph{multiple Mahler measure}, introduced in \cite{klo}, is defined by
\begin{equation*}
\mu(P_1,P_2, \ldots, P_k)
:= \int_0^1 \cdots \int_0^1 \prod_{j=1}^k \log \left| P_j\left(e^{2\pi i t_1},
\ldots, e^{2\pi i t_n}\right)\right| \mathrm{d} t_1 \mathrm{d} t_2 \ldots \mathrm{d} t_n.
\end{equation*}
When $P=P_1=P_2= \cdots =P_k$ this devolves to a \emph{higher Mahler measure},
$\mu_k(P)$, as introduced and examined in \cite{klo}. When $k=1$ both reduce
to the standard (logarithmic) \emph{Mahler measure} \cite{boyd}.
We also recall \emph{Jensen's formula}:
\begin{equation}\label{eq:sl}
\int_0^1 \log \left|\alpha-e^{2\pi i\,t}\right|\,\mathrm{d} t
= \log \left(\max\{|\alpha|,1\}\right).
\end{equation}
An easy consequence of Jensen's formula is that for complex constants $a$ and
$b$ we have
\begin{equation}\label{eq:mlin}
\mu(ax+b) = \log|a|\vee\log |b|.
\end{equation}
In the following development,
\begin{equation*}
\op{Li}_{a_1,\ldots,a_k}(z)
:= \sum_{n_1>\cdots>n_k>0} \frac{z^{n_1}}{n_1^{a_1} \cdots n_k^{a_k}}
\end{equation*}
denotes the \emph{generalized polylogarithm} as is studied in \cite{mcv} and in
\cite[Ch. 3]{bbg}. For example, $\op{Li}_{2,1}(z) =\sum_{k=1}^\infty
\frac{z^k}{k^2}\sum_{j=1}^{k-1} \frac1j$. In particular, $\op{Li}_k(x) :=
\sum_{n=1}^\infty \frac{x^n}{n^k}$ is the \emph{polylogarithm of order $k$} and
\begin{equation*}
\Ti{k}{x} := \sum_{n=0}^\infty \,(-1)^n\frac{ x^{2n+1}}{(2n+1)^{k}}
\end{equation*}
is the related \emph{inverse tangent of order $k$}. We use the same notation for
the analytic continuations of these functions.
Moreover, \emph{multiple zeta values} are denoted by
\begin{equation*}
\zeta(a_1,\ldots,a_k) := \op{Li}_{a_1,\ldots,a_k}(1).
\end{equation*}
Similarly, we consider the \emph{multiple Clausen functions} ($\op{Cl}$) and
\emph{multiple Glaisher functions} ($\op{Gl}$) of depth $k$ defined as
\begin{align}
\Cl{a_1,\ldots,a_k}{\theta} &= \left\{ \begin{array}{ll}
\Im \op{Li}_{a_1,\ldots,a_k}(e^{i\theta}) & \text{if $w$ even}\\
\Re \op{Li}_{a_1,\ldots,a_k}(e^{i\theta}) & \text{if $w$ odd}
\end{array} \right\}, \label{eq:defcl}\\
\Gl{a_1,\ldots,a_k}{\theta} &= \left\{ \begin{array}{ll}
\Re \op{Li}_{a_1,\ldots,a_k}(e^{i\theta}) & \text{if $w$ even}\\
\Im \op{Li}_{a_1,\ldots,a_k}(e^{i\theta}) & \text{if $w$ odd}
\end{array} \right\}, \label{eq:defgl}
\end{align}
where $w=a_1+\ldots+a_k$ is the weight of the function.
As illustrated in \eqref{eq:defcl2}, the Clausen and Glaisher functions
alternate between being cosine and sine series with the parity of the
dimension. Of particular importance will be the case of $\theta = \pi/3$ which
has also been considered in \cite{mcv}.
Our other notation and usage is largely consistent with that in \cite{lewin2}
and the newly published \cite{NIST}, in which most of the requisite
material is described. Finally, a recent elaboration of what is meant when we
speak about evaluations and ``closed forms'' is to be found in \cite{closed}.
\section{Log-sine integrals} \label{sec:logsin}
For $n=1,2, \ldots$, we consider the \emph{log-sine integrals} defined by
\begin{equation}\label{eq:slx}
\Ls{n}{\sigma} := -\int_{0}^{\sigma}\log^{n-1} \left|2\,\sin \frac\theta 2\right|\,{\mathrm{d}\theta}
\end{equation}
and their moments for $k\ge 0$ given by
\begin{equation}\label{eq:slxm}
\LsD{n}{k}{\sigma}
:= -\int_{0}^{\sigma}\theta^k\,\log^{n-1-k} \left|2\,\sin \frac\theta 2\right| \,{\mathrm{d}\theta}.
\end{equation}
In each case the modulus is not needed for $0 \le \sigma \le 2\pi$. Various
log-sine integral evaluations may be found in \cite[\S7.6 \& \S7.9]{lewin2}.
We observe that $\Ls{1}{\sigma} = -\sigma$ and that $\LsD{n}{0}{\sigma} =
\Ls{n}{\sigma}$. This is the notation used by Lewin \cite{lewin, lewin2}. In
particular,
\begin{equation}\label{eq:defcl2}
\Ls{2}{\sigma} = \Cl{2}{\sigma}
:= \sum_{n=1}^\infty \frac{\sin(n\sigma)}{n^2}
\end{equation}
is the Clausen function introduced in \eqref{eq:defcl}.
\subsection[Log-sine integrals at pi]{Log-sine integrals at $\pi$}\label{ssec:lspi}
We first recall that the log-sine integrals at $\pi$ can always be evaluated in
terms of zeta values. This is a consequence of the exponential generating
function \cite[Eqn. (7.109)]{lewin2}
\begin{equation}\label{eq:lsatpiogf}
-\frac1\pi\,\sum_{m=0}^{\infty} \Ls{m+1}{\pi} \,\frac{u^m}{m!}
=\frac{\Gamma\left(1+u\right)}{\Gamma^2\left(1+\frac u2\right)}
= \binom{u}{u/2}.
\end{equation}
This will be revisited and further explained in Section \ref{sec:walks}.
\begin{example}[Values of $\Ls{n}{\pi}$]\label{ex:pi}
For instance, we have $\Ls{2}{\pi} = 0$ as well as
\begin{align*}
-\Ls{3}{\pi} &= {\frac{1}{12}}\,\pi^3\\
\Ls{4}{\pi} &= \frac32\pi\,\zeta(3)\\
-\Ls{5}{\pi} &= \frac{19}{240}\,\pi^5\\
\Ls{6}{\pi} &= {\frac{45}{2}}\,\pi \,\zeta(5)+\frac54\,\pi^3\zeta(3)\\
- \Ls{7}{\pi} &= {\frac{275}{1344}}\,\pi^7+{\frac{45}{2}}\,\pi \, \zeta^2(3)\\
\Ls{8}{\pi} &= {\frac{2835}{4}}\,\pi \,\zeta(7)+ {\frac{315}{8}}\,\pi^3\zeta(5)
+{\frac{133}{32}}\,\pi^5\zeta(3),
\end{align*}
and so forth. Note that these values may be conveniently obtained from
\eqref{eq:lsatpiogf} by a computer algebra system as the following snippet of
\emph{Maple} code demonstrates:\\
\texttt{for k to 6 do simplify(subs(x=0,diff(Pi*binomial(x,x/2),x\$k))) od}\\
More general log-sine evaluations with an emphasis on automatic evaluations
have been studied in \cite{logsin-evaluations}.
\hspace*{\fill}$\Diamond$\medskip
\end{example}
For general log-sine integrals, in \cite{logsin-evaluations} the following
computationally effective exponential generating function was obtained.
\begin{theorem}[Generating function for $\LsD{n+k+1}{k}{\pi}$]\label{prop:gfb}
For $2|\mu| <\lambda <1$ we have
\begin{align}
\sum_{n,k \ge 0}\LsD{n+k+1}{k}{\pi}\frac{\lambda^n}{n!}\frac{(i\mu)^k}{k!}
&= -i \sum_{n\ge0} \binom{\lambda}{n}
\frac{(-1)^n\, \mathrm{e}^{i\pi\frac\lambda2} - \mathrm{e}^{i\pi\mu}}{\mu-\frac\lambda2+n}. \label{eq:gfsum}
\end{align}
\end{theorem}
One may extract one-variable generating functions from \eqref{eq:gfsum}. For instance,
\begin{equation*}
\sum_{n=0}^\infty \LsD{n+2}{1}{\pi}\frac{\lambda^n}{n!}
= \sum_{n=0}^{\infty} \binom{\lambda}{n} \frac{-1 + (-1)^n \cos\frac{\pi\lambda}2}{\left(n -\frac\lambda2 \right)^2}.
\end{equation*}
The log-sine integrals at $\pi/3$ are especially useful as illustrated in
\cite{mcv} and are discussed at some length in \cite{logsin1} where other
applications to Mahler measures are given.
\subsection{Extensions of the log-sine integrals}\label{ssec:ls3}
It is possible to extend some of these considerations to the log-sine-cosine integrals
\begin{equation}\label{eq:deflsc}
\Lsc{m}{n}{\sigma} := -\int_{0}^{\sigma}\log^{m-1} \left|2\,\sin \frac \theta 2\right|\,\log^{n-1} \left|2\,\cos \frac \theta 2\right| \,\mathrm{d}\theta.
\end{equation}
Then $\Lsc{m}{1}{\sigma}= \Ls{m}{\sigma}$.
Let us set
\begin{equation}
\mu_{m,n}(1-x,1+x) :=\mu(\underbrace{1-x, \cdots,1-x}_m,\underbrace{1+x, \cdots,1+x}_n).
\end{equation}
Then immediately from the definition we obtain the following:
\begin{theorem}[Evaluation of $\mu_{m,n}(1-x,1+x)$]\label{thm:gls}
For non-negative integers $m,n$,
\begin{equation}\label{eq:gslmahler}
\mu_{m,n}(1-x,1+x) = -\frac1\pi\,\Lsc{m+1}{n+1}{\pi}.
\end{equation}
\end{theorem}
In every case this is evaluable in terms of zeta values. Indeed, using the result in \cite[\S7.9.2, (7.114)]{lewin2}, we obtain the
generating function
\begin{equation}\label{eq:glsatpiogf}
{\rm gs}(u,v):=-\frac1\pi\, \sum_{m,n=0}^\infty \Lsc{m+1}{n+1}{\pi} \,\frac{u^m}{m!}\frac{v^n}{n!}
=\frac{2^{u+v}}\pi\,\frac{\Gamma \left( \frac{1+u}2 \right)\Gamma \left(\frac{1+v}2 \right)}{\Gamma \left( 1+\frac{u+v}2 \right)}.
\end{equation}
From the duplication formula for the Gamma function this can be rewritten as
\[{\rm gs}(u,v) = \binom{u}{\frac u2} \binom{v}{\frac v2} {\frac{\Gamma \left(
1+\frac u2 \right) \Gamma \left(1+\frac v2 \right) }{\Gamma \left( 1+\frac{u+v}2 \right) }},
\]
so that
\[{\rm gs}(u,0) = \binom{u}{\frac u2} = {\rm gs}(u,u).\]
From here it is apparent that \eqref{eq:glsatpiogf} is an extension of
\eqref{eq:lsatpiogf}:
\begin{example}[Values of $\Lsc{n}{m}{\pi}$]\label{ex:gpi}
For instance,
\begin{align*}\label{ex:gls}
\mu_{2,1}(1-x,1+x) &= \mu_{1,2}(1-x,1+x) = \frac{1}{4}\zeta(3), \\
\mu_{3,2}(1-x,1+x) &= \frac{3}{4}\zeta(5)-\frac{1}{8}\pi^2\zeta(3), \\
\mu_{6,3}(1-x,1+x) &=
\frac{315}{4}\zeta(9) + \frac{135}{32}\pi^2\zeta(7) + \frac{309}{128}\pi^4\zeta(5)
- \frac{45}{256}\pi^6\zeta(3) - \frac{1575}{32}\zeta^3(3).
\end{align*}
As in Example \ref{ex:pi} this can be easily obtained with a line of code
in a computer algebra system such as \emph{Mathematica} or \emph{Maple}.
\hspace*{\fill}$\Diamond$\medskip
\end{example}
\begin{remark}
Also ${\rm gs}(u,-u)=\sec(\pi/(2u)).$ From this we deduce for $n=0,1,2,\ldots$ that
\[\sum_{k=0}^n (-1)^k \mu_{k,n-k}(1-x,1+x) = |E_{2n}|\,\frac{\left(\frac\pi2\right)^{2n}}{(2n)!} =\frac4\pi \,L_{-4}(2n+1),\]
where $E_{2n}$ are the even Euler numbers: $1,-1,5,-61,1385 \ldots$\,.
\hspace*{\fill}$\Diamond$\medskip
\end{remark}
A more recondite \emph{extended log-sine integral of order three} is developed
in \cite[\S 8.4.3]{lewin2} from properties of the trilogarithm. It is defined by
\begin{align}
\Ls3{\theta,\omega}
&:=-\int_0^\theta\,\log \left|2\sin \frac\sigma 2\right|\, \log\left|2\sin \frac{ \sigma+\omega}2 \right| \,\mathrm{d} \sigma,
\end{align}
so that $\Ls3{\theta,0} = \Ls3{\theta}.$
This extended log-sine integral reduces as follows:
\begin{align}\label{eq:ls3}
-\Ls3{2\theta, 2\omega}
&= \frac12 \Ls3{2\omega} -\frac12 \Ls3{2\theta}
-\frac12 \Ls3{2\theta+2\omega} \nonumber\\
&\quad- 2\Im\op{Li}_3\left(\frac{\sin(\theta) \mathrm{e}^{i\omega}}{\sin(\theta+\omega)}\right)
+\theta\log^2 \left( \frac{\sin(\theta)}{\sin(\theta+\omega)}\right) \nonumber \\
&\quad+ \log\left( {\frac{\sin(\theta)}{\sin(\theta+\omega)}} \right)
\left\{ \Cl{2}{2\theta} + \Cl{2}{2\omega} - \Cl{2}{2\theta+2\omega} \right\}.
\end{align}
We note that $-\frac{1}{2\pi} \Ls3{2\pi,\omega} = \mu(1-x,1-e^{i\omega}x)$ but
this is more easily evaluated by Fourier techniques. Indeed one has:
\begin{proposition}[A dilogarithmic measure, part I \cite{klo}]\label{prop:diloga}
For two complex numbers $u$ and $v$ we have
\begin{equation}\mu(1-u\,x,1-v\,x) =\left\{
\begin{array}{ll}
\frac 12\,\Re\op{Li}_2\left(v\overline{u} \right), & \text{if $|u| \le 1,|v| \le 1$,} \\
\frac 12\,\Re\op{Li}_2\left(\frac v{\overline{u}}\right), & \text{if $|u| \ge 1,|v| \le 1$,} \\
\frac 12\,\Re\op{Li}_2\left(\frac 1{v\overline{u}} \right)+ \log |u| \log |v| , & \text{if $|u|\ge 1 ,|v| \ge 1$.}
\end{array}
\right.
\end{equation}
\end{proposition}
This is proven much as is \eqref{eq:mu2} of Proposition \ref{prop:mu2}. In
Lewin's terms \cite[A.2.5]{lewin2} for $0 < \theta\le 2\pi$ and $r\ge0$ we may
write
\begin{equation}\label{def:Li}
\Re\op{Li}_2\left(re^{i\theta}\right) =: \op{Li}_2\left(r,\theta\right)= -\frac 12\int_0^r \log\left(t^2+1-2t \cos \theta\right)\frac{\mathrm{d} t}{t},
\end{equation}
with the reflection formula
\begin{equation}
\op{Li}_2\left(r,\theta\right) + \op{Li}_2\left(\frac1r,\theta\right)
= \zeta(2)- \frac12\,\log^2 r+\frac12\,(\pi-\theta)^2.
\end{equation}
This leads to:
\begin{proposition}[A dilogarithmic measure, part II]\label{prop:dilogb}
For complex numbers $u=re^{i\theta}$ and $v=se^{i\tau}$ we have
\begin{equation}\mu(1-u \, x,1-v \, x) =\left\{
\begin{array}{ll}
\frac 12\,\op{Li}_2\left(rs,\theta-\tau\right) & \text{if $r \le 1,s \le 1$,} \\
\frac 12\,\op{Li}_2\left(\frac s r,\theta+\tau\right), & \text{if $r\ge 1,s \le 1$,} \\
\frac 12\,\op{Li}_2\left(\frac1{sr},\theta-\tau\right)+ \log r \log s , & \text{if $r\ge 1 ,s \ge 1$.}
\end{array}
\right.
\end{equation}
\end{proposition}
\section{Mahler measures and moments of random walks}\label{sec:walks}
The $s$-th moments of an $n$-step uniform random walk are given by
\[ W_n (s) = \int_0^1 \ldots \int_0^1 \left| \sum_{k = 1}^n e^{2 \pi
i t_k} \right|^s \mathrm{d} t_1 \cdots \mathrm{d} t_n \]
and their relation with Mahler measure is observed in \cite{bswz-densities}.
In particular,
\[ W_n' (0) = \mu (1 + x_1 + \ldots + x_{n-1}), \]
with the cases $2 \le n \le 6$
discussed in \cite{logsin1}.
Higher derivatives of $W_n$ correspond to higher Mahler measures:
\begin{equation}\label{eq:diffW}
W_n^{(m)}(0) = \mu_m (1 + x_1 + \ldots + x_{n-1}).
\end{equation}
The evaluation $W_2(s) = \binom{s}{s/2}$ thus
explains and proves the generating function \eqref{eq:lsatpiogf}; in other
words, we find that
\begin{align}\label{w3dm}
W_2^{(m)}(0) &= -\frac1\pi \Ls{m+1}{\pi}.
\end{align}
As a consequence of the study of random walks in \cite{bswz-densities} we
record the following generating function for $\mu_k(1+x+y)$ which follows from
\eqref{eq:diffW} and the hypergeometric expression for $W_3$ in \cite[Thm.
10]{bswz-densities}. There is a corresponding expression, using a single
Meijer-$G$ function, for $W_4$ (i.e., $\mu_m(1+x+y+z)$) given in \cite[Thm.
11]{bswz-densities}.
\begin{theorem}[Hypergeometric form for $W_3(s)$]\label{thm:W3}
For complex $|s|<2$, we may write
\begin{align}\label{eq:hyper}
W_3(s) = \sum_{n=0}^\infty \mu_n(1+x+y)\frac{s^n}{n!}
&=\frac {\sqrt{3}}{2\pi}\, 3^{s+1}
\,\frac{\Gamma(1+\frac s2)^2}{\Gamma(s+2)}\,
\pFq32{\frac{s+2}2, \frac{s+2}2,\frac{s+2}2}{1,\frac{s+3}2}{\frac14} \\
&=\frac{\sqrt {3}}{\pi }\left(\frac 32\right)^{s+1}\int _{0}^{1}\!{\frac {{z}^{1+s}
{\pFq21{1+\frac s2,1+\frac s2}{1}{\frac{z^2}4}}}{\sqrt
{1-{z}^{2}}}}\,{\,\mathrm{d} z}. \label{eq:hyperb}
\end{align}
\end{theorem}
\begin{proof}
Equation \eqref{eq:hyper} is proven in \cite[Thm. 10]{bswz-densities}, while
\eqref{eq:hyperb} is a consequence of \eqref{eq:hyper} and \cite[Eqn.
(16.5.2)]{NIST}.
\end{proof}
We shall exploit Theorem \ref{thm:W3} next, in Section \ref{sec:epsilon}. For
integers $n \ge 1$ we also have
\begin{equation}\label{eq:mu-trigintegral}
\mu_n(1+x+y) =\frac{1}{4\pi^2} \int_0^{2\pi} \mathrm{d}\theta
\int_0^{2\pi} \left(\Re\log\left(1-2\sin(\theta)\mathrm{e}^{i\,\omega}\right)\right)^n\,\mathrm{d} \omega,
\end{equation}
as follows directly from the definition and some simple trigonometry, since
$\Re\log =\log|\cdot|$. This is the basis for the evaluations of Section
\ref{sec:mun}. In particular, in Section \ref{sec:mun} we will evaluate the
inner integral in terms of multiple harmonic polylogarithms.
\section{Epsilon expansion of $W_3$}\label{sec:epsilon}
In this section we use known results from the $\varepsilon$-expansion of
hypergeometric functions \cite{dk1-eps, davydychev-bin} to obtain
$\mu_n(1+x+y)$ in terms of multiple inverse binomial sums. We then derive
complete evaluations of $\mu_1(1+x+y)$, $\mu_2(1+x+y)$ and $\mu_3(1+x+y)$. An
alternative approach will be pursued in Sections \ref{sec:mun} and
\ref{sec:mu23}.
In light of Theorem \ref{thm:W3}, the evaluation of $\mu_n(1+x+y)$
is essentially reduced to the Taylor expansion
\begin{align}\label{eq:alphadef}
\pFq32{\frac{\varepsilon+2}2, \frac{\varepsilon+2}2,\frac{\varepsilon+2}2}{1,\frac{\varepsilon+3}2}{\frac14}
= \sum_{n=0}^\infty \alpha_n \varepsilon^n.
\end{align}
Indeed, from \eqref{eq:hyper} and Leibniz' rule we have
\begin{align}\label{eq:mun-sum}
\mu_n(1+x+y) &= \frac {\sqrt{3}}{2\pi} \sum _{k=0}^n \binom{n}{k} \alpha_k \beta_{n-k}
\end{align}
where $\beta_k$ is defined by
\begin{align}\label{eq:beta-def}
3^{ \varepsilon+1} \,\frac{\Gamma(1+\frac \varepsilon2)^2}{\Gamma(2+\varepsilon)}
= \sum_{n=0}^\infty \beta_n \varepsilon^n.
\end{align}
Note that $\beta_k$ is easy to compute as illustrated in Example \ref{ex:pi}.
The expansion of hypergeometric functions in terms of their parameters as in
\eqref{eq:alphadef} occurs in physics \cite{dk1-eps, davydychev-bin} in the
context of the evaluation of Feynman diagrams and is commonly referred to
as \emph{epsilon expansion}, thus explaining the choice of variable in
\eqref{eq:alphadef}.
\begin{remark}
From \eqref{eq:beta-def} we see that $\beta_k$ may be computed directly from
the coefficients $\gamma_k$ defined by the Taylor expansion
\begin{equation*}
\frac{\Gamma(1+\frac\varepsilon2)^2}{\Gamma(1+\varepsilon)}
= \frac{1}{\binom{\varepsilon}{\varepsilon/2}} = \sum_{n=0}^\infty \gamma_n \varepsilon^n.
\end{equation*}
Appealing to \eqref{eq:lsatpiogf} we find that $\gamma_n$ is recursively determined
by $\gamma_0=1$ and
\begin{equation*}
\gamma_n = \frac{1}{\pi} \sum_{k=1}^n \Ls{k+1}{\pi} \frac{\gamma_{n-k}}{k!}.
\end{equation*}
In particular, the results of Section \ref{ssec:lspi} show that $\gamma_n$ can
always be expressed in terms of zeta values. Accordingly, $\beta_n$ evaluates
in terms of $\log3$ and zeta values.
\hspace*{\fill}$\Diamond$\medskip
\end{remark}
Let $S_k(j) := \sum_{m=1}^j \frac{1}{m^k}$ denote the harmonic numbers of order
$k$. Following \cite{davydychev-bin} we abbreviate $S_k:=S_k(j-1)$ and
$\bar{S}_k:=S_k(2j-1)$ in order to make it more clear which results in this
reference contribute to the evaluations below. As in \cite[Appendix
B]{dk1-eps}, we use the duplication formula $(2a)_{2j} = 4^j (a)_j (a+1/2)_j$
as well as the expansion
\begin{align}
\frac{(m+a\varepsilon)_j}{(m)_j} = \exp\left[ -\sum_{k=1}^\infty \frac{(-a\varepsilon)^k}{k} \left[S_k(j+m-1) - S_k(m-1)\right] \right]
\end{align}
where $m$ is a positive integer
to write:
\begin{align}
\nonumber \pFq32{\frac{\varepsilon+2}2,
\frac{\varepsilon+2}2,\frac{\varepsilon+2}2}{1,\frac{\varepsilon+3}2}{\frac14}
&= \sum_{j=0}^\infty \frac{(1+\varepsilon/2)_j^3}{4^j (j!)^2 (3/2+\varepsilon/2)_j}
\\ \nonumber
&= \sum_{j=0}^\infty \frac{(1+\varepsilon/2)_j^4}{(j!)^2 (2+\varepsilon)_{2j}}
\\ \nonumber
&= \sum_{j=0}^\infty \frac{2}{j+1} \frac{1}{\binom{2(j+1)}{j+1}}
\left[\frac{(1+\varepsilon/2)_j}{j!}\right]^4 \left[\frac{(2+\varepsilon)_{2j}}{(2j+1)!}\right]^{-1} \\
&= \sum_{j=1}^\infty \frac{2}{j} \frac{1}{\binom{2j}{j}}
\exp\left[ \sum_{k=1}^\infty \frac{(-\varepsilon)^k}{k} A_{k,j} \right] \label{eq:hypeps}
\end{align}
where
\begin{align}\label{eq:akj}
A_{k,j} :=S_k(2j-1)-1 -
4\frac{S_k(j-1)}{2^k} =
\sum_{m=2}^{2j-1}\frac{2(-1)^{m+1}-1}{m^k}.
\end{align}
We can now read off the terms $\alpha_n$ of the $\varepsilon$-expansion:
\begin{theorem}\label{thm:an}
For $n =0,1,2,\ldots$
\begin{align}\label{eq:alphan}
\alpha_n = [\varepsilon^n] \;
\pFq32{\frac{\varepsilon+2}2, \frac{\varepsilon+2}2,\frac{\varepsilon+2}2}{1,\frac{\varepsilon+3}2}{\frac14}
&= (-1)^n \sum_{j=1}^\infty \frac{2}{j} \frac{1}{\binom{2j}{j}} \sum
\prod_{k=1}^n \frac{A_{k,j}^{m_k}}{m_k! k^{m_k}}
\end{align}
where the inner sum is over all non-negative integers $m_1,\ldots,m_n$ such
that $m_1+2m_2+\ldots+n m_n = n$.
\end{theorem}
\begin{proof}
Equation \eqref{eq:alphan} may be derived from \eqref{eq:hypeps} using, for
instance, Fa\`{a} di Bruno's formula for the $n$-th derivative of the composition
on two functions.
\end{proof}
\begin{example}[$\alpha_0$, $\alpha_1$ and $\alpha_2$]
In particular,
\begin{align*}
\alpha_1 = [\varepsilon] \;
\pFq32{\frac{\varepsilon+2}2, \frac{\varepsilon+2}2,\frac{\varepsilon+2}2}{1,\frac{\varepsilon+3}2}{\frac14}
&= -\sum_{j=1}^\infty \frac{2}{j} \frac{1}{\binom{2j}{j}} A_{1,j} \\
&= -\sum_{j=1}^\infty \frac{2}{j} \frac{1}{\binom{2j}{j}} \left[ \bar{S_1}-2S_1-1
\right].
\end{align*}
Such \emph{multiple inverse binomial sums} are studied in
\cite{davydychev-bin}. In particular, using \cite[(2.20),
(2.21)]{davydychev-bin} we find
\begin{align}
\alpha_0 &= \frac{2\pi}{3\sqrt{3}}, \label{eq:eps0}\\
\alpha_1 &= \frac{2}{3\sqrt{3}} \left[ \pi - \pi\log3 + \Ls{2}{\frac\pi3} \right].\label{eq:eps1}
\end{align}
For the second term $\alpha_2$ in the $\varepsilon$-expansion
\eqref{eq:hypeps} produces
\begin{align*}
[\varepsilon^2] \;
\pFq32{\frac{\varepsilon+2}2, \frac{\varepsilon+2}2,\frac{\varepsilon+2}2}{1,\frac{\varepsilon+3}2}{\frac14}
&= \sum_{j=1}^\infty \frac{1}{j} \frac{1}{\binom{2j}{j}} \left[ A_{1,j}^2 + A_{2,j} \right] \\
&= \sum_{j=1}^\infty \frac{1}{j} \frac{1}{\binom{2j}{j}} \left[ \bar{S}_2-S_2 + (\bar{S}_1-2S_1)^2 -2\bar{S}_1+4S_1 \right].
\end{align*}
Using \cite[(2.8),(2.22)-(2.24)]{davydychev-bin} we obtain
\begin{align}\label{eq:eps2}
\alpha_2
&= \frac{2}{3\sqrt{3}} \left[ \frac{\pi}{72} - \pi\log3 + \frac12 \pi\log3
+ (1-\log3)\Ls{2}{\frac\pi3} \right. \nonumber\\
&\quad\left.+ \frac32\Ls{3}{\frac\pi3}
+ \frac32\Ls{3}{\frac{2\pi}{3}} - 3\Ls{3}{\pi} \right].
\end{align}
\hspace*{\fill}$\Diamond$\medskip
\end{example}
These results provide us with evaluations of $\mu_1(1+x+y)$ and
$\mu_2(1+x+y)$ as given next. As expected, the result for $\mu_1(1+x+y)$ agrees
with Smyth's original evaluation, and the result for
$\mu_2(1+x+y)$ agrees with our prior evaluation in \cite{logsin1}.
The latter evaluation will be recalled in Section \ref{sec:mu2}.
\begin{theorem}[Evaluation of $\mu_1(1+x+y)$ and $\mu_2(1+x+y)$]\label{thm:mu21eps}
We have
\begin{align}
\mu_1(1+x+y) &= \frac{1}{\pi} \Ls{2}{\frac\pi3}, \label{eq:mu1eps}\\
\mu_2(1+x+y) &= \frac3\pi \Ls{3}{\frac{2\pi}{3}} + \frac{\pi^2}{4}. \label{eq:mu2eps}
\end{align}
\end{theorem}
\begin{proof}
Using Theorem \ref{thm:W3} we obtain
\begin{align}
\mu_1(1+x+y) = \frac{3\sqrt{3}}{2\pi} \left[ (\log3-1)\alpha_0 + \alpha_1 \right].
\end{align}
Combining this with equations \eqref{eq:eps0}, \eqref{eq:eps1} yields (\ref{eq:mu1eps}).
Again, using Theorem \ref{thm:W3} we obtain
\begin{align}
\mu_2(1+x+y) = \frac{3\sqrt{3}}{2\pi} \left[ (\log^2 3 - 2\log3 + 2 - \tfrac{\pi^2}{12})\alpha_0
+ 2(\log3 - 1) \alpha_1 + 2 \alpha_2 \right].
\end{align}
Combining this with equations \eqref{eq:eps0}, \eqref{eq:eps1}, and \eqref{eq:eps2} yields
\begin{align}
\pi\mu_2(1+x+y)
&= 3\Ls{3}{\frac{2\pi}{3}} + 3\Ls{3}{\frac\pi3} - 6\Ls{3}{\pi} - \frac{\pi^3}{18} \nonumber\\
&= 3\Ls{3}{\frac{2\pi}{3}} + \frac{\pi^3}{4}.
\end{align}
The last equality follows, for instance, automatically from the results in
\cite{logsin-evaluations}.
\end{proof}
\begin{example}[$\alpha_3$]
The evaluation of $\alpha_3$ is more involved and we omit some
details. Again, \eqref{eq:hypeps} produces
\begin{align*}
[\varepsilon^3] \;
\pFq32{\frac{\varepsilon+2}2, \frac{\varepsilon+2}2,\frac{\varepsilon+2}2}{1,\frac{\varepsilon+3}2}{\frac14}
&= -\frac13 \sum_{j=1}^\infty \frac{1}{j} \frac{1}{\binom{2j}{j}}
\left[ A_{1,j}^3 + 3A_{1,j}A_{2,j} + 2A_{3,j} \right].
\end{align*}
Using \cite[(2.25)-(2.28),(2.68)-(2.70),(2.81),(2.89)]{davydychev-bin} as well as results from \cite{logsin-evaluations} we obtain
\begin{align}\label{eq:eps3}
\alpha_3
&= \frac{2}{3\sqrt{3}} \left[ \frac{5\pi^3}{108}(1-\log3) + \frac12 \pi\log^2 3
- \frac16 \pi \log^3 3 + \frac{11}{9} \pi \zeta(3) \right. \nonumber\\
&\quad + \Cl{2}{\frac\pi3} \left( \frac{5}{36} \pi^2 - \log3 + \frac12 \log^2 3 \right)
- 3 \Gl{2,1}{\frac{2\pi}{3}} \left( 1 - \log3 \right) \nonumber\\
&\quad\left.- \frac{35}{6} \Cl{4}{\frac\pi3} + 15 \Cl{2,1,1}{\frac{2\pi}{3}}
- 3 \Lsc{2}{3}{\frac\pi3} \right].
\end{align}
Observe the occurrence of the log-sine-cosine integral
$\Lsc{2}{3}{\frac\pi3}$. The log-sine-cosine integrals were defined
in \eqref{eq:deflsc}.
\hspace*{\fill}$\Diamond$\medskip
\end{example}
Proceeding as in the proof of Theorem \ref{thm:mu21eps} we obtain:
\begin{theorem}[Evaluation of $\mu_3(1+x+y)$]\label{thm:mu3eps}
We have
\begin{align}\label{eq:mu3eps}
\pi\mu_3(1+x+y)
&= 15\Ls{4}{\frac{2\pi}{3}} - 18\Lsc{2}{3}{\frac\pi3} - 15\Cl{4}{\frac\pi3} \nonumber\\
&\quad- \frac14\pi^2\Cl{2}{\frac\pi3} - 17\pi\zeta(3).
\end{align}
\end{theorem}
The log-sine-cosine integral $\Lsc{2}{3}{\frac\pi3}$ appears to reduce further as
\begin{align}
12\Lsc{2}{3}{\frac\pi3}
&\stackrel{?[1]}{=} 6\Ls{4}{\frac{2\pi}{3}} - 4\Cl{4}{\frac\pi3} - 7\pi\zeta(3) \\
&= 6\Ls{4}{\frac{2\pi}{3}} - \frac89\Ls{4}{\frac\pi3} - \frac{59}{9}\pi\zeta(3). \label{eq:lsc23}
\end{align}
This conjectural reduction also appears in \cite[(A.30)]{dk1-eps} where it was
found via PSLQ. Combining this with \eqref{eq:mu3eps}, we obtain an conjectural
evaluation of $\mu_3(1+x+y)$ equivalent to \eqref{eq:mu3ls}.
On the other hand, it follows from \cite[(2.18)]{davydychev-bin} that
\begin{align}
12\Lsc{2}{3}{\frac\pi3}
&= \Ls{4}{\frac{2\pi}{3}} - 4\Ls{4}{\frac\pi3} - \frac{1}{12}\pi\log^3 3 \nonumber\\
&\quad+ 24\Ti{4}{\frac{1}{\sqrt{3}}} + 12\log3 \Ti{3}{\frac{1}{\sqrt{3}}} + 3\log^2 3 \Ti{2}{\frac{1}{\sqrt{3}}}.
\end{align}
Using the known evaluations --- see for instance \cite[(76),(77)]{logsin1} --- for
the inverse tangent integrals of order two and three, we find that
\eqref{eq:lsc23} is equivalent to
\begin{align}\label{eq:lsc23ti}
\Ti{4}{\frac{1}{\sqrt{3}}}
&\stackrel{?[1]}{=} \frac{5}{24}\Ls{4}{\frac{2\pi}{3}} + \frac{7}{54}\Ls{4}{\frac\pi3}
- \frac{59}{216}\pi\zeta(3) - \frac{1}{288}\pi\log^3 3 \nonumber\\
&\quad- \frac12\log3 \Ti{3}{\frac{1}{\sqrt{3}}} - \frac18\log^2 3 \Ti{2}{\frac{1}{\sqrt{3}}}.
\end{align}
\section[Trigonometric analysis of mun(1+x+y)]{Trigonometric analysis of $\mu_n(1+x+y)$}\label{sec:mun}
As promised in \cite{logsin1} --- motivated by the development outlined above
--- we take the analysis of $\mu_n(1+x+y)$ for $n \ge 3$ a fair distance. In
light of \eqref{eq:mu-trigintegral} we define
\begin{align}\label{eq:rho-n}
\rho_n(\alpha) &:= \frac1{2\pi} \int_{-\pi}^{\pi}
\left(\Re\log\left(1-\alpha\,\mathrm{e}^{i\,\omega}\right)\right)^n \,\mathrm{d}\omega
\end{align}
for $n \ge 0$ so that
\begin{align}\label{eq:mu-rhointegral}
\mu_n(1+x+y) &= \frac{1}{2\pi} \int_{-\pi}^{\pi}\,\rho_n(|2 \sin \theta|) \,\mathrm{d}\theta.
\end{align}
We thus typically set $\alpha = |2\sin\theta|$. Note that $\rho_0(\alpha)=1$,
$\rho_1(\alpha)=\log(|\alpha| \vee 1))$.
\begin{proposition}[Properties of $\rho_n$]\label{prop:rho}
Let $n =1,2, \ldots$.
\begin{enumerate}[(a)]
\item\label{prop:ra} For $|\alpha| \le 1$ we have
\begin{align} \label{eq:zn}
\rho_n(\alpha) &=(-1)^n \sum_{m=1}^\infty\,\frac{\alpha^m}{m^n} \,\omega_n(m),
\end{align}
where $\omega_n$ is defined as
\begin{align}\label{eq:cn}
\omega_n(m) &=\sum_{\sum_{j=1}^n k_j=m}\,\frac{1}{2\pi}\,
\int_{-\pi}^{\pi}\, \prod_{j=1}^n\,\frac{m}{k_j}\cos (k_j\omega) \,\mathrm{d} \omega.
\end{align}
\item\label{prop:rb} For $|\alpha| \ge 1$ we have
\begin{align} \label{eq:mun}
\rho_n(\alpha) &= \sum_{k=0}^n \binom{n}{k} \log^{n-k}|\alpha|\,\rho_k\left(\frac 1 \alpha \right).
\end{align}
\end{enumerate}
\end{proposition}
\begin{proof}
For part \eqref{prop:ra} we use \eqref{eq:rho-n} to write
\begin{align*}
\rho_n(\alpha)
&= \frac1{2\pi} \int_{-\pi}^{\pi}\, \left(\Re\,\log\left(1-\alpha\mathrm{e}^{i\,\omega}\right)\right)^n\,\mathrm{d} \omega\\
&=\frac1{2\pi} \int_{-\pi}^{\pi}\, \left\{-\sum_{ k \ge 1}\frac{\alpha^k}{k}\,\cos (k\omega) \right\}^n\,\mathrm{d} \omega\\
&=(-1)^n\sum_{m=1}^\infty\,\frac{\alpha^m}{m^n} \,\omega_n(m),
\end{align*}
as asserted. We note that $|\omega_n(m)| \le m^n$ and so the sum is convergent.
For part \eqref{prop:rb} we now use \eqref{eq:rho-n} to write
\begin{align*}
\rho_n(\alpha)
&= \frac{1}{2\pi} \int_{-\pi}^{\pi}\,\log^n \left(|\alpha|\,\left|1-\alpha^{-1}\mathrm{e}^{i\,\omega}\right| \right)\,\mathrm{d}\omega\\
&= \frac{1}{2\pi} \int_{-\pi}^{\pi}\,\left( \log |\alpha|+\log\left|1-\alpha^{-1}\mathrm{e}^{i\,\omega}\right|\right)^n \,\mathrm{d}\omega\\
&= \sum_{k=0}^n \binom{n}{k} \log^{n-k}|\alpha|\,\frac{1}{2\pi} \int_{-\pi}^{\pi}\,\log^k\left|1-\alpha^{-1}\mathrm{e}^{i\,\omega}\right| \,\mathrm{d}\omega\\
&= \sum_{k=0}^n \binom{n}{k} \log^{n-k}|\alpha|\,\rho_k\left(\frac 1 \alpha \right),
\end{align*}
as required.
\end{proof}
\begin{example}[Evaluation of $\omega_n$ and $\rho_n$ for $n \le 2$]\label{ex:rho2}
We have $\omega_0(m)= 0$, $\omega_1(m) = \delta_0(m)$, and
\begin{align}\label{eqn:omega-ex}
\omega_2(0) =1,\quad
\omega_2(2m) = 2,\quad
\omega_2(2m+1) = 0.
\end{align}
Likewise, $\rho_0(\alpha)= 1$, $\rho_1(\alpha) = \log\left(|\alpha| \vee 1\right)$, and
\begin{align}
\rho_2(\alpha)&= \frac12\op{Li}_2(\alpha^2), \qquad \mbox {for~} |\alpha| \le 1, \label{eqn:rho2}\\
&= \frac12\op{Li}_2\left(\frac{1}{\alpha^2}\right)+\log^2 |\alpha|,\qquad \mbox {for~} |\alpha| \ge 1,\label{eqn:rho-ex}
\end{align}
where the case $n=2$ follows from \eqref{eqn:omega-ex} and Proposition
\ref{prop:rho}.
\hspace*{\fill}$\Diamond$\medskip
\end{example}
We have arrived at an effective description of $\mu_n(1+x+y)$:
\begin{theorem}[Evaluation of $\mu_n(1+x+y)$]\label{thm:mun}
Let $n =1,2, \ldots$. Then
\begin{align}\label{eq:mugen}
\mu_n(1+x+y) &= \frac{1}{\pi}\left\{ \Ls{n+1}{\frac{\pi}{3}}-\Ls{n+1}{\pi} \right\}
+ \frac{2}{\pi} \int_{0}^{\pi/6} \rho_n\left(2\sin\theta\right) \,\mathrm{d}\theta \nonumber\\
&\quad+ \frac{2}{\pi} \sum_{k=2}^{n} \binom{n}{k} \int_{\pi/6}^{\pi/2}\,
\log^{n-k}\,(2\sin\theta)\,\rho_k\left(\frac 1 {2\sin \theta} \right)\mathrm{d} \theta.
\end{align}
\end{theorem}
\begin{proof}
Since $|\alpha| <1$ exactly when $|\theta| < \pi/6$ we start with
\eqref{eq:mu-rhointegral} to get
\begin{align*}
\mu_n(1+x+y) &= \frac{1}{2\pi} \int_{-\pi}^{\pi}\,\rho_n(|2 \sin \theta|) \,\mathrm{d}\theta \\
&= \frac{2}{\pi} \int_{0}^{\pi/6}\,\rho_n(2 \sin \theta) \,\mathrm{d}\theta
+ \frac{2}{\pi} \int_{\pi/6}^{\pi/2}\,\rho_n(2 \sin \theta) \,\mathrm{d}\theta \\
&= \frac{2}{\pi} \int_{0}^{\pi/6}\,\rho_n(2 \sin \theta) \,\mathrm{d}\theta \\
&\quad+ \sum_{k=0}^n \binom{n}{k} \frac{2}{\pi} \int_{\pi/6}^{\pi/2}
\log^{n-k}(2\sin \theta)\,\rho_k\left(\frac 1 {2\sin \theta} \right) \,\mathrm{d}\theta.
\end{align*}
We observe that for $k=1$ the contribution is zero since $\rho_1$ is zero
for $|\alpha|<1$. After evaluating the term with $k=0$ we arrive at
\eqref{eq:mugen}.
\end{proof}
\begin{remark}\label{rem:sasaki}
As is shown in \cite{logsin1},
\begin{equation*}
\frac{1}{\pi}\left\{ \Ls{n+1}{\frac{\pi}{3}} - \Ls{n+1}{\pi} \right\} = \mu_n(1+x+y_*)
\end{equation*}
is a multiple Mahler measure and is recursively evaluable for all $n$.
\hspace*{\fill}$\Diamond$\medskip
\end{remark}
We note the occurrence of values at $\pi/3$ and so record the following:
\begin{example}[Values of $\Ls{n}{\pi/3}$]\label{ex:lspi3}
The following evaluations may be obtained with the help of the implementation accompanying
\cite{logsin-evaluations}.\footnote{These are available for download from \url{http://arminstraub.com/pub/log-sine-integrals}}
\begin{align*}
\Ls{2}{\frac\pi3} &= \mcl{2} \\
-\Ls{3}{\frac\pi3} &= \frac{7}{108}\, \pi^3 \\
\Ls{4}{\frac\pi3} &= \frac12\pi\,\zeta(3)+\frac 92\,\mcl{4} \\
-\Ls{5}{\frac\pi3} &= \frac{1543}{19440}\pi^5 - 6\mgl{4,1} \\
\Ls{6}{\frac\pi3} &= \frac{15}2\pi \,\zeta(5) + \frac{35}{36}\,\pi^3
\zeta(3) + \frac{135}{2}\,\mcl{6} \\
-\Ls{7}{\frac\pi3} &= \frac{74369}{326592}\pi^7 + \frac{15}{2}\,\pi
\zeta(3)^2 - 135\,\mgl{6,1} \\
\Ls{8}{\frac\pi3} &= \frac{13181}{2592}\pi^5\zeta(3) + \frac{1225}{24}\pi^3\zeta(5)
+ \frac{319445}{864}\pi\zeta(7) \\
&\quad+ \frac{35}{2}\pi^2\mcl{6}
+ \frac{945}{4}\mcl{8} + 315\mcl{6,1,1}
\end{align*}\hspace*{\fill}$\Diamond$\medskip
\end{example}
\subsection[Further evaluation of rho]{Further evaluation of $\rho_n$}\label{ssec:rhon}
To make further progress, we need first to determine $\rho_n$ for $n \ge 3$.
It is instructive to explore the next few cases.
\begin{example}[Evaluation of $\omega_3$ and $\rho_3$]\label{ex:rho3}
We use{\small \[4\,\cos \left( a \right) \cos \left( b \right) \cos \left( c \right) =
\cos \left( a+b+c \right) +\cos \left( a-b-c \right) +\cos \left( a-b+
c \right) +\cos \left( a-c+b \right)\]}
and so derive
\begin{align*}
\omega_3(m) &= \frac 14\,\sum\, \left \{\frac{m^3}{ijk} \colon i\pm j\pm k =0, i+j+k=m\right\}.
\end{align*}
Note that we must have exactly one of $i=j+k, j=k+i$ or $k=i+j$.
We thus learn that $\omega_3(2m+1)=0$. Moreover, by symmetry,
\begin{align}\label{eq:omega3}
\omega_3(2m) &= \frac34\,\sum_{j+k=m} \frac{(2m)^3}{jk(j+k)} \nonumber\\
&= 6\,\sum_{j+k=m} \frac{m^2}{jk} = 12\,m\sum_{k=1}^{m-1} \frac 1k .
\end{align}
Hence, by Proposition \ref{prop:rho},
\begin{align}\label{eq:rho3}
\rho_3(\alpha) &= -\frac32\,\sum_{m=1}^\infty \frac{\sum_{k=1}^{m-1} \frac 1k}{m^2}\, \alpha^{2m}
= -\frac32\,\op{Li}_{2,1}(\alpha^2)
\end{align}
for $|\alpha|<1$.
\hspace*{\fill}$\Diamond$\medskip
\end{example}
\subsection[A general formula for rho]{A general formula for $\rho_n$}
In the general case we have
\begin{equation}\label{eq:prodcos}
\prod_{j=1}^n \cos(x_j) = 2^{-n} \sum_{\varepsilon\in\{-1,1\}^n}
\cos\left( \sum_{j=1}^n \varepsilon_j x_j \right)
\end{equation}
which follows inductively from $2\cos(a)\cos(b) = \cos(a+b)+\cos(a-b)$.
\begin{proposition}\label{prop:omegaodd}
For integers $n,m\ge0$ we have $\omega_n(2m+1) = 0$.
\end{proposition}
\begin{proof}
In light of \eqref{eq:prodcos} the summand corresponding to the indices
$k_1,\ldots,k_n$ in \eqref{eq:cn} for $\omega_n(2m+1) = 0$ is nonzero if and
only if there exists $\varepsilon\in\{-1,1\}^n$ such that $\varepsilon_1 k_1
+ \ldots + \varepsilon_n k_n = 0$.
In other words, there is a set $S\subset\{1,\ldots,n\}$ such that
\[ \sum_{j\in S} k_j = \sum_{j\not\in S} k_j. \]
Thus $k_1+\ldots+k_n=2\sum_{j\in S} k_j$ which contradicts $k_1+\ldots+k_n=2m+1$.
\end{proof}
\begin{example}[Evaluation of $\omega_4$ and $\rho_4$]\label{ex:rho4}
Proceeding as in Example \ref{ex:rho3} and employing \eqref{eq:prodcos}, we find
\begin{align}\label{eq:o4}
\omega_4(2m) &= \frac38 \sum_{\substack{i+j=m \\ k+\ell=m}} \frac{(2m)^4}{ijk\ell}
+ \frac12 \sum_{i+j+k=m} \frac{(2m)^4}{ijk\ell} \nonumber \\
&= 24 m^2 \sum_{\substack{i<m \\ j<m}} \frac1{ij}
+ 24 m^2 \sum_{i+j<m} \frac1{ij} \nonumber\\
&= 48 m^2 \sum_{i=1}^{m-1} \frac1i \sum_{j=1}^{i-1} \frac1j
+ 24 m^2 \sum_{i=1}^{m-1} \frac1{i^2}
+ 48 m^2 \sum_{i=1}^{m-1} \frac1i \sum_{j=1}^{i-1} \frac1j.
\end{align}
Consequently, for $|\alpha|<1$ and appealing to Proposition \ref{prop:rho},
\begin{align}\label{eq:rho4}
\rho_4(\alpha) = \sum_{m=1}^\infty \frac{\alpha^{2m}}{(2m)^4}\, \omega_4(2m)
= 6\,\op{Li}_{2,1,1}(\alpha^2) + \frac32\,\op{Li}_{2,2}(\alpha^2).
\end{align}
This suggests that $\rho_n(\alpha)$ is generally expressible as a sum of
polylogarithmic terms, as will be shown next.
\hspace*{\fill}$\Diamond$\medskip
\end{example}
To help general evaluation of $\omega_n(2m)$, for integers $j\ge0$ and $m\ge1$,
let us define
\begin{equation}\label{eq:defsigma1}
\sigma_j(m) :=\sum_{m_1+\ldots+m_j= m}\frac{1}{m_1\cdots m_j}.
\end{equation}
\begin{proposition}\label{prop:sigma}
For positive integers $n$, $m$ we have
\begin{equation}\label{eq:os}
\frac{\omega_n(2m) }{m^n}= \sum_{j=1}^{n-1} \binom{n}{j} \sigma_{j}(m)\sigma_{n-j}(m)
\end{equation}
where $\sigma_{j}$ is as defined in \eqref{eq:defsigma1}.
\end{proposition}
\begin{proof}
It follows from \eqref{eq:prodcos} that
\begin{equation*}
\omega_n(2m) = \sum_{k_1+\ldots+k_n=2m} \sum_{\stack{\varepsilon\in\{-1,1\}^n}{\sum_j \varepsilon_j k_j = 0}} \prod_{j=1}^n \frac{m}{k_j}.
\end{equation*}
Arguing as in Proposition \ref{prop:omegaodd} we therefore find that
\begin{equation*}
\omega_n(2m) = \sum_{j=1}^{n-1} \binom{n}{j} \sum_{\substack{k_1+\ldots+k_j=m \\ k_{j+1}+\ldots+k_n=m}} \prod_{j=1}^n \frac{m}{k_j}.
\end{equation*}
This is equivalent to \eqref{eq:os}.
\end{proof}
Moreover, we have a simple useful recursion:
\begin{proposition}\label{prop:srec}
Let $m \ge 1$. Then $\sigma_{1}(m)=1/m$ while for $j \ge2$ we have
\begin{align*}
\sigma_{j}(m) = \frac{j}{m} \sum_{r=1}^{m-1} \sigma_{j-1}(r).
\end{align*}
\end{proposition}
\begin{proof}
We have
\begin{align*}
\sigma_j(m) &= \sum_{m_1+\ldots+m_j= m}\frac{1}{m_1\cdots m_j} \\
&= \frac{j}{m} \sum_{m_1+\ldots+m_j = m}\frac{1}{m_1\cdots m_{j-1}} \\
&= \frac{j}{m} \sum_{r=1}^{m-1} \sum_{m_1+\ldots+m_{j-1} = r}\frac{1}{m_1\cdots m_{j-1}}
\end{align*}
which yields the claim.
\end{proof}
\begin{corollary}\label{cor:sigmajm}
We have
\begin{align*}
\sigma_{j}(m) = \frac{j!}{m} \sum_{m>m_1>\ldots>m_{j-1}>0} \frac{1}{m_1 \cdots m_{j-1}}.
\end{align*}
\end{corollary}
Thus, for instance, $\sigma_2(m)=2H_{m-1}/m$. From here, we easily re-obtain the
evaluations of $\omega_3$ and $\omega_4$ given in Examples \ref{ex:rho3} and
\ref{ex:rho4}. To further illustrate Propositions \ref{prop:sigma} and \ref{prop:srec}, we
now compute $\rho_5$ and $\rho_6$.
\begin{example}[Evaluation of $\rho_5$ and $\rho_6$]\label{ex:rho5}
From Proposition \ref{prop:sigma},
\begin{equation*}
\frac{\omega_5(2m) }{m^5}= 10\sigma_1(m)\sigma_4(m) + 20\sigma_2(m)\sigma_3(m).
\end{equation*}
Consequently, for $|\alpha|<1$,
\begin{align}\label{eq:rho5}
-\rho_5(\alpha) &= \sum_{m=1}^\infty \frac{\alpha^{2m}}{(2m)^5}\, \omega_5(2m) \nonumber\\
&= \frac{10\cdot4!}{32}\,\op{Li}_{2,1,1,1}(\alpha^2)
+ \frac{20\cdot2!\cdot3!}{32}\left( 3\op{Li}_{2,1,1,1}(\alpha^2) + \op{Li}_{2,1,2}(\alpha^2) + \op{Li}_{2,2,1}(\alpha^2) \right) \nonumber\\
&= 30\,\op{Li}_{2,1,1,1}(\alpha^2)
+ \frac{15}{2}\left(\op{Li}_{2,1,2}(\alpha^2) + \op{Li}_{2,2,1}(\alpha^2) \right).
\end{align}
Similarly, we have for $|\alpha|<1$,
\begin{align}\label{eq:rho6}
\rho_6(\alpha)\nonumber
&= 180\,\op{Li}_{2,1,1,1,1}(\alpha^2)
+ 45\left(\op{Li}_{2,1,1,2}(\alpha^2) + \op{Li}_{2,1,2,1}(\alpha^2) + \op{Li}_{2,2,1,1}(\alpha^2) \right) \\
&\quad+ \frac{45}{4}\op{Li}_{2,2,2}(\alpha^2).
\end{align}
From these examples the general pattern, established next, begins to
transpire.
\hspace*{\fill}$\Diamond$\medskip
\end{example}
In general, $\rho_n$ evaluates as follows:
\begin{theorem}[Evaluation of $\rho_n$]\label{thm:rhon}
For $|\alpha|<1$ and integers $n\ge2$,
\begin{align*}
\rho_n(\alpha) = \frac{(-1)^n n!}{4^n}
\sum_w 4^{\ell(w)} \op{Li}_{w}(\alpha^2)
\end{align*}
where the sum is over all indices $w=(2,a_2,a_3,\ldots,a_{\ell(w)})$ such
that $a_2,a_3,\ldots\in\{1,2\}$ and $|w|=n$.
\end{theorem}
\begin{proof}
From Proposition \ref{prop:sigma} and Corollary \ref{cor:sigmajm} we have
\begin{align*}
\rho_n(\alpha) = \frac{(-1)^n n!}{2^n} \sum_{m=1}^\infty \frac{\alpha^{2m}}{m^2}
\sum_{j=0}^{n-2} \sum_{\stack{m>m_1>\ldots>m_j>0}{m>m_{j+1}>\ldots>m_{n-2}>0}}
\frac{1}{m_1\cdots m_{n-2}}.
\end{align*}
Combining the right-hand side into harmonic polylogarithms yields
\begin{align*}
\rho_n(\alpha) = \frac{(-1)^n n!}{2^n}
\sum_{k=0}^{n-2} \sum_{\stack{a_1,\ldots,a_k\in\{1,2\}}{a_1+\ldots+a_k=n-2}}
2^{c(a)} \op{Li}_{2,a_1,\ldots,a_k}(\alpha^2)
\end{align*}
where $c(a)$ is the number of $1$s among $a_1,\ldots,a_k$.
The claim follows.
\end{proof}
\begin{example}[Special values of $\rho_n$]\label{ex:rho-vals}
Given Theorem \ref{thm:rhon}, one does not expect to be able to evaluate
$\rho_n(\alpha)$ explicitly at most points. Three exceptions are $\alpha=0$
(which is trivial), $\alpha=1$, and $\alpha=1/\sqrt{2}$. For instance we
have $\rho_4(1)= \frac{19}{240}\,\pi^4$ as well as $-\rho_5(1) =
\frac{45}{2}\zeta(5) + \frac{5}{4}\zeta(3)$ and $\rho_6(1) =
\frac{275}{1344}\pi^6 + \frac{45}{2}\zeta(3)^2$. At $\alpha= 1/\sqrt{2}$ we have
\begin{align}
\rho_4\left(\frac1{\sqrt{2}}\right)
&= \frac{7}{16}\log^4 2 + \frac{3}{16}\pi^2 \log^2 2 - \frac{39}{8}\zeta(3) \log 2
+ \frac{13}{192}\pi^4 - 6\op{Li}_4\left(\tfrac12\right).
\end{align}
For $n \ge 5$ the expressions are expected to be more complicated.
\hspace*{\fill}$\Diamond$\medskip
\end{example}
\subsection{Reducing harmonic polylogarithms of low weight}\label{sec:lireductions}
Theorems \ref{thm:mun} and \ref{thm:rhon} take us closer to a closed form for $\mu_n(1+x+y)$. As $\rho_n$ are expressible in terms of multiple harmonic polylogarithms of weight $n$, it remains to supply reductions those of low
weight. Such polylogarithms are reduced by the use of the differential operators
\begin{equation*}
(D_0 f)(x)= x f'(x) \quad \mbox{and} \quad (D_1 f)(x)=(1-x)f'(x)
\end{equation*}
depending on whether the outer index is `$2$' or `$1$' \cite{b3l}.
\begin{enumerate}
\item As was known to Ramanujan, and as studied further in
\cite[\S8.1]{goldbach}, for $0<x<1$,
\begin{align}\label{eq:z21}
\op{Li}_{2,1}(x) &= \frac12\, \log^2(1-x)\log(x)+ \log(1-x)\op{Li}_2(1-x) \nonumber\\
&\quad -\op{Li}_3(1-x) + \zeta(3).
\end{align}
Equation \eqref{eq:z21}, also given in \cite{lewin2}, provides a useful
expression numerically and symbolically. For future use, we also record
the relation, obtainable as in \cite[\S6.4 \& \S6.7]{lewin2},
\begin{align}\label{eq:z21r}
\Re\op{Li}_{2,1}\left(\frac{1}{x}\right) + \op{Li}_{2,1}(x)
&= \zeta(3) -\frac16 \log^3 x + \frac12\pi^2 \log x \nonumber\\
&\quad -\op{Li}_2(x) \log x + \op{Li}_3(x) \qquad \mbox{for~} 0 < x < 1.
\end{align}
\item For $\op{Li}_{2,2}$ we work as follows.
As $(1-x) \op{Li}'_{1,2}(x) =\op{Li}_2 \left(x \right) $, integration yields:
\begin{align}
\op{Li}_{1,2}(x) &= 2\op{Li}_3(1-x)-\log(1-x)\op{Li}_2(x) -2\log(1-x)\op{Li}_2(1-x) \nonumber\\
&\quad- \log(1-x)^2\log(x)-2\zeta(3).
\end{align}
Then, since $x\op{Li}'_{2,2}(x) =\op{Li}_{1,2}(x)$, on integrating again we obtain
$\op{Li}_{2,2}(x)$ in terms of polylogarithms up to order four. We appeal to
various formulae in \cite[\S6.4.4]{lewin2} to arrive at
\begin{align*}\op{Li}_{2,2}(t) &=\nonumber \frac 12\,
\log^2 (1-t) \log^2 t
-2\zeta(2)\log (1-t) \log t -2\zeta(3) \log t -\frac12\,\op{Li}_2^2 (t) \\& +2\op{Li}_3 \left(1-t
\right)\log t -2 \int_{0}^{t}\!{\frac{ \op{Li}_2 \left(x \right)\log x
}{1-x}}{\,\mathrm{d} x}-\int_{0}^{t}\!{\frac{\log \left( 1-x \right)\log^2 x}{1-x}}{\mathrm{d} x}.
\end{align*}
Expanding the penultimate integral as a series leads to
\[\int_{0}^{t}\!{\frac{ \op{Li}_2 \left(x \right)\log x
}{1-x}}{\,\mathrm{d} x}=\op{Li}_{1,2}(t) \log t-\op{Li}_{2,2}(t),\]
and so to
\begin{align}
\op{Li}_{2,2}(t) &= \frac 12 \op{Li}_2^2 (t)-2\zeta(3)\ln t +2 \op{Li}_3 (1-t)\log t-\frac 12\,
\log^2 (1-t) \log^2 t \nonumber\\
&\quad- 2 \log (1-t) \op{Li}_2 (1-t) \log t+\int_{0}^{t}\!\frac{\log\left( 1-x \right) \log^2 x}{1-x}{\,\mathrm{d} x}.
\end{align}
Now \cite[A3.4 Eq. (12)]{lewin2} will evaluate the final integral and we deduce:
\begin{align}
\op{Li}_{2,2}(t) &= -\frac1{12}\, \log^4(1-t)+ \frac13 \log^3(1-t) \log t -\zeta(2) \log^2 (1-t) \nonumber\\
&\quad+ 2\log(1-t) \op{Li}_3 (t) -2\,\zeta ( 3) \log(1-t) -2\,\op{Li}_4(t) \nonumber \\
&\quad- 2\op{Li}_4 \left({\frac{t}{t-1}} \right) +2\,\op{Li}_4(1-t) -2\zeta(4) +\frac12\, \op{Li}_2^2 (t).
\end{align}
\item The form for $\op{Li}_{3,1}(t)$ is obtained in the same way but starting
from $\op{Li}_{2,1}(t)$ as given in \eqref{eq:z21}. This leads to:
\begin{align}
2\,\op{Li}_{3,1}(t)+\op{Li}_{2,2}(t)= \frac12\, \op{Li}_2^2(t).
\end{align}
This symmetry result, and its derivative
\begin{align}
2\,\op{Li}_{2,1}(t)+\op{Li}_{1,2}(t)= \op{Li}_1(t) \op{Li}_2(t),
\end{align}
are also obtained in \cite[Cor. 2 \& Cor. 3]{zlobin} by other methods.
\item Since $\op{Li}_{2,1,1}(x) =\int_0^x\op{Li}_{1,1,1}(t)/t\,\mathrm{d} t$ and
$\op{Li}_{1,1,1}(x) =\int_0^x\op{Li}_{1,1}(t)/(1-t)\,\mathrm{d} t$, we first compute
$\op{Li}_{1,1}(x)=\log^2(1-x)/2$ and so $\op{Li}_{1,1,1}(x)=-\log^3(1-x)/6$ (the
pattern will be clear). Hence
\begin{align}
\op{Li}_{2,1,1}(x) &=-\frac16 \int_0^x\,\log^3(1-t)\frac{\mathrm{d} t}t \nonumber\\
&=\frac{\pi^4}{90}-\frac16 \log(1-t)^3 \log t - \frac12 \log(1-t)^2 \op{Li}_2(1-t) \nonumber\\
&\quad+ \log(1-t)\op{Li}_3(1-t)-\op{Li}_4(1-t).
\end{align}
\item In general,
\begin{equation}
\op{Li}_{\{1\}_n}(x) = \frac{(-1)^n}{n!} \log(1-x)^n,
\end{equation}
and therefore
\begin{align}
\op{Li}_{2,\{1\}_{n-1}}(x) &= \frac{(-1)^n}{n!} \int_0^x \log(1-t)^n \frac{\mathrm{d} t}{t} \nonumber \\
&= \zeta(n+1)-\sum_{m=0}^n \frac{(-1)^{n-m}}{(n-m)!} \log(1-x)^{n-m} \op{Li}_{m+1}(1-x).
\end{align}
\end{enumerate}
We have, inter alia, provided closed reductions for all multiple polylogarithms
with weight less than five. One does not expect such complete results
thereafter.
The reductions presented in this section allow us to express $\rho_3$ and
$\rho_4$ in terms of polylogarithms of depth $1$. Equation \eqref{eq:z21}
treats $\rho_3$ while \eqref{eq:rho4} leads to
\begin{align}\label{eq:rho4b}
\rho_4\left(\alpha^2\right)&=3\left(\op{Li}_3 \left(\alpha^{2} \right)-\zeta(3) +\op{Li}_3
\left( 1-{\alpha}^{2}\right) \right) \log \left( 1-{\alpha}^{2} \right)-\frac18\,\log^4 \left(1 -\alpha^{2} \right) \nonumber \\
&\quad+ 3\zeta(4)- 3\,\op{Li}_4\left(\frac{-\alpha^2}{1-\alpha^2} \right)- 3\,\op{Li}_4\left(\alpha^{2} \right)-
3\,\op{Li}_4\left(1-\alpha^{2} \right)+\frac34\op{Li}_2^2\left(1-\alpha^{2} \right) \nonumber\\
&\quad- \log \alpha \,\log^3 \left( 1-\alpha^{2}\right)
-\left( \frac{\pi^2}4+3\op{Li}_2 \left( 1-\alpha^{2} \right)\right)\log^2 \left( 1-\alpha^{2} \right).
\end{align}
\section[Explicit evaluations of mun(1+x+y) for small n]{Explicit evaluations of $\mu_n(1+x+y)$ for small
$n$}\label{sec:mu23}
We now return to the explicit evaluation of the multiple Mahler measures
$\mu_k(1+x+y)$. The starting point for this section is the evaluation of
$\mu_2(1+x+y)$ from \cite{logsin1} which is reviewed in Section \ref{sec:mu2}
and was derived alternatively in Theorem \ref{thm:mu21eps}. Building on this,
we then present an informal evaluation of $\mu_3(1+x+y)$ in Section
\ref{sec:mu3}. A conjectural evaluation of $\mu_4(1+x+y)$ is presented in
equation \eqref{eq:mu4} of the Conclusion.
\subsection[Evaluation of mu2(1+x+y)]{Evaluation of $\mu_2(1+x+y)$}\label{sec:mu2}
\begin{theorem}[Evaluation of $\mu_2(1+x+y)$]\label{thm:mu21xy}
We have
\begin{align}\label{eq:mu2ls}
\mu_2(1+x+y) = \frac3\pi \Ls{3}{\frac{2\pi}{3}} + \frac{\pi^2}{4}.
\end{align}
\end{theorem}
By comparison, Smyth's original result may be written as (see \cite{logsin1})
\begin{equation}\label{eq:mu1xy}
\mu_1(1+x+y) = \frac{3}{2\pi} \Ls{2}{\frac{2\pi}{3}}
= \frac{1}{\pi}\mcl{2}.
\end{equation}
We recall from \cite{logsin1} that the evaluation in Theorem
\ref{thm:mu21xy} proceeded by first establishing the following dilogarithmic
form:
\begin{proposition}[A dilogarithmic representation]\label{prop:mu2}
We have:
\begin{enumerate}[(a)]
\item\label{propa}
\begin{equation}\label{eq:z2}
\frac 2\pi \int_{0}^{\pi} \Re\op{Li}_2\left(4\sin^2\theta\right) \mathrm{d}\theta = 2\zeta(2).
\end{equation}
\item\label{propb}
\begin{equation}\label{eq:mu2}
\mu_2(1+x+y) =
\frac{\pi^2}{36}+\frac 2\pi \int_{0}^{\pi/6} \op{Li}_2\left(4\sin^2\theta\right) \mathrm{d}\theta.
\end{equation}
\end{enumerate}
\end{proposition}
We include the proof from \cite{logsin1} as it is instructive for evaluation of
$\mu_3(1+x+y)$.
\begin{proof}
For \eqref{propa} we define $\tau(z):= \frac 2\pi \int_{0}^{\pi}
\op{Li}_2\left(4z\sin^2\theta\right) \mathrm{d}\theta.$ This is an analytic function of
$z$. For $|z|<1/4$ we may use the original series for $\op{Li}_2$ and expand
term by term using Wallis' formula to derive
\begin{align*}
\tau(z) &= \frac 2\pi\,\sum_{n\ge 1}\frac{(4z)^n}{n^2}\int_{0}^{\pi}\,\sin^{2n}\theta\,\mathrm{d}\theta
= 4z\,\pFq43{1, 1, 1, \frac 32}{2, 2, 2}{4z} \\
&= 4\op{Li}_2\left( \frac12-\frac12\sqrt{1-4z} \right) - 2 \log\left( \frac12+\frac12\sqrt{1-4z} \right)^2.
\end{align*}
The final equality can be obtained in \emph{Mathematica} and then verified by
differentiation. In particular, the final function provides an analytic
continuation and so we obtain $\tau(1)=2\zeta(2) + 4i\mcl{2}$ which yields
the assertion.
For \eqref{propb}, commencing much as in \cite[Thm. 11]{klo}, we write
\begin{equation*}
\mu_2(1+x+y) = \frac{1}{4\pi^2} \int_{-\pi}^{\pi} \int_{-\pi}^{\pi}
\Re\log\left(1-2\sin(\theta)\mathrm{e}^{i\,\omega}\right)^2 \,\mathrm{d}\omega\,\mathrm{d}\theta.
\end{equation*}
We consider the inner integral
$\rho(\alpha):=\int_{-\pi}^{\pi}\left(\Re\log\left(1-\alpha\,\mathrm{e}^{i\,\omega}\right)\right)^2\,\mathrm{d} \omega$ with $\alpha := 2\sin\theta$. For
$|\theta| \le \pi/6$ we directly apply Parseval's identity to obtain
\begin{equation}\label{eq:mrhoA}
\rho(2\sin \theta) = \pi \op{Li}_2\left(4\sin^2\theta\right).
\end{equation}
which is equivalent to \eqref{eqn:rho2} since $\rho(\alpha) =
2\pi\rho_2(\alpha)$. In the remaining case we write
\begin{align}\label{eq:mrhoB}
\rho(\alpha) &=
\int_{-\pi}^{\pi}\left\{\log|\alpha| + \Re\log \left(1-\alpha^{-1}\,\mathrm{e}^{i\,\omega}\right)\right\}^2 \,\mathrm{d}\omega \nonumber\\
&= 2\pi\,\log^2 |\alpha|- 2\log|\alpha|\int_{-\pi}^{\pi}\log \left|1-\alpha^{-1}\,\mathrm{e}^{i\,\omega}\right|\,\mathrm{d} \omega
+ \pi \op{Li}_2\left(\frac{1}{\alpha^2}\right) \nonumber\\
&= 2\pi\,\log^2 |\alpha| + \pi \op{Li}_2\left(\frac{1}{\alpha^2}\right),
\end{align}
where we have appealed to Parseval's and Jensen's formulae. Thus,
\begin{equation}\label{eq:mrhoC}
\mu_2(1+x+y) = \frac 1\pi \int_{0}^{\pi/6} \op{Li}_2\left(4\sin^2\theta\right) \mathrm{d}\theta
+ \frac 1\pi \int_{\pi/6}^{\pi/2} \op{Li}_2\left(\frac{1}{4\sin^2\theta}\right) \mathrm{d}\theta
+ \frac{\pi^2}{54},
\end{equation}
since $\frac 2\pi\int_{\pi/6}^{\pi/2}\log^2 \alpha\,\mathrm{d}\theta = \mu_2(1+x+y_*) =
\frac{\pi^2}{54}$. Now, for $\alpha >1$, the functional equation in
\cite[A2.1 (6)]{lewin}
\begin{align}\label{eq:dilog}
\op{Li}_2 (\alpha) +\op{Li}_2 (1/\alpha) +\frac 12\,\log^2 \alpha &=2\zeta(2)+ i\pi \log \alpha
\end{align}
gives:
\begin{equation}\label{eq:fed1}
\int_{\pi/6}^{\pi/2}\left\{\Re\op{Li}_2 \left(4 \sin^2 \theta \right)
+ \op{Li}_2 \left(\frac1{4 \sin^2 \theta} \right)\right\}\,\mathrm{d}\theta
= \frac{5}{54} \pi^3.
\end{equation}
We now combine \eqref{eq:z2}, \eqref{eq:fed1} and \eqref{eq:mrhoC} to deduce
the desired result in \eqref{eq:mu2}.
\end{proof}
\subsection[Evaluation of mu3(1+x+y)]{Evaluation of $\mu_3(1+x+y)$}\label{sec:mu3}
In this section we provide a remarkably concise closed form of
$\mu_3(1+x+y)$. We were led to this form by the integer relation algorithm
PSLQ \cite{bbg} (see Example \ref{ex:num} for some comments on obtaining high
precision evaluations), and by considering the evaluation \eqref{eq:mu2ls} of
$\mu_2(1+x+y)$.
The details of formalization are formidable --- at least by the route
chosen here --- and so we proceed more informally leaving three conjectural identities.
\begin{conjecture}[Evaluation of $\mu_3(1+x+y)$]\label{thm:mu3f2} We have
\begin{align}\label{eq:mu3ls}
\mu_3(1+x+y) \stackrel{?[1]}{=} \frac{6}{\pi}\Ls{4}{\frac{2\pi}{3}} - \frac9\pi\mcl{4}
- \frac\pi4\mcl{2} - \frac{13}{2}\zeta(3).
\end{align}
\end{conjecture}
This evaluation is equivalent to the conjectural identities \eqref{eq:lsc23}
and \eqref{eq:lsc23ti}.
\begin{proof}
We first use Theorem \ref{thm:mun} to write
\begin{align}\label{eq:mu3}
\mu_3(1+x+y) &= \frac{2}{\pi} \int_{0}^{\pi/6} \rho_3(2\sin\theta) \,\mathrm{d}\theta
+ \frac{2}{\pi} \int_{\pi/6}^{\pi/2} \rho_3\left(\frac{1}{2\sin\theta}\right) \,\mathrm{d}\theta \\
&\quad+ \frac{3}{\pi} \int_{\pi/6}^{\pi/2} \log(2\sin\theta)\,\op{Li}_2\left(\frac 1 {4\sin^2 \theta} \right) \,\mathrm{d}\theta
- \zeta(3) + \frac{9}{2\pi}\mcl{4}, \nonumber
\end{align}
on appealing to Examples \ref{ex:pi} and \ref{ex:lspi3}.
Now the functional equation for the dilogarithm \eqref{eq:dilog} as used above
and knowledge of $\Ls{n}{\pi/3}$ (see \cite{logsin1,logsin-evaluations}) allow
us to deduce
\begin{align}
\frac3\pi \int_{0}^{\pi/6} \log(2\sin\theta)&\,\op{Li}_2\left(4\sin^2 \theta \right) \mathrm{d}\theta
+ \frac3\pi \int_{0}^{\pi/6} \log(2\sin\theta)\,\op{Li}_2\left(\frac{1}{4\sin^2 \theta}\right) \mathrm{d}\theta \nonumber\\
&= \frac32\zeta(3)-\frac{\pi}2 \mcl{2}+\frac{27}{2\pi}\mcl{4}, \label{eq:pi6}\\
\frac3\pi \int_{\pi/6}^{\pi/2} \log(2\sin\theta)&\,\Re\op{Li}_2\left(4\sin^2 \theta\right) \mathrm{d}\theta
+ \frac3\pi \int_{\pi/6}^{\pi/2} \log(2\sin \theta)\,\Re\op{Li}_2\left(\frac{1}{4\sin^2 \theta}\right) \mathrm{d}\theta \nonumber\\
&= 3\zeta(3)+\frac{\pi}2\,\mcl{2}-\frac{27}{2\pi}\mcl{4} \label{eq:pi6b}
\end{align}
Moreover, we have
\begin{align}
\frac3\pi \left\{\int_{0}^{\pi/6} + \int_{\pi/6}^{\pi/2}\right\}
\log(2\sin\theta)\,\Re\op{Li}_2\left(4\sin^2 \theta \right) \mathrm{d}\theta
&\stackrel{?[2]}{=} \frac72\zeta(3) - \pi\mcl{2}, \label{eq:pi2a} \\
\frac3\pi \left\{\int_{0}^{\pi/6} + \int_{\pi/6}^{\pi/2}\right\}
\log(2\sin\theta)\,\Re\op{Li}_2\left(\frac{1}{4\sin^2 \theta}\right) \mathrm{d}\theta
&\stackrel{?[2]}{=} \zeta(3) + \pi\mcl{2}, \label{eq:pi2b}
\end{align}
which are provable as was \eqref{eq:z2} because for $|z| < 1/2$ we have
\[ \frac 1\pi\, \int_{0}^{\pi }\!\log \left( 2\,\sin \frac\theta2 \right) \op{Li}_2\left(4\,{z}^{2} \sin^2 \frac\theta2\right) \,\mathrm{d}\theta
=\sum_{n=1}^{\infty } \binom{2n}{n} \frac{\sum_{k=1}^{2n} \frac{(-1)^k}{k}}{n^2}\,z^{2n}. \]
(The latter is derivable also from \eqref{eq:pi6}, \eqref{eq:pi6b} and \eqref{eq:pi2a}.)
Thence, \eqref{eq:pi6}, \eqref{eq:pi6b} and \eqref{eq:pi2a} together establish
that the equality
\begin{align}\label{eq:mu3f}
\frac3\pi \int_{\pi/6}^{\pi/2}\log(2\sin\theta)\,\op{Li}_2\left(\frac{1}{4\sin^2 \theta} \right) \mathrm{d}\theta
&=\frac23\zeta(3)+\frac{7\pi}{12}\mcl{2}-\frac{17}{2\pi}\mcl{4}
\end{align}
is true as soon as we establish
\begin{align}\label{eq:I3}
I_3 := \frac3\pi \int_{0}^{\pi/6} \log(2\sin\theta)
\op{Li}_2\left(4\sin^2 \theta \right) \mathrm{d}\theta
\stackrel{?[3]}{=}\frac76\,\zeta(3)-\frac{11\pi}{12}\,\mcl{2}+5\mcl{4}.
\end{align}
This can be achieved by writing the integral as
\begin{align*}
I_3 &= \frac3\pi\,\sum_{n=1}^\infty\frac 1{n^2} \int_0^1\!\frac{ s^{2n}}{\sqrt {4-{s}^{2}}}\log s\,{\mathrm{d} s}
\end{align*}
and using the binomial series to arrive at
\begin{align}\label{3sum}
I_3&=-\frac{3}{2\pi}\,\sum_{m=0}^\infty\frac{\binom{2m}{m}}{4^{2m}} \sum_{n=1}^\infty\frac 1{n^2 \left(1+2(n+m)\right)^2}.
\end{align}
Now \eqref{3sum} can in principle be evaluated to the asserted result
in \eqref{eq:I3}; so this leaves us to deal with the two terms involving
$\rho_3$.
These two terms are in turn related by
\begin{align}
\frac2\pi \int_{0}^{\pi/6}\op{Li}_{2,1}\left(4\sin^2 \theta \right) \mathrm{d}\theta
&+ \frac2\pi \int_{0}^{\pi/6} \Re\op{Li}_{2,1}\left(\frac{1}{4\sin^2 \theta} \right) \mathrm{d}\theta \nonumber\\
&= \frac1{9}\left\{\zeta(3) - \pi\mcl{2} + \frac{6}{\pi}\mcl{4}\right\}, \label{eq:mu36}
\end{align}
as we see by integrating \eqref{eq:z21r}. Likewise,
\begin{align}
\frac2\pi \int_{\pi/6}^{\pi/2} \Re\op{Li}_{2,1}\left(4\sin^2 \theta \right) \mathrm{d}\theta
&+ \frac2\pi \int_{\pi/6}^{\pi/2} \op{Li}_{2,1}\left(\frac{1}{4\sin^2 \theta} \right) \mathrm{d}\theta \nonumber\\
&= \frac1{9}\left\{2\zeta(3) - 5\pi\mcl{2} - \frac{6}{\pi}\mcl{4}\right\}. \label{eq:mu36b}
\end{align}
Also, using \eqref{eq:z21} we arrive at
\begin{align}\label{eq:rint}
\frac2\pi \int_{0}^{\pi/6} \op{Li}_{2,1}\left(4\sin^2 \theta \right) \mathrm{d}\theta
&= \frac{20}{27}\zeta(3) - \frac{8\pi}{27}\mcl{2} + \frac{4}{9\pi}\mcl{4} \nonumber\\
&\quad+ \frac1\pi \int_{0}^{\pi/3} \log^2 \left(1-4\sin^2\frac\theta 2 \right)
\log \left( 2\sin\frac\theta 2 \right) \mathrm{d}\theta,
\end{align}
and
\begin{equation}\label{eq:rintb}
\frac2\pi \int_{0}^{\pi/2} \Re\op{Li}_{2,1}\left(4\sin^2 \theta \right)\mathrm{d} \theta
= \frac{1}{3}\zeta(3) - \frac{2\pi}{3}\mcl{2}.
\end{equation}
We may now establish --- from \eqref{eq:mu3f}, \eqref{eq:mu36},
\eqref{eq:mu36b}, \eqref{eq:rint}, \eqref{eq:rintb} and \eqref{eq:mu3} ---
that
\begin{align}
\mu_3(1+x+y) &= \frac{43}{18}\,\zeta(3)-\frac{47\pi}{36}\,\mcl{2}-\frac{13}{3\pi}\,\mcl{4} \nonumber \\
&\quad+ \frac{2}\pi\int_{0}^{\pi/3} \log^2 \left(1-4\sin^2\frac\theta 2 \right)
\log \left( 2\sin\frac\theta2 \right) \mathrm{d}\theta.
\end{align}
Hence, to prove \eqref{eq:mu3ls} we are reduced to verifying that
\begin{align}
- \frac{1}\pi \Ls{4}{\frac{2\pi}3}
&\stackrel{?[4]}{=} -\frac{37}{54}\zeta(3) + \frac{7\pi}{27}\mcl{2} - \frac7{9\pi}\mcl{4} \nonumber\\
&\quad+ \frac{1}{2\pi} \int_{0}^{\pi/3} \log^2 \left(1-4\sin^2\frac\theta 2 \right)
\log \left( 2\sin\frac\theta 2 \right) \mathrm{d}\theta.
\end{align}
which completes the evaluation.
\end{proof}
\begin{remark}
By noting that, for integers $n\ge2$,
\begin{equation*}
\Cl{n}{\frac\pi3} = \left( \frac{1}{2^{n-1}} + (-1)^n \right) \Cl{n}{\frac{2\pi}{3}},
\end{equation*}
the arguments of the Clausen functions in the evaluation \eqref{eq:mu3ls} of
$\mu_3(1+x+y)$ may be transformed to $\frac{2\pi}{3}$.
Many further variations are possible. For instance, it follows from \cite{logsin-evaluations} that
\begin{align}
\Ls{4}{\frac{2 \pi} 3}
& = \frac{31}{18}\,\pi\zeta(3)+\frac{\pi^2}{12}\Cl{2}{\frac{2\pi}3}
-\frac32\Cl{4}{\frac{2\pi}3}+6\Cl{2,1,1}{\frac{2\pi}3}
\end{align}
in terms of multi Clausen values.
\hspace*{\fill}$\Diamond$\medskip
\end{remark}
\section{Proofs of two conjectures of Boyd}\label{sec:boyd}
We now use log-sine integrals to recapture the following evaluations
conjectured by Boyd in 1998 and first proven in \cite{two} using \emph{Bloch-Wigner}
logarithms. Below ${\rm L}_{-n}$ denotes a primitive $\rm L$-series and $\mathrm{G}$
is Catalan's constant.
\begin{theorem}[Two quadratic evaluations]\label{thm:boyd} We have
\begin{align}
\mu(y^2(x+1)^2+y(x^2+6x+1)+(x+1)^2)
&= \frac{16}{3\pi}\,{\rm L}_{-4}(2) = \frac{16}{3\pi}\,\mathrm{G}, \label{ex:bl1}\\
\intertext{as well as}
\mu(y^2(x+1)^2+y(x^2-10x+1)+(x+1)^2)
&= \frac{5\sqrt{3}}{\pi}\,{\rm L}_{-3}(2) = \frac{20}{3\,\pi}\mcl{2}. \label{ex:bl2}
\end{align}
\end{theorem}
\begin{proof}
Let $P_c = y^2(x+1)^2+y(x^2+2cx+1)+(x+1)^2$ and $\mu_c = \mu(P_c)$ for a real
variable $c$. We set $x=\mathrm{e}^{2\pi it}$, $y=\mathrm{e}^{2\pi iu}$ and note that
\begin{align*}
|P_c| &= |(x+1)^2(y^2+y+1)+2(c-1)xy| \\
&= \left|(x+x^{-1}+2)(y+1+y^{-1})+2(c-1)\right| \\
&= |2(\cos(2\pi t)+1)(2\cos(2\pi u)+1)+2(c-1)| \\
&= 2\,|c+2\cos(2\pi u)+(1+2\cos(2 \pi u))\cos(2\pi t)|.
\end{align*}
It is known that (see \cite[\S4.224, Ex. 9]{gr}), for real $a, \,b$ with $|a|\ge |b|>0,$
\begin{align}\label{eq:lab}
\int_0^1 \log\left|2a+2b\cos(2\pi \theta)\right| \,\mathrm{d}\theta
= \log\left( |a|+\sqrt{a^2-b^2}\right).
\end{align}
Applying this, with $a=c+2\cos(2\pi u)$ and $b=1+2\cos(2 \pi u))$ to $\int_0^1 |P_c| \,\mathrm{d} t,$ we get
\begin{equation}\label{eq:mc}
\mu_c = \int_0^1 \log\left|c+2\cos(2\pi u) + \sqrt{({c}^{2}-1)+4(c-1) \cos(2\pi u)}\right| \,\mathrm{d} u.
\end{equation}
If $c^2-1=\pm 4(c-1)$, that is if $c=3$ or $c=-5$, then the surd is a perfect square and also $|a|\ge |b|.$
(a) When $c=3$ for (\ref{ex:bl1}), by symmetry, after factorization we obtain
\begin{align*}
\mu_3 &= \frac 1\pi \int_0^\pi \log(1+4|\cos \theta|+4|\cos^2 \theta|) \,\mathrm{d}\theta
= \frac4\pi \int_0^{\pi/2} \log(1+2\cos\theta) \,\mathrm{d}\theta \\
&= \frac4\pi \int_0^{\pi/2} \log\left(\frac{2\sin\frac{3\theta}2}{2\sin\frac{\theta}2}\right)\,\mathrm{d}\theta
= \frac4{3\,\pi}\,\left(\Ls{2}{\frac{3\pi}2}-3\Ls{2}{\frac{\pi}2}\right) \\
&= \frac{16}{3}\,\frac{{\rm L}_{-4}(2)}{\pi}
\end{align*}
as required, since $\Ls{2}{\frac{3\pi}2}=-\Ls{2}{\frac{\pi}2}= {\rm L}_{-4}(2)$, which is Catalan's constant $\mathrm{G}$.
(b) When $c=-5$ for (\ref{ex:bl2}), we likewise obtain
\begin{align*}
\mu_{-5} &= \frac 2\pi \int_0^\pi \log\left(\sqrt 3+2\sin\theta\right) \,\mathrm{d}\theta
= \frac2\pi \int_{\pi/3}^{4\pi/3} \log\left(\sqrt 3+2\sin\left(\theta-\frac\pi 3\right)\right) \,\mathrm{d}\theta \\
&= \frac2\pi \int_{\pi/3}^{4\pi/3} \left\{\log2\left(\sin\frac\theta2\right)
+ \log2\left(\sin\frac{\theta+\frac\pi3}2\right)\right\} \,\mathrm{d}\theta \\
&= \frac2\pi \int_{\pi/3}^{4\pi/3} \log2\left(\sin\frac\theta2\right) \,\mathrm{d}\theta
+ \frac2\pi \int_{2\pi/3}^{5\pi/3}\log2\left(\sin\frac\theta2\right) \,\mathrm{d}\theta \\
&= \frac4\pi \mcl{2} - \frac4\pi \Cl{2}{\frac{4\pi}{3}}
= \frac{20}{3\pi} \mcl{2},
\end{align*}
since $\Cl{2}{\frac{4\pi}{3}} = -\frac23 \mcl{2}$ and so we are done.
\end{proof}
When $c=1$ the cosine in the surd disappears, and we obtain $\mu_1=0$, which is
trivial as in this case the polynomial factorizes as $(1+x)^2(1+y+y^2)$. For
$c=-1$ we are able, with some care, to directly integrate \eqref{eq:mc} and so
to obtain an apparently new Mahler measure:
\begin{theorem}We have
\begin{align}
\mu_{-1} &= \mu\left((x+1)^2(y^2+y+1)-2xy\right) \\
&= \frac{1}{\pi}\,\left\{ \frac12 B\left( \frac14, \frac14 \right)
\pFq32{\frac14,\frac14,1}{\frac34,\frac54}{\frac14}
- \frac16 B\left( \frac34, \frac34 \right)
\pFq32{\frac34,\frac34,1}{\frac54,\frac74}{\frac14}\right\}\nonumber.
\end{align}
\end{theorem}
We observe that an alternative form of $\mu_{-1}$ is given by
\[ \mu_{-1} = \mu\left(\left(x+1/x+2\sqrt{1/x}\right)(y+1/y+1)-2\right). \]
\begin{remark}
Equation \eqref{eq:lab} may be applied to other conjectured Mahler measures.
For instance, $\mu(1+x+y+1/x+1/y)=.25133043371325\ldots$ was
conjectured by
Deninger \cite{finch} to evaluate in $L$-series terms as
\begin{equation}
\mu(1+x+y+1/x+1/y) =15 \sum_{n=1}^\infty \frac{a_n}{n^2},
\end{equation}
where $ \sum_{n=1}^\infty a_n q^n = \eta(q) \eta(q^3) \eta(q^5) \eta(q^{15}).$
Here $\eta$ is the \emph{Dirichlet eta-function}:
\begin{equation}\label{eq:eta}
\eta(q):=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}
This has very recently been proven in \cite{rz}.
Application of \eqref{eq:lab} shows that
\begin{equation*}
\mu(1+x+y+1/x+1/y) =
\frac1\pi\, \int_0^{\pi/3}
\log\left(\frac{1 + 2\cos\theta}2 + \sqrt{\left(\frac{1+2\cos\theta}2\right)^2 - 1}
\right) \,\mathrm{d}\theta,
\end{equation*}
but the surd remains an obstacle to a direct evaluation.
\qed
\end{remark}
\section{Conclusion}
To recapitulate, $\mu_k(1+x+y)=W_3^{(k)}(0)$ has been evaluated in terms of
log-sine integrals for $1 \le k \le 3$. Namely,
\begin{align}\label{eq:mu2A2}
\mu_1(1+x+y) &= \frac{3}{2\pi} \Ls{2}{\frac{2\pi}{3}}, \\
\mu_2(1+x+y) &= \frac3\pi \Ls{3}{\frac{2\pi}{3}} + \frac{\pi^2}{4}, \\
\mu_3(1+x+y) &\stackrel{?[1]}{=} \frac{6}{\pi}\Ls{4}{\frac{2\pi}{3}} - \frac9\pi\mcl{4}
- \frac\pi4\mcl{2} - \frac{13}{2}\zeta(3).
\end{align}
Hence it is reasonable to ask whether $\mu_4(1+x+y)$ and higher Mahler measures
have evaluations in similar terms.
\begin{example}[Evaluation of $\mu_4(1+x+y)$]\label{ex:mu4}
This question is taken up in \cite{logsin4} where it is found that
\begin{align}\label{eq:mu4}
\pi \mu_4(1+x+y) &\stackrel{?[5]}{=} 12 \Ls{5}{\frac{2\pi}{3}} -
\frac{49}{3} \Ls{5}{\frac\pi3} + 81\Gl{4,1}{\frac{2\pi}{3}} \\
&\quad+ 3\pi^2 \Gl{2,1}{\frac{2\pi}{3}}+ 2\zeta(3)\Cl{2}{\frac\pi3} +
\pi\Cl{2}{\frac\pi3}^2 - \frac{29}{90}\pi^5. \nonumber
\end{align}
in terms of generalized Glaisher and Clausen values.
\hspace*{\fill}$\Diamond$\medskip
\end{example}
We close with numerical values for these quantities.
\begin{example
\label{ex:num}
By computing higher-order finite differences in the right-hand side of
\eqref{eq:hyper} we have obtained values for $\mu_n(1+x+y)$ to several
thousand digits. To confirm these values we have evaluated the
double-integral \eqref{eq:mu-trigintegral} to about $250$ digits for all
$n\le8$. These are the results for $\mu_k:=\mu_k(1+x+y)$ to fifty digits:
\begin{align}
\mu_2 &= 0.41929927830117445534618570174886146566170299117521,\\
\mu_3 &= 0.13072798584098927059592540295887788768895327503289,\\
\mu_4 &= 0.52153569858138778267996782141801173128244973155094,\\
\mu_5 &= -0.46811264825699083401802243892432823881642492433794.
\end{align}
These values will allow a reader to confirm many of our results numerically.
\hspace*{\fill}$\Diamond$\medskip
\end{example}
\paragraph{Acknowledgements}
We thank David Bailey for his assistance with the two-dimensional quadratures
in Example \ref{ex:num}. Thanks are due to Yasuo Ohno and Yoshitaka Sasaki for
introducing us to the relevant papers during their recent visit to CARMA.
|
\section{Introduction}
Andrei Sakharov~\cite{Sak} pointed out that a primordial particle number asymmetry is produced given three ingredients: symmetry violation, C and CP violation, as well as a departure from thermal equilibrium. The observed baryon asymmetry suggests that these conditions have been met in the history of the Universe. Since global symmetry violation is expected to occur in the early Universe, we expect the existence of primordial asymmetries to be rather generic.
The dominant contributions to the observed matter density in the Universe are due to dark matter and baryons. Their density fractions are remarkably well determined from cosmic microwave background data from WMAP7~\cite{WMAP7}:
\begin{eqnarray}\label{Omegas}
\Omega_{DM} h^{2} =0.1109 \pm 0.0056\quad\quad\quad\Omega_{B}h^{2} =0.02258^{+0.00057}_{-0.00056}.
\end{eqnarray}
Yet, very little remains known about the nature of dark matter (DM). In most theoretical efforts, the problems of dark matter and baryogenesis are treated separately, with the implicit assumption that the comparable densities of these two types of matter is simply a coincidence. However, the fact that the abundances of baryons and DM are just a factor $\sim5$ apart may be an indication of a common underlying origin. The paradigm of asymmetric dark matter (ADM) has been suggested as a way of linking the asymmetries, and thus the abundances in the dark and visible sectors; see~\cite{ADM} and references therein~\footnote{There are also many other models seeking to link the abundances of baryons and DM~\cite{Dodelson,Kaplan,Thomas,enqvist,Kitano,GravQ,Hylogenesis,Allahverdi,gaugedb,mcd,gu}.}.
In ADM two distinct scenarios are typically considered~\cite{nussinov,Barr:1990ca,hooper,Darkogenesis, adm16,adm15,adm14,adm13,adm12,adm11,adm10,adm9,adm8,adm7,adm6,adm5,adm4,adm3,adm2,adm1,Falko,Xogenesis,f1,f2,f3,Kohri:2009yn,Agashe:2004bm,Farrar:2004qy,Khlopov,Wise}. In one case, a primordial asymmetry in one sector is transferred to the other sector. Here the primordial DM ($\eta_{DM}$) and baryonic ($\eta_B$) asymmetries generally satisfy either $\eta_{DM}\sim\eta_B$ or $\eta_{DM}\ll\eta_B$ depending on whether the transfer mechanism decouples when the DM is relativistic or non-relativisti
, respectively. Whereas if both asymmetries are generated by the same physical process then $\eta_{DM}\gg\eta_B$ is also possible. Now, if, in analogy with the visible sector, the present-day DM population is totally asymmetric, one can explain the coincidence expressed in~(\ref{Omegas}) by \emph{tuning} the DM mass to the value
\begin{eqnarray}\label{mDM}
m_{ADM}=\frac{\Omega_{DM}}{\Omega_B}\frac{\eta_B}{\eta_{DM}}m_p,
\end{eqnarray}
with $m_p$ denoting the proton mass.
The final DM abundance is rendered asymmetric by removing the conventional symmetric component that arises from thermal freeze-out. In the existing literature, this is typically achieved by the introduction of either a strong coupling, in analogy with QCD for the visible sector, or new light states. Both ingredients effectively induce an annihilation cross section sufficiently large to suppress the thermal symmetric component. Our results alleviate this hurdle to ADM model building and bolsters the case for their further study.
Here we consider the coupled evolution of the symmetric and asymmetric populations via the Boltzmann equations and quantify how large the annihilation cross section needs to be in order for the present DM densities to be asymmetric. We will see that the symmetric population depends \emph{exponentially} on the annihilation cross section, and hence that a weak scale force typically suffices to remove the symmetric component. In this sense, the weakly interacting massive particle (WIMP) thermal freeze-out scenario is not incompatible with the paradigm of ADM.
We then discuss the implications for ADM and point out that the spectrum of possible ADM scenarios is much richer than previously thought. In particular, we claim that most weakly coupled extensions of the standard model are likely to interpolate between the extreme cases of purely asymmetric DM and symmetric DM.
In this {\it asymmetric WIMP} scenario the present-day DM abundance is determined by a combination of its thermal annihilation cross section, its mass, and the primordial asymmetry. This relation among the short-distance and primordial parameters generalizes what is required in the extremely asymmetric or symmetric scenarios. In fact, for an asymmetric WIMP the total dark matter abundance is allowed to depend on a new observable, namely the ratio of the anti-dark matter to dark matter number densities. In contrast, in the asymmetric scenarios considered so far this ratio is fixed to be zero, and consequently the present abundance is set by fixing the mass as in~(\ref{mDM}). Likewise, in the symmetric scenario the dark matter and anti-dark matter abundances are assumed to be exactly equal, and the present abundance is then set by fixing the annihilation cross section. For asymmetric WIMP dark matter, these two scenarios are recovered as limiting cases of either small or large present-day anti-DM particle abundances.
We then apply our results to some existing models of ADM transfer operators \cite{ADM}, and investigate more generally the utility of weak-scale suppressed higher dimensional operators. This is motivated by our findings that
an asymmetric WIMP needs an annihilation cross-section only a few times larger than a picobarn to obtain the correct dark matter abundance. As an illustration, for the Higgs portal we find that requiring the correct annihilation cross-section leads to spin-independent direct detection rates at currently observable levels. The next round of direct detection experiments can cover a substantial part of the parameter space.
The outline of the paper is as follows. In Sec.~\ref{origin} we solve the Boltzmann equations for a generic species in the presence of a primordial particle/anti-particle asymmetry. In Sec.~\ref{adm} we specialize our results to the case in which the species is the DM, and discuss the implications of the asymmetric WIMP. In Sec.~\ref{adm:app} we apply our results to existing scenarios of ADM and determine the limits imposed by direct detection experiments. We summarize our results in Sec.~\ref{conc}.
In Appendix~\ref{CT} we present accurate results for $2\rightarrow n$ collision terms in the approximation that the incoming particles are non-relativistic on average and the final state threshold is much larger than the masses of the incoming particles. In Appendix~\ref{colapps} we apply the results of the previous Appendix to the specific transfer operators considered in the text. Finally in Appendix~\ref{appDD} we compute the elastic scattering cross section for direct detection arising from the Higgs portal operator.
\section{On the origin of asymmetric species}
\label{origin}
We begin with an analysis of the effect of a primordial particle-antiparticle asymmetry on a generic species $X$ of mass $m$. Previous work on the relic abundance in the presence of an asymmetry appears in \cite{GS,ST}.
We assume that the particle $X$ is \emph{not} self-conjugate~\footnote{This is certainly the case if $X$ carries a $U(1)_X$ global number, but our results apply more generally.}, and that a particle/anti-particle asymmetry in the $X$-number is generated at high temperatures. As the Universe expands, the number violating effects decouple at a temperature $T_{D}$ and the asymmetry is frozen in for $T< T_{D}$. At this stage the number density of particles ($n^+$) and antiparticles ($n^-$) is controlled by a set of coupled Boltzmann equations. Under reasonable assumptions~\footnote{In writing~(\ref{Boltz}) we assumed that i) our particle species is in a bath of particles in thermal equilibrium, ii) that no mass degeneracy between the two sectors is present, iii) that the dominant process changing the $n^\pm$ densities is annihilation, and iv) that the annihilation process occurs far from a resonant threshold~\cite{GS1}.} these reduce to:
\begin{eqnarray}\label{Boltz}
\frac{dn^\pm}{dt}+3Hn^\pm=-\langle\sigma_{\textrm{ann}}v\rangle\left(n^+n^--n_{eq}^+n_{eq}^-\right),
\end{eqnarray}
where $\langle\sigma_{\textrm{\small{ann}}}v\rangle\equiv\sigma_0 (T/m)^n$ is the thermally averaged annihilation cross section times the relative particle velocity. In the absence of significant entropy production it is convenient to introduce the quantity $Y^\pm\equiv n^\pm/s$, where $s$ denotes the total entropy density. In terms of this quantity, the particle-antiparticle asymmetry can be expressed by
\begin{eqnarray}
\eta\equiv Y^+-Y^-.
\end{eqnarray}
Without loss of generality, we conventionally define \emph{particles} to be more abundant than \emph{anti-particles} such that $\eta \ge 0$. Notice that, as anticipated, $\eta$ stays constant in the thermal evolution described by~(\ref{Boltz}).
The present densities $Y^\pm_\infty$ turn out to be strong functions of the primordial asymmetry $\eta$ \emph{and} the annihilation cross section $\sigma_0$. We distinguish between two limiting scenarios:
\begin{itemize}
\item In the first regime the dynamics can be considered ``strong" and the final abundance is dominated by particles over antiparticles. In this case $Y^-_\infty\ll Y^+_\infty\simeq\eta$. This is typical of most baryogenesis schemes if $X$ is identified with the ordinary baryons.
\item In the second regime the dynamics is ``weak" and the final abundances of particles and antiparticles are comparable, and separately much bigger than the asymmetry. In this case $Y^-_\infty\simeq Y^+_\infty\gg\eta$. This is generally assumed in the standard thermal WIMP scenario if $X$ is identified with the dark matter.
\end{itemize}
In the remainder of this section we will present a careful study of the Boltzmann equations~(\ref{Boltz}) in the presence of a primordial asymmetry, and provide a quantitative assessment of what ``strong" and ``weak" dynamics mean (see Sec.~\ref{analyticsol}). We will see that there is a broad range in parameter space between the two extremes discussed above in which the final abundances are comparable, namely
\begin{eqnarray}\label{intermediate}
Y^-_\infty\sim Y^+_\infty\sim\eta.
\end{eqnarray}
\subsection{The fractional asymmetry}
In order to solve the coupled system~(\ref{Boltz}) it is useful to introduce the quantity
\begin{eqnarray}
r=\frac{n^-}{n^+}=\frac{Y^-}{Y^+}.
\end{eqnarray}
Because the asymmetry $\eta$ is conserved, knowledge of the above quantity
suffices to determine the relative abundances at any time via~\footnote{As anticipated, the following discussion does not depend on our convention that particles are more abundant that anti-particles. In particular, the physics is invariant under the transformation $\eta\rightarrow-\eta$, $r\rightarrow1/r$.}
\begin{eqnarray}\label{YY}
Y^+=\eta\,\frac{1}{1-r}\quad\quad Y^-=\eta\,\frac{r}{1-r}.
\end{eqnarray}
Notice that by definition $r$ satisfies $0\leq r\leq1$.
We refer to $r$ as the \emph{fractional asymmetry} because it gives a measure of the effective particle-antiparticle asymmetry of a given species at any time. This should be compared to the primordial asymmetry $\eta$, which provides such information \emph{only} in the ultra-relativistic limit. The functional dependence of $r$ on the primordial asymmetry $\eta$ and the strength of the particle interactions, $\langle\sigma_{\textrm{ann}} v\rangle$, will be given below. The parameter $r$ is in principle an observable quantity as it controls the expected indirect signal coming from particle annihilation. Small $r$ corresponds to an extremely asymmetric case in which indirect annihilation signals are small or even absent, whereas in the large $r$ regime such signals are unsuppressed.
We assume that the Universe is radiation dominated in the epoch of interest, in which case the Hubble parameter is
\begin{eqnarray}\label{H}
H(T)=\left(\frac{8 \pi^{3}}{90}\right)^{1/2} g_{\textrm{\small{eff}}}^{1/2}(T)\frac{T^2}{M_{Pl}}=\frac{1}{2t},
\end{eqnarray}
with $M_{Pl}\approx 1.22\times10^{19}$ GeV.
If the annihilation cross section for the species $X$ is not too small compared to a typical weak scale process, the dynamical effects encoded in~(\ref{Boltz}) become relevant when the particle is non-relativistic. In this case the equilibrium distributions can be taken to be
\begin{eqnarray}
n_{eq}^\pm=n_{eq}e^{\pm\xi}\quad\quad\quad n_{eq}=g\left(\frac{mT}{2\pi}\right)^{3/2}e^{-m/T}.
\end{eqnarray}
In the above expression $g$ counts the internal degrees of freedom and $\mu=\xi T$ is the chemical potential. The explicit expression of the entropy density per comoving volume
\begin{eqnarray}
s(T)=\frac{2\pi^2}{45}h_{\textrm{\small{eff}}}(T)T^3,
\end{eqnarray}
allows us to introduce the convenient notation
\begin{eqnarray}
Y_{eq}\equiv\frac{n_{eq}}{s}=ax^{3/2}e^{-x},~~~~~ x=\frac{m}{T},
\end{eqnarray}
with $a\equiv 45g/(4 \sqrt{2} \pi^{7/2}h_{\textrm{\small{eff}}}) $.
With these definitions the dynamical equation for $r(x)$ following from~(\ref{Boltz}) reads
\begin{eqnarray}\label{Boltz'}
\frac{dr}{dx}&=&-\lambda\eta g_*^{1/2}\, x^{-n-2}\left[r-\frac{Y_{eq}^2}{\eta^2}(1-r)^2\right] \nonumber \\
&=& -\lambda\eta g_*^{1/2}\, x^{-n-2} \left[ r - r_{eq}\left( \frac{1-r}{1-r_{eq}}\right)^{2} \right], \end{eqnarray}
where
\begin{eqnarray}
\label{lambda}
\lambda&=&\left(\frac{\pi}{45}\right)^{1/2}M_{Pl}m\sigma_0,
\end{eqnarray}
and~\footnote{By requiring that $sR^3$ stays constant in time, taking into account the temperature dependence of $h_{\textrm{\small{eff}}}$ and using the relation $t=t(T)$ given by the definition of $H$, one finds that $d\log h_{\textrm{\small{eff}}}=\frac{3}{4}d\log g_\textrm{\small{eff}}$. With the help of this result one verifies that~(\ref{gstar}) coincides with the expression used in~\cite{darksusy}.}
\begin{eqnarray}
\label{gstar}
g_*^{1/2}=\frac{h_\textrm{\small{eff}}}{g_\textrm{\small{eff}}^{1/2}}\left(1-\frac{1}{4}\frac{x}{g_\textrm{\small{eff}}}\frac{dg_\textrm{\small{eff}}}{dx}\right).
\end{eqnarray}
In the above we have also defined $r_{eq} \equiv e^{-2\xi(x)}$, where $\xi$ is determined by
\begin{eqnarray}\label{xi}
2\sinh\xi=\frac{\eta}{Y_{eq}}.
\end{eqnarray}
Notice from (\ref{gstar}) that we have taken into account the temperature dependence in $h_\textrm{\small{eff}}$ and $g_\textrm{\small{eff}}$. Because $g_\textrm{\small{eff}}$ is monotonically increasing with $T$, we see that the parentheses in the definition of $g^{1/2}_*$ is positive definite. In the numerical results that follow we use the data table from DarkSUSY~\cite{darksusy} for the temperature dependence of $g_{*}(T)$ and $h_{\textrm{\small{eff}}}(T)$.~\footnote{Note that we use the notation of DarkSUSY for the massless degrees of freedom parameters. To translate our notation to that of Kolb and Turner~\cite{KT} one should make the substitutions $g_\textrm{\small{eff}} \rightarrow g_{*}$ and $h_\textrm{\small{eff}} \rightarrow g_{*S}$.}
Eq.~(\ref{Boltz'}) reproduces the well known case $\eta=0$ for which one finds that $r=1$ for any $x$. We will instead focus on scenarios with $\eta\neq0$ in the following. As shown in Fig.~\ref{fig0}, the effect of nonzero $\eta$ is to deplete the less abundant species more efficiently compared to $\eta = 0$ for the same annihilation cross section and mass.
\begin{figure
\begin{center}
\includegraphics[width=4.5in]{yx2.pdf}
\caption{\small Evolution of $Y^{\pm}(x)$ illustrating the effect of the asymmetry $\eta$. After freeze-out both $Y^-$ and $Y^+$ continue to evolve as the anti-particles find the particles and annihilate. The $Y^\pm_{\eta =0}$ curve shows the abundance for $\eta=0$, a mass $m=10$ GeV and annihilation cross-section $\sigma_0=2$ pb. In contrast, with a non-zero asymmetry $\eta=\eta_B = 0.88 \times 10^{-10}$ and same mass and cross-section, the more abundant species (here $Y^+$) is depleted less than when $\eta=0$. Also shown is the equilibrium solution $Y_{eq}(x)$.
\label{fig0}}
\end{center}
\end{figure}
\subsection{The relic abundance of asymmetric species}
\label{analyticsol}
Equation~(\ref{Boltz'}) can be solved by numerical methods and imposing an appropriate initial condition at a scale $x=x_i\geq10$, where the non-relativistic approximation works very well. Although we have chosen $x_i=10$, we have checked that larger values ($10<x_i<x_f$, where $x_f$ is defined below) do not alter the final result. From (\ref{Boltz'}) one sees that in the early Universe $r = r_{eq}$, provided the cross-section is not too small. In our numerical solutions $r(x_i)$ is chosen to equal its equilibrium value $r(x_i)\equiv r_{eq}(x_i)=e^{-2\xi(x_i)}$.
We now present a very accurate analytic approximation of the numerical results. The numerical, exact solution will be compared to it shortly.
While there are no analytic solutions of~(\ref{Boltz'}), it is not difficult to guess the qualitative behavior of $r(x)$. As already mentioned, for small $x$ the solution is very close to the equilibrium solution $r_{eq}(x)$.
As the Universe expands, $r$ eventually has difficulty tracking the equilibrium expression which is exponentially decreasing. The easiest way to see this is to inspect the first equation in~(\ref{Boltz'}) and observe that the second term on the right-side is proportional to $Y_{eq} \sim e^{-x}$. The deviation of $r$ from the equilibrium solution begins to grow.
The \textit{freeze-out} scale $x\equiv x_f$ is then defined by the condition that the two terms on the right hand side of~(\ref{Boltz'}) no longer balance each other, and $d r /dx \approx d r_{eq}/dx $ becomes comparable to the terms on the right-side of the Boltzmann equation.
The behavior for $x>x_f$ is then dominated by the first term on the right hand side of~(\ref{Boltz'}) whatever the asymmetry is because of the exponential dependance of $Y_{eq}$ on $x$. We can therefore write
\begin{eqnarray}\label{rx}
r(x)=r_{eq,f}\, e^{-\lambda\eta\,\Phi(x,m)}
\label{Exponentialr}
\end{eqnarray}
with
\begin{eqnarray}\label{Phi}
\Phi(x,m)&\equiv& \int_{x_f}^x dx'\, x'^{-n-2}g_*^{1/2}\,\\\nonumber
&=& \int_{m/x}^{m/x_f} \frac{dT }{m}\left(\frac{T}{m}\right)^ng_*^{1/2} \\\nonumber
\end{eqnarray}
and $r_{eq,f} \equiv r_{eq}(x_f) $.
The present value of the fractional asymmetry is thus controlled by the quantities $r_{eq,f}$, $x_f$, $m$, and ultimately by $\xi_{f}$. Importantly, observe from (\ref{Exponentialr}) that the fractional asymmetry is \emph{exponentially} sensitive to the annihilation cross-section.
It remains to find an estimate for the freeze-out scale. An analytic expression for $x_f$ is obtained by the condition that $d r /dx \approx d r_{eq}/dx$ is comparable to either the first or second term in eq.~(\ref{Boltz'}).
This leads to
\begin{eqnarray}
\frac{d r_{eq}}{d x } \approx -\delta\, \lambda \eta g^{1/2}_* x^{-n-2} r_{eq},
\end{eqnarray}
or equivalently using the approximate relation~(\ref{xi}),
\begin{eqnarray}\label{FO}
\left(1-\frac{3}{2x_f}\right)\tanh\xi_f=\left(1-\frac{3}{2x_f}\right)\frac{1-r_{eq,f}}{1+r_{eq,f}}\approx\delta\frac{\lambda\eta g_{*,f}^{1/2}}{2 x_f^{n+2}},
\end{eqnarray}
with $\delta$ a number to be determined by fitting the numerical solution. Because $x_f$ turns out to be $\gtrsim20$ for the scenarios of interest (see the end of this section), we can ignore the $3/2x_f$ factor in the above expression.
As long as $g_{*,f}^{1/2}\lambda\eta<2x_f^{n+2}$, the condition~(\ref{FO}) can be solved iteratively. After one iteration, and with the help of equation~(\ref{xi}), we find
\begin{eqnarray}\label{xf}
x_f\approx \log(\delta g_{*,f}^{1/2}a_f\lambda)+\frac{1}{2}\log\frac{\log^3(\delta g_{*,f}^{1/2}a_f\lambda)}{\log^{2n+4}(\delta g_{*,f}^{1/2}a_f\lambda)-g_{*,f}\left(\delta \frac{\lambda\eta}{2}\right)^2}+\dots.
\end{eqnarray}
This expression is accurate at the percent level, so we will use it in our applications. Notice that the effect of a nonzero asymmetry $\eta$ on the freeze-out temperature is very mild (see for example Fig.~\ref{fig0}). In fact, for all $g_{*,f}^{1/2}\lambda\eta<2x_f^{n+2}$ the scale $x_f$ is very close to the result found in the symmetric limit, in which case $\delta$ is typically taken to be $\delta=n+1$~\cite{ST}. As already emphasized, however, the effect of $\eta$ on the anti-particle population is considerable.
The asymmetry can only be considered small if $g_{*,f}^{1/2}\lambda\eta\ll2x_f^{n+2}$, in which case one has $Y^+_\infty\simeq Y^-_\infty\gg\eta$, as anticipated in the introduction. More generally, in the ``weak" coupling regime $g_{*,f}^{1/2}\lambda\eta<2x_f^{n+2}$ we find that to a high accuracy $r_{eq,f}\simeq1$. The final estimate for the present fractional asymmetry in this limit thus reads
\begin{eqnarray}\label{estimate1}
r_\infty= e^{-\lambda\eta\,\Phi}\quad\quad \left[\textrm{if}\quad\frac{\lambda\eta g_{*,f}^{1/2}}{2 x_f^{n+2}}<
\right],
\end{eqnarray}
where $\Phi$ is understood to be $\Phi=\Phi(\infty,m)$. This result agrees with~\cite{GS}.
While it is clear that Eq.~(\ref{estimate1}) describes the regime where $r_{\infty} \lesssim 1$, it is less clear how far this approximation extends to the ``strong" coupling regime in which $r_{\infty} \ll1$. In order to appreciate the range of validity of~(\ref{estimate1}), we estimate the integral~(\ref{Phi}) assuming that $g_*^{1/2}$ is approximately constant in the relevant range of temperatures. With this approximation one finds that $\Phi\approx g_{*,f}^{1/2}/[(n+1)x_f^{n+1}]$, and~(\ref{estimate1}) becomes
\begin{eqnarray}\label{gg}
r_\infty\simeq\exp\left(-\frac{\lambda\eta g_{*,f}^{1/2}}{(n+1)x_f^{n+1}}\right)=\exp\left(-\frac{\lambda\eta g_{*,f}^{1/2}}{2x_f^{n+2}}\,\frac{2x_f}{(n+1)}\right).
\end{eqnarray}
To see how far the approximation extends, first note that for $r_{\infty} \gtrsim O(10^{-4})$, the exponent in~(\ref{gg}) should be at most $O(10)$. Since for a weak-scale cross section $x_f\sim20$, the upper bound on the exponent translates into an upper bound of $\lambda\eta g_{*,f}^{1/2}/[2x_f^{n+2}]<O(0.2)$, in agreement with our assumption~(\ref{estimate1}). We thus conclude that~(\ref{estimate1}) not only incorporates the intermediate limit anticipated in~(\ref{intermediate}), but also the ``strong" limit in which $r_{\infty} \gtrsim 10^{-4}$.
As $2x_f^{n+2}\rightarrow \lambda\eta g_{*,f}^{1/2}$ the fractional asymmetry gets further suppressed since now $r_{eq,f}$ can be much smaller than 1. Because in this regime $r_\infty$ is essentially vanishing for all practical purposes, we will be mainly concerned with the regime~(\ref{estimate1}) in the following sections.
Finally, the estimate~(\ref{estimate1}) turns out to be a very good approximation of the exact (numerical) result. We scanned a large range of parameter space ranging from $\eta=O(10^{\pm4})\eta_B$, $m=O(10^{\pm4})m_p$, and $\sigma_0$ several orders of magnitude above and below the ``thermal WIMP" value $\sigma_{0,WIMP}=O(1)$ pb for both s-wave and p-wave processes. Based on this analysis, we have found that our analytic expression departs by at most $5\%$ from the numerical result. Consistent with the derivation presented above, within the scanned range we have $\lambda\eta g_{*,f}^{1/2}/[2x_f^{n+2}]<O(0.1)$.
\section{Asymmetric WIMP dark matter}
\label{adm}
In asymmetric dark matter (ADM) scenarios the dark matter is assumed to have a primordial asymmetry $\eta$. For definiteness we will measure the DM asymmetry in units of the baryon asymmetry via the relation
\begin{eqnarray}\label{eta}
\eta=\epsilon\eta_B
\end{eqnarray}
where $\eta_B = (0.88 \pm 0.021)\times10^{-10}$~\cite{WMAPetab} \footnote{$\eta$ and $\eta_B$ are normalized relative to the total entropy density (i.e., neutrinos $+$ photons).}. The present-day abundance of baryons is $\rho_{B} = m_{p}~s~\eta_{B}$, whereas the DM abundance is
\begin{eqnarray} \label{presconst} \rho_{DM} &=& m~s \left(Y^{+} + Y^{-} \right) \nonumber \\
&=& m ~s \left( Y^{+} - Y^{-} + 2 Y^{-} \right) \nonumber \\
&=& m~s \left( \eta + 2 \frac{\eta ~r_{\infty} }{1- r_{\infty}}\right), \end{eqnarray}
where we have used~(\ref{YY}) to express the present day density of anti-DM, $Y^{-}$, in terms of $\eta$ and $r_\infty$. The first term of (\ref{presconst}) corresponds to the asymmetric component which survives in the limit $r_{\infty} \rightarrow 0$, where the abundance scales as $\rho_{DM} \sim \eta $. Whereas the second term corresponds to the symmetric component, in which case the abundance scales as $\rho_{DM} \sim 1/ \langle \sigma_{ann} v \rangle$ when $r_{\infty} \rightarrow 1$. In the intermediate regime, one may expect the total abundance to scale as simply the sum of these two approximate scalings. However, this is not the case in general since the symmetric component depends exponentially on $\eta \, \langle \sigma_{ann} v \rangle$.
We write the observational constraint~(\ref{Omegas}) as
\begin{eqnarray}\label{const}
\frac{m}{m_p}\epsilon = \left(\frac{1-r_{\infty}}{1+r_{\infty}}\right)\frac{\Omega_{DM}}{\Omega_B},
\end{eqnarray}
with $m_p$ and $m$ the proton and DM masses, respectively, and $r_\infty\equiv Y^-_\infty/Y^+_\infty$ the present fractional asymmetry. The present-day constraints (\ref{presconst}) and $\rho_{B} = m_{p}~s~\eta_{B}$ are equivalent to satisfying (\ref{const}), generating the correct baryon asymmetry $\eta_B$, and having a theory for (\ref{eta}).
The literature so far has focused on the following two extreme scenarios. In the first scenario (ADM) the DM is totally asymmetric, namely $r_\infty=0$. The observed relic abundance is then explained if the DM mass is \emph{tuned} to the value
\begin{eqnarray}\label{ADMm}
\textrm{ADM:}\quad m_{ADM}= \frac{\Omega_{DM}}{\Omega_B}\frac{m_p}{\epsilon},
\end{eqnarray}
and there is no strong constraint on the cross section. The cross section simply needs to be large enough (see below) to suppress $r_\infty$.
In the second extreme scenario (thermal WIMP) the DM is totally symmetric, namely $r_\infty=1$. From our analytic expression~(\ref{estimate1}) we find that in this limit the relation~(\ref{const}) becomes
\begin{eqnarray}\label{WIMP}
\textrm {Thermal WIMP:}\quad\quad\sigma_{0,WIMP}=\frac{1}{\eta_Bm_p}\frac{\Omega_B}{\Omega_{DM}}\frac{1}{M_{Pl}\Phi_{WIMP}}\sqrt{\frac{180}{\pi}}.
\end{eqnarray}
If $g_*$ is assumed to be constant, $\Phi=g_{*,f}^{1/2}/[(n+1)x_f^{n+1}]$, and the above formula becomes the one commonly used in the literature. The DM abundance is now determined by \emph{tuning} the annihilation cross section, the dependence on $m$ being very mild.
For the intermediate scenarios in which $0<r_\infty<1$ the physics interpolates between the above extremes. In Fig.~\ref{fig1} we plot the allowed parameter space in the $m-\sigma_0$ plane of a DM candidate satisfying the condition~(\ref{const}) for different choices of the asymmetry parameter $\epsilon$ (see Eq.~(\ref{eta})). The two figures refer to purely s-wave and p-wave processes, respectively. For a given asymmetry $\epsilon$, we see that the allowed region is bounded from below by the WIMP cross section $\sigma_{0,WIMP}$~(\ref{WIMP}), and bounded from the right as a maximum on the mass at the ADM mass value~(\ref{ADMm}).
For $\epsilon\rightarrow0$ and $m\gtrsim20$ GeV the lines asymptote to a horizontal line representing the thermal WIMP scenario, where the cross section is fixed but the mass can vary. The regime $m\lesssim20$ GeV is instead strongly sensitive to the abrupt change in the number of degrees of freedom around the QCD phase transition. The latter appears as a large gradient in $g_*^{1/2}$ at $T^*\lesssim1$ GeV~\cite{darksusy}. The effect on the relic abundance of the species is relevant if $T_f=m/x_f\lesssim T^*$, and can be seen in Fig.~\ref{fig1} as the bump at $m\lesssim x_fT^*\simeq20$ GeV. This change in the degrees of freedom also tends to reduce the exponent in~(\ref{rx}) for light species.
In Fig.~\ref{fig1} the fractional asymmetry varies from $r_\infty=0$ in the upper part of the curves to $r_\infty=1$ when the curves overlap with the thermal WIMP scenario ($\epsilon=0$). The same information is presented differently in Fig. \ref{fig2b}, where we trade $\epsilon$ for $r_{\infty}$.
In this figure the exponential sensitivity of $r_{\infty}$ to the annihilation cross-section is evident.
\begin{figure}
\begin{center}
\includegraphics[width=3.5in]{SigmaMn0.pdf}\\
\vspace{5mm}
\includegraphics[width=3.5in]{SigmaMn1.pdf}
\caption{\small Here we plot the annihilation cross section $\sigma_0$ required to reproduce the correct DM abundance $\Omega_{DM}$ via a s-wave process $n=0$ (above plot) and p-wave $n=1$ (bottom plot) for a given dark matter mass $m$, and for various values of the primordial asymmetry $\eta=\epsilon\eta_B$. The line for $\epsilon=0$ corresponds to the usual thermal WIMP scenario. Notice that the fractional asymmetry runs from $r_\infty=0$ in the upper part of the curves to $r_\infty=1$ when the lines converge on the standard thermal WIMP curve. The effect of the QCD phase transition appears as a bump at $m\lesssim20$ GeV, as anticipated in the text. Note that the bottom plot is basically enhanced by a factor $\Phi_{n=0}/\Phi_{n=1}\sim (n+1)x_f$ compared to the former. As a reference, recall that $1$ pb $\simeq2.6\times10^{-9}$ GeV$^{-2}$.
\label{fig1}}
\end{center}
\end{figure}
A generic model for beyond the SM physics is expected to lie in between the two extreme scenarios described above. In fact as emphasized in the introduction, a primordial DM asymmetry ($\epsilon\neq0$) is generally expected; meanwhile most models beyond the SM naturally lead to WIMP-like cross sections. The resulting \emph{asymmetric WIMP} scenario will typically have $\sigma_0\gtrsim\sigma_{0,WIMP}$ and $r_\infty\lesssim1$.
It is instructive to re-express the fractional asymmetry for an asymmetric DM species as a function of its thermally averaged cross section $\sigma_0$ and the quantity $\Phi=\Phi(\infty,m)$ defined in~(\ref{Phi}):
\begin{eqnarray}\label{rinf}
r_{\infty}= e^{-\lambda\eta\Phi}=\exp\left[-2\left(\frac{\sigma_0}{\sigma_{0,WIMP}}\right) \left(\frac{\Phi}{\Phi_{WIMP}}\right) \frac{1-r_{\infty}}{1+r_{\infty}}\right],
\end{eqnarray}
where $\sigma_{0,WIMP}$ and $\Phi_{WIMP}\approx\Phi$ refer to a symmetric DM candidate of the same mass. The first equality in~(\ref{rinf}) is a good analytic approximation to the numerical solution (see~(\ref{estimate1})), whereas the second is exact and follows from eliminating $\eta$, $\lambda$ and $m$ using (\ref{lambda}) and the empirical formul\ae~(\ref{const}) and~(\ref{WIMP}).
From (\ref{rinf}) we see explicitly that for WIMP-like cross-sections and larger, the dependence of $r_{\infty}$ on the cross section is very sensitive.
For a more quantitative evaluation, we show in Fig.~\ref{fig3} the solution of the above implicit equation. Indeed, we see that an $O(1)$ departure from the thermal WIMP cross section implies a significant change in $r_\infty$.
Indirect signals of annihilating asymmetric DM in the Universe are suppressed compared to those of a symmetric thermal candidate of the same mass by the quantity
\begin{eqnarray}\label{sup}
\frac{\sigma_0}{\sigma_{0,WIMP}}\times r_\infty\times \left(\frac{2}{1+r_\infty}\right)^2 \leq1.
\end{eqnarray}
One can understand the above formula as follows. The rate for indirect events is proportional to $\langle\sigma_{\textrm{\small{ann}}}v\rangle r_\infty(n^+_{\eta\neq0})^2$ for an asymmetric candidate, or similarly to $\langle\sigma_{\textrm{\small{ann}}}v\rangle (n^+_{\eta=0})^2$ for a symmetric (thermal WIMP) candidate. Notice, however, that the number densities $n^+_{\eta\neq0}$ and $n^+_{\eta=0}$ are not the same. In fact, for DM candidates of the same mass it is the total (particle plus antiparticle) densities that must be equal, namely $n_{\eta\neq0}^+(1+r_\infty)=2n^+_{\eta=0}$. Taking the ratio of these rates gives~(\ref{sup}).
We plot the quantity~(\ref{sup}) as a function of the cross section as a dashed line in Fig.~\ref{fig3}. Interestingly, we find that potentially important \emph{indirect} signals are still effective in the intermediate regime $\sigma_0\gtrsim\sigma_{0,WIMP}$ in which $r_\infty\lesssim1$. For example, a suppression of $\,\lesssim0.5$ is found for values of the fractional asymmetry below $\sim0.1$ (i.e. for $\sigma_0/\sigma_{0,WIMP}\gtrsim1.5$), when the DM could already be considered asymmetric.
\begin{figure
\begin{center}
\includegraphics[width=4.5in]{sigrinf.pdf}
\caption{\small Same as in Figure \ref{fig1} showing contours of constant $r_{\infty}$, for s--wave process. The completely asymmetric scenario $r_{\infty} \ll 1$ corresponds to the top region of the plot having cross-sections always larger than the thermal WIMP cross-section. For reference we show the $\epsilon=0$ line which corresponds to the usual thermal WIMP scenario ($r_{\infty}=0$). \label{fig2b}}
\end{center}
\end{figure}
\begin{figure
\begin{center}
\includegraphics[width=4.5in]{rinf.pdf}
\caption{\small Present fractional asymmetry $r_\infty=Y^-_\infty/Y^+_\infty$ (see (\ref{rinf})) for various values of the thermally averaged cross section $\sigma_0$ (defined as $\langle\sigma_{\textrm{\small{ann}}}v\rangle\equiv\sigma_0 (T/m)^n$ in the first section) in units of the value $\sigma_{0,WIMP}$ obtained for an exactly symmetric species of the same mass, see eq.(\ref{WIMP}). All the points in the curve, i.e. all solutions of~(\ref{rinf}), account for the observed DM density. The dashed line is the suppression factor~(\ref{sup}) for indirect detection. \label{fig3}}
\end{center}
\end{figure}
\section{Asymmetric WIMP dark matter: applications}\label{adm:app}
In the following subsections we apply our results to a few ADM models, choosing our examples from among those already existing in the literature~\cite{ADM,Xogenesis}. We demonstrate there is a large allowed region in these models in which the present-day fractional dark matter asymmetry is in the range $0<r_\infty<1$ and, {\em significantly}, the dark matter annihilation cross-section in the early Universe can be comparable to the typical WIMP cross-section. Our results generalize the conclusions found in the previous literature, which only discusses completely asymmetric dark matter, corresponding to a present fractional asymmetry of $r_\infty=0$ in our notation.
In the examples we discuss the MSSM is assumed to couple to dark matter superfields $X,\bar X$ via a heavy messenger sector. The dark matter chiral superfields $X, \overline{X}$ carry $\pm 1$ charge under a hidden sector global $U(1)$, are Standard Model gauge singlets, and have a SUSY invariant mass $m_X$. For definiteness we assume that after SUSY breaking the fermionic component $\psi_X$ is the dark matter. Similar results hold for scalar DM candidates.
The overall system is assumed to violate either lepton ($L$) number or baryon ($B$) number together with the dark ($X$) number, but to (perturbatively) preserve a linear combination of these symmetries. When the number-violating interactions are in thermal equilibrium, chemical equilibrium forces a relation between the asymmetries of the visible and dark sectors and therefore determine the quantity $\epsilon$ introduced earlier~(\ref{eta}). We will refer to this effect as a transfer mechanism.
\subsection{Example 1: Lepton-number violating transfer operator}
\label{example1}
In the first example we consider lepton number violation and take the leading transfer mechanism between the dark and leptonic sectors, after having integrated out the messengers, to be described by a nonrenormalizable term in the superpotential~\cite{ADM}:
\begin{eqnarray}\label{transf3}
\Delta W_{asym}= \frac{XXH_uL}{\Lambda},
\end{eqnarray}
with $\Lambda$ taken to be close to the TeV scale. The operator~(\ref{transf3}) preserves a global $L'=X-2L$ number, but violates both $L$ and $X$. The quantum numbers $L,X$ become good quantum numbers in the early Universe as soon as the transfer operator~(\ref{transf3}) freezes out, assumed to occur for simplicity at an instantaneous temperature $T_D$. One can show that for $\Lambda=O(1)$ TeV $T_D$ is below the scale $T_{sph}$ at which the sphalerons shut off, in which case the only source of lepton number violation at low temperatures is induced by~(\ref{transf3}).
This model has either one or two stable particles, depending on the relative mass difference between the $\psi_X$'s and the lightest superpartner of the SM. To see that note that the model has a $Z_4$ symmetry (see also~\cite{ADM}) which is the usual $Z_2$ $R-$parity in the visible sector, plus a $Z_4$ in the $X$ sector that acts as $+i$ on the fermionic component $\psi_X$ and $-i$ on the scalar component $\widetilde{X}$ (the $\overline{X}$ fields have the opposite charge). This symmetry remains unbroken by soft SUSY and electroweak symmetry breaking, and guarantees that at least one component of $X$ is always stable. The other SUSY component of $X$ generically decays through the transfer operator to the lightest component of $X$ and SM fields.
The transfer operator typically causes the lightest superpartner (LSP) of the SM to be unstable, if the channel is kinematically allowed. (The decay $\hbox{LSP} \rightarrow XX + SM ~particles$ is consistent with the $Z_4$ symmetry).
If $m_{LSP} < 2 m_X$ then this decay is forbidden and the model has two stable particles. In what follows we will assume $m_{LSP} > 2 m_X$ for simplicity.
\subsubsection{Chemical Potential Analysis}
To estimate the primordial DM asymmetry parameter $\epsilon$, see~(\ref{eta}), we proceed in two steps. For a useful review on chemical potential analysis see~\cite{HT}.
As a first step we impose chemical equilibrium and charge neutrality at the scale $T=T_{sph}>T_D$ at which the sphalerons decouple, assumed to be below the Higgs condensate scale $T_{c}$. At these temperatures the only dynamical particles are the DM and the Standard Model fields (including the top quark), and we have:
\begin{eqnarray}\label{chem}
Q&\propto&18\mu_u-12\mu_d-6\mu_e=0\\\nonumber
\mu_u-\mu_d&=&\mu_\nu-\mu_e\\\nonumber
\mu_u+2\mu_d+\mu_\nu&=&0\\\nonumber
2\mu_X+\mu_\nu&=&0,
\end{eqnarray}
where the chemical potentials $\mu_u, \mu_d, \mu_\nu, \mu_e,$ and $\mu_X$ refer to the Standard Model up, down quarks, the (purely left handed) neutrinos, the charged leptons, and the DM respectively. The first equation in~(\ref{chem}) imposes electrical neutrality, the second accounts for the $W^\pm$ exchange, the third for the sphaleron process, and finally the last equation follows from the transfer operator~(\ref{transf3}). The solution of the system~(\ref{chem}) can be used to derive the primordial baryon and $L'$ asymmetries:
\begin{eqnarray}\label{BL'}
\eta_B&= &-\frac{36}{7}\mu_e\\\nonumber
\eta_{L'}&=&\left(\frac{25}{6}+\frac{11}{36} \frac{f(m/T_{sph})}{f(0)}\right)\eta_B,
\end{eqnarray}
where the function $f(x)$ is defined by
\begin{eqnarray}\label{boltz}
f(x)=\frac{1}{4\pi^2}\int_0^\infty dy \frac{y^2}{\cosh^2\left(\frac{1}{2}\sqrt{x^2+y^2}\right)},
\end{eqnarray}
and accounts for the possibility that the DM be non-relativistic at the relevant temperature~\cite{Barr:1990ca}. If the DM is instead bosonic the $\cosh(x)$ function should be replaced by $\sinh(x)$.
Below the sphaleron temperature the asymmetries for $B$ and $L'$ given above are conserved, while $L$ and $X$ are still violated by~(\ref{transf3}). In the second step of our chemical potential analysis we thus compute the present asymmetry for the $X$ number at the temperature $T_D$ at which the transfer operator decouples.
Chemical equilibrium at $T_D$ imposes a number of conditions similar to~(\ref{chem}). The two main differences are that at this scale the sphaleron process (see third line in~(\ref{chem})) is no longer effective, and that the top quark is not included. The system of equations for the chemical potentials is now solved by replacing the sphaleron process constraint with the initial conditions provided by~(\ref{BL'}), and not including the top quark contribution. The result is finally
\begin{eqnarray}\label{epsilonHuLXX}
\epsilon\equiv\frac{\eta_X}{\eta_B}=\left(\frac{309+22 \frac{f(m/T_{sph})}{f(0)}}{1026+72 \frac{f(m/T_D)}{f(0)}}\right)\frac{f(m/T_D)}{f(0)}.
\end{eqnarray}
Our result~(\ref{epsilonHuLXX}) reduces to the one found in~\cite{ADM} if the top quark is instead decoupled above the sphaleron scale, and if one assumes $f(m/T_D)=f(m/T_{sph})=f(0)$. We also checked that the result changes only mildly if one assumes that the massive vector bosons are decoupled at the scale $T_D$. In all these cases one finds that~(\ref{epsilonHuLXX}) can be well approximated as $\epsilon\approx0.3 f(m/T_D)/f(0)$.
\subsubsection{Transfer Operator Decoupling}
In order for the analysis of the previous sections to be applicable, the freeze-out temperature $T_f$ must be below the scale $T_D$ at which the transfer operator decouples~\footnote{We will comment on the limiting case $T_D=T_f$ in Section~\ref{discuss}. For the moment, notice that the case $T_D<T_f$ is clearly meaningless: freeze-out always occurs when the strongest interaction decouples.}: only under this condition there exists a temperature range $T_f\leq T\leq T_D$ in which Eq.~(\ref{Boltz}), and the subsequent analysis, is fully reliable. We will see in subsection~\ref{anndd} that, in typical UV completions of the model~(\ref{transf3}), this condition is ensured by higher dimensional operators that preserve the dark number and $L$, and which are therefore not involved in transferring asymmetries.
The Boltzmann equation governing the evolution of the $X$ number at $T \gtrsim T_{D}$ is given by Eq.~(\ref{Boltz}), but with additional, $X$-number changing terms induced by~(\ref{transf3}). These latter involve the DM and the SM particles as final states, with typical reactions of the form
\begin{eqnarray}\label{XXnu'}
\psi_X \psi _X \leftrightarrow \overline{\psi}_X \overline{\psi}_X \overline{\nu} \overline{ \nu}.
\end{eqnarray}
The decoupling temperature $T_D$ is then estimated as the scale at which the strongest interaction mediated by the transfer operator -- formally described by a ``collision term" $C$ on the right hand side of~(\ref{Boltz}) -- becomes comparable to the Hubble expansion rate, namely when
\begin{eqnarray}\label{TD}
C=\,H(T_D)n_X,
\end{eqnarray}
with $n_X$ the DM equilibrium number density.
If the transfer operator decouples when the DM is relativistic, the Boltzmann suppression in~(\ref{epsilonHuLXX}) is not effective and the parameter $\epsilon$ does not depend on the DM mass. In this case the details of the transfer mechanism are irrelevant.
The Boltzmann suppression in~(\ref{epsilonHuLXX}) is effective if the transfer operator decouples when the DM is non-relativistic. In this case only the precise value of $T_D$, and hence the details of the transfer operator matter. In the specific case in which the neutralino is the LSP, we find that the dominant number-violating interaction mediated by~(\ref{transf3}) in the non-relativistic limit is associated to the process~(\ref{XXnu'}), and occurs via the production of an on-shell intermediate neutralino which eventually decays with a branching ratio BR $=1/2$ into an $X$- and $L$-number violating final state $\psi_X \psi _X \rightarrow \tilde{\chi} \overline{\nu}\rightarrow\overline{\psi}_X \overline{\psi}_X \overline{\nu} \overline{ \nu}$.
If we neglect the DM mass, then at leading order in $T/m_{ \tilde{\chi}}$ the collision term $C$ reduces to (see Appendix \ref{lhu})
\begin{eqnarray}\label{C}
C=-\frac{\sqrt{2\pi}}{4(2\pi)^5}\left(\frac{g_w v_u}{m_{\tilde\nu}^2\Lambda}\right)^2\, m_{ \tilde{\chi}}^8\left(\frac{T}{m_{ \tilde{\chi}}}\right)^{9/2}\,e^{-m_{ \tilde{\chi}}/T}.
\end{eqnarray}
In the above expression, $g_w$ is a weak gauge coupling, $v_u=v\sin\beta$ is the vacuum expectation value of the up-type Higgs, and $m_{\tilde\nu}$ is the sneutrino mass. The exponential suppression in~(\ref{C}) reflects the fact that the process occurs far in the Boltzmann tail of the thermal distribution of the incoming DM particles.
The factors of $T$ in (\ref{C}) arise from a partial integration in the process of taking the thermal average, and reflect the fact that the phase space integral vanishes when the energy of the DM particles is at threshold. Another way to understand these factors of $T$ is to note that for the inverse process $\nu \tilde{\chi} \rightarrow \psi_{X} \psi_{X}$ the powers of $T$ arise solely from the number densities of the incoming particles simply because there is no threshold and the matrix element is set by the mass of the LSP. The inverse collision term is therefore proportional to $n_{\tilde{\chi}} n_{\nu}\sim T^{3/2} T^{3} e^{-m_{\tilde{\chi}}/T}$.
The decoupling temperature is finally estimated using~(\ref{TD}), with $C$ given by~(\ref{C}) and $n_X$ the non-relativistic DM equilibrium density. We find that, up to negligible logarithmic corrections, $T_D$ is given by
\begin{eqnarray}
T_D\approx \frac{m_{ \tilde{\chi}}-m_X}{\log\gamma},~~\quad\gamma=\frac{1}{4(2\pi)^3}\left(\frac{g_wv_u}{m_{\tilde\nu}^2\Lambda}\right)^2\frac{(m_{ \tilde{\chi}}-m_X) M_{Pl}m_{ \tilde{\chi}}^{7/2}}{\left(\frac{8 \pi^{3}}{90}\right)^{1/2} g_{\textrm{\small{eff,D}}}^{1/2}\,m_X^{3/2}}
\end{eqnarray}
which approximately reads $T_D\sim(m_{ \tilde{\chi}}-m_X)/20$ if the mass scales are of $O($GeV-TeV$)$. We now have all the ingredients required to estimate the primordial asymmetry parameter $\epsilon$, see eq.~(\ref{epsilonHuLXX}), as a function of the DM mass and the LSP mass $m_{ \tilde{\chi}}$. This estimate has been found here working under the assumption that the transfer operator decouples when the DM is already non-relativistic. Yet, because the Boltzmann suppression in~(\ref{epsilonHuLXX}) ceases to be effective when the non-relativistic limit is still a good approximation, it is clear that our estimate of $\epsilon$ should also apply to the case in which the transfer operator decouples when the DM is relativistic. As emphasized above, indeed, in this latter case the details of the transfer mechanism do not matter.
\subsubsection{\label{discuss}Discussion}
The main purpose of the present subsection is to scan the parameter space of the model~(\ref{transf3}) and show that there is a sizeable region in the allowed parameter space with a present-day fractional asymmetry in the range $0.1<r_\infty<1$. This is the asymmetric WIMP regime.
The relevant parameters of the model~(\ref{transf3}) at low temperatures reduce to the decoupling temperature $T_D$, the DM mass $m_X$, and the annihilation cross section (encoded in $r_\infty$). The requirement that the particle $\psi_X$ saturates the observed DM density is implemented by~(\ref{const}) and it amounts to a constraint on these parameters. In Fig.~\ref{figHuLXX} we show the resulting contour plot for the quantity $r_\infty$ \emph{defined} by~(\ref{const}) as a function of $m_X$ and $m_{ \tilde{\chi}}$.
\begin{figure
\begin{center}
\includegraphics[width=4.5in]{nicerplot.pdf}
\caption{\small Contour plot for $r_\infty$ in the allowed parameter space for the model~(\ref{transf3}). Below the lower line and above the upper line the DM density is smaller and higher than experimentally observed, respectively. In the dark blue area $0.9\leq r_\infty<1$, in the blue area $0.1\leq r_\infty\leq0.9$, and in the light blue area $0\leq r_\infty\leq0.1$. See the text for more details. Here we have taken $m_{\tilde{\nu}} = \Lambda = 1$ TeV and $\tan \beta = 20$. \label{figHuLXX}}
\end{center}
\end{figure}
The upper curve shows the line~(\ref{ADMm}) at which $r_\infty=0$. In the area above this curve the DM abundance exceeds the observed value. This region of the parameter space is therefore excluded. The lowest curve is determined by the condition $T_D= T_f\sim m_X/20$, which approximately reads $m_{ \tilde{\chi}}\sim2 m_X$ in our case, at which $r_\infty\simeq1$ (and $\epsilon\ll1$). In the area below this latter curve the relic abundance of the species $\psi_X$ is smaller than the observed value. This area is excluded here because $\psi_X$ is assumed for simplicity to be the only DM candidate. The intermediate lines in Fig.~\ref{figHuLXX} separate regions with different present-day fractional asymmetry $r_\infty$.
The plot can be understood as follows. For any $r_\infty$, a given value for the DM mass completely determines $\epsilon$ with~(\ref{const}). This fixes $T_D$ and in turn it determines a function $m_{ \tilde{\chi}}(m_X)$. As long as the Boltzmann suppression in~(\ref{epsilonHuLXX}) is active the curve is approximately a straight line determined by $m_X/T_D\sim const\gtrsim1$. This behavior is visible in the large DM mass regime of Fig.~\ref{figHuLXX}. The Boltzmann suppression shuts off when the previous lines hit the boundary of the non-relativistic regime $m_X/T_D\sim1$. For DM masses below this point the primordial asymmetry parameter is given by $\epsilon\sim0.3$ and $m_X$ is fully determined by $r_\infty$. This effect is seen as a vertical asymptote in the low DM mass regime of Fig.~\ref{figHuLXX}.
It is evident from Figure~\ref{figHuLXX} that a large region in parameter space is associated to the asymmetric WIMP scenario $0.1\lesssim r_\infty\lesssim1$. Here indirect signals are unsuppressed.
\subsubsection{Annihilation and direct detection}
\label{anndd}
In all the cases discussed above -- either the strongly asymmetric $r_\infty\ll1$ or strongly symmetric $r_\infty\simeq1$ limit -- we see from Fig.~\ref{fig3} that the annihilation cross section can be of thermal WIMP magnitude. Such a relatively low cross section may be induced by annihilation into SM particles mediated by non-renormalizable operators suppressed by the very same scale $\Lambda$ suppressing~(\ref{transf3}). From an effective field theory perspective in fact we expect that a generic UV completion of~(\ref{transf3}) also leads to direct couplings between the DM and the SM particles. For example, the following corrections to the superpotential and the K\"{a}hler are allowed by all the symmetries
\begin{eqnarray}\label{Wother}
\Delta W_{sym}&=&{\lambda}^2\frac{X\overline{X}H_uH_d}{\Lambda}+\dots\\
\Delta K_{sym} &=& y^2\frac{X^\dagger XL^\dagger L}{\Lambda^2}+\dots, \label{Kother}
\end{eqnarray}
and are therefore generic
We now show that both operators~(\ref{Wother}) and~(\ref{Kother}) provide annihilation cross sections of the right strength if $\Lambda=O( \hbox{TeV})$. These operators can therefore be responsible for keeping the DM in thermal equilibrium below $T_D$, and in particular for accounting for the values of $r_\infty$ required to generate the correct relic abundance. As we will see, they can also provide direct detection signals. Let us now discuss a few annihilation modes in more detail.
The correction to the K\"{a}hler triggers the process $\psi_X\overline{\psi_X}\rightarrow l\bar l$, where $l$ are SM leptons and $\psi_X$ is a Dirac DM fermion. In order to avoid large flavor violating effects that might arise from~(\ref{Kother}), we follow~\cite{ADM} and assume that the only relevant coupling is with the third SM lepton generation. Neglecting the $\tau$ and $\nu_\tau$ masses we find~\cite{ADM}:
\begin{eqnarray}\label{annl}
\langle\sigma_{\textrm{ann}}v\rangle=\frac{y^4}{16\pi}\frac{m_X^2}{\Lambda^4}.
\end{eqnarray}
It is easy to see that this channel can provide an annihilation cross section of the correct order for several DM masses. For example, for a light DM candidate, say with $m_X\sim20$ GeV one should require $\Lambda/y\lesssim200$ GeV, whereas for weak-scale DM masses, say $m_X\sim200$ GeV one sees that $\Lambda/y\lesssim700$ GeV suffices. If~(\ref{Wother}) provides the dominant annihilation mode, one finds no constraining direct detection bounds~\cite{ADM}.
The dominant annihilation mode mediated by the operator $X\overline{X}H_uH_d$, on the other hand, occurs via the s-wave annihilation either to final state fermions -- through the exchange of the (virtual) CP odd scalar $A^0$ -- or to a $h A^0$ final state. For the first channel the largest rate occurs if the final state fermions are top quarks, if kinematically allowed. Specifically, for $\overline{\psi_X}\psi_X\rightarrow {A^0}^{*}\rightarrow\bar t t$ and $ m_X > m_t$, we have
\begin{eqnarray}\label{XXtt}
\langle\sigma_{\textrm{ann}}v\rangle=\frac{\lambda^4}{\tan^2\beta}\frac{N_c}{8\pi}\frac{m_t^2}{\Lambda^2}\frac{m_X^2}{(4m_X^2-m_{A^0}^2)^2} \sqrt{1-\frac{m_t^2}{m_X^2}}.
\end{eqnarray}
This can be of the correct magnitude to account for the DM abundance if $A^0$ is not too heavy, $m_X$ is not too small, and $\tan\beta$ not too large.
For example, if we take $m_X=200$ GeV $>m_{A^0}/2$ and $\tan^2\beta\approx1$
the averaged cross section is close to the WIMP value; for these values $\langle\sigma_{\textrm{ann}}v\rangle\approx\lambda^4\times4\times 10^{-9}$ GeV$^{-2} (1.2 \hbox{ TeV}/\Lambda)^2$. If this channel is kinematically closed the other channel may be open and provide a large enough rate. For $2m_X > m_{h} + m_{A^0}$, we have for the s-wave annihilation $\overline{X}X \rightarrow h A^0$,
\begin{eqnarray}\label{XXAh}
\langle\sigma_{\textrm{ann}}v\rangle=\lambda^4 \frac{\sin^2 (\beta -\alpha) }{64 \pi \Lambda^2}
\sqrt{1-\frac{(m_{A^0}+m_{h})^2}{4 m^2_X} } \sqrt{1-\frac{(m_{A^0}-m_{h})^2}{4 m^2_X} } .
\end{eqnarray}
Unlike the previous cross-section, this one is not suppressed at large $\tan \beta$.
If we take $\sin^2 (\beta -\alpha) \approx 1$, the averaged cross-section is
$\langle\sigma_{\textrm{ann}}v\rangle\approx\lambda^4 \times4\times 10^{-9}$ GeV$^{-2} (1.5 \hbox{ TeV}/\Lambda)^2$.
If~(\ref{Wother}) provides the dominant annihilation mode, the DM direct detection bounds are not generally negligible. In the absence of light degrees of freedom other than the MSSM particles and the DM, the ``single nucleon'' cross section for spin-independent elastic DM scattering mediated by virtual exchanges of CP-even Higgses is
\begin{equation}
\label{sigxn}
\sigma_{Xn} = \left(\frac{1~{\rm GeV}}{\mu_{T}}\right)^{2} \frac{\sigma_{N}}{A^{2}},
\end{equation}
where the spin-independent dark matter-nucleus cross section is
\begin{equation} \sigma_{N} =\frac{\mu_{T}^{2}}{\pi} \left(Z f_{p} + (A-Z) f_{n}\right)^{2} \end{equation}
with $f_{p}$ and $f_{n}$ controlling the coupling between nucleons and the dark matter, and $\mu_{T}$ the DM-nucleus reduced mass. The nucleon coupling parameters $f_{p}$ and $f_{n}$ can be found in Appendix~\ref{appDD} and depend on the scale in Eq.~(\ref{Wother}) as well as parameters in the Higgs sector.
Requiring that the annihilation cross section is larger than the WIMP value implies a lower bound on DM-nucleon cross section. Similarly, the bounds from direct detection can be translated into an upper bound for the annihilation cross section and thus a lower bound for $r_\infty$.
\begin{figure
\begin{center}
\mbox{\subfigure{\includegraphics[width=2.8in]{ddbounds700.pdf}}~~~
\subfigure{ \includegraphics[width=2.8in]{ddbounds1000.pdf}}}
\caption{\small Here we show the impact of the CDMS-II bound~\cite{cdms} $\sigma_{Xn} = 5\times 10^{-44}~\rm{cm}^{2}$ and the new XENON100 bounds~\cite{xenon} $\sigma_{Xn} = 3.3\times 10^{-44}~\rm{cm}^{2}$ on the parameter space of our model for $m_{X}=400$ GeV and $\Lambda = 700$ GeV (left plot) and $\Lambda = 1000$ GeV (right plot). To be consistent with the null results of CDMS-II, a parameter point must lie to the right of the red curve, while to be consistent with XENON100 bounds a point must lie to the right of the dashed blue curve. Obtaining the correct relic abundance requires $\langle\sigma_{\textrm{ann}}v\rangle\gtrsim \sigma_{0,WIMP}$ which occurs for points to the left of the solid black curve. For reference we include the dashed black curve corresponding to $\langle\sigma_{\textrm{ann}}v\rangle= 2 ~\sigma_{0,WIMP}$, which is not present in the $\Lambda = 1000$ GeV plot. Note that the $\Lambda = 1000$ GeV case is strongly constrained by direct detection bounds. \label{figdd}}
\end{center}
\end{figure}
To see that, let us assume that the dominant annihilation mode is $\overline{X}X \rightarrow h A^0$ and that $m_X$ is at the weak scale. We also know from Fig.~\ref{fig1} that $\langle\sigma_{\textrm{ann}}v\rangle\gtrsim \sigma_{0,WIMP} = 4.5\times10^{-9}$ GeV$^{-2} $. The CDMS-II bound along with this constraint on the annihilation cross section is shown in Fig.~\ref{figdd}. We also include the new XENON100 bound $\sigma_{Xn} = 3.3\times 10^{-43}~\rm{cm}^{2}$~\cite{xenon}. In Fig.~\ref{figdd} we have taken a dark matter mass $m_{X} = 400$ GeV, a transfer operator scale $\Lambda = 700$ GeV (left plot) and $\Lambda = 1000$ GeV (right plot) and a light Higgs mass $m_{h} = 120$ GeV. We have also assumed $m_{A^{0}}^{2} \gg m_{Z}^{2}$ such that the mixing angle $\alpha$ is determined at tree-level by $\alpha \approx \beta - \pi/2$ and the heavy Higgs mass is $ m_{H}^{2} \approx m_{A^{0}}^{2}$. The relation between $\alpha$ and $\beta$ implies that the annihilation cross section contours appear as straight vertical lines. The interested reader can find additional details in Appendix~\ref{appDD}.
Our analysis of the annihilation modes has been model-independent so far. We end this section by emphasizing that, once a UV completion for the operator~(\ref{transf3}) is known, alternative reactions, and constraints, might arise. As an instance, consider the UV completion proposed in~\cite{ADM}. One introduces two electro-weak doublets $D, \bar D$ with hypercharge $\mp\frac{1}{2}$ and generalized lepton number $L'=\mp1$, and two SM singlets $X, \bar X$ with $L'=\pm1$. One readily sees that integrating out the $D$ field one gets the operator~(\ref{transf3}) as well as~(\ref{Wother}) and~(\ref{Kother}). A typical feature of this UV completion is the occurrence of a mixing between the neutral component of the doublet $D$ and the DM after electro-weak symmetry breaking. This mixing induces a coupling of the DM $\psi_X$ to the $Z^0$ of order $\sim g\lambda_{u,d}v_{u,d}/\Lambda$, and thus implies new annihilation and direct detection modes mediated by a virtual gauge boson. Given the lower bound on the annihilation cross section, one finds that the direct detection signal is always too large if the mixing-induced $Z^0$ exchange is the dominant channel. One should then suppress this channel by taking $\lambda_{u,d}<O(10^{-1})$. At this point the dominant annihilation mode would be given by~(\ref{annl}). Alternatively one could consider the NMSSM, where a large contribution to~(\ref{Wother}) might be obtained integrating out the singlet $S$, also coupled to the DM via $\lambda'S\bar X X$. In this latter case the coupling of~(\ref{Wother}) would be naturally at the TeV scale, and freeze-out would thus be controlled by either $\overline{\psi_X}\psi_X\rightarrow {A^0}^{*}\rightarrow\bar t t$ or $\overline{X}X \rightarrow h A^0$. Notice that in all of the cases discussed above the DM is assumed to be the lightest state beyond the MSSM.
\subsection{Example 2: Baryon-number violating transfer operator}
We now consider an example in which the transfer operator violates $B$ and $X$, but not $L$. The model is again an extension of the MSSM, and it is described by the following nonrenormalizable term~\cite{ADM}:
\begin{eqnarray}\label{op}
\Delta W_{asym}=\frac{XXu^c d^c d^c}{\Lambda^2}.
\end{eqnarray}
The superpotential operator~(\ref{op}) preserves a global $B'=X+2B$ number, but violates $B$ and $X$ separately. The physics of this model is very similar to the lepton violating operator~(\ref{transf3}), see for example the subsection~\ref{example1}.
A chemical potential analysis similar to the one presented for the $L$-violating operator in this case gives:
\begin{eqnarray}\label{epsilonXXqqq}
\epsilon\equiv\frac{\eta_X}{\eta_B}=\left(\frac{1756 + 253 \frac{f(m/T_{sph})}{f(0)}}{5472+ 836 \frac{f(m/T_{sph})}{f(0)}- 50 \frac{f(m/T_D)}{f(0)}}\right)\frac{f(m/T_D)}{f(0)}.
\end{eqnarray}
Similarly to~(\ref{epsilonHuLXX}), this result is very robust and approximately reads $\epsilon\approx0.3 f(m/T_D)/f(0)$.
The typical processes transferring the asymmetry between the visible and the dark sectors
can occur through on-shell or off-shell neutralinos or squarks. The processes involving neutralinos are always strongly suppressed by phase space, and we find that on-shell production of squarks dominates for typical masses; see Appendix~\ref{appudd} for details. Using again our approximate expression~(\ref{Capp}) the collision term for the process $\psi_X \psi_X \rightarrow\tilde{u}\tilde{d}\tilde{d}$ is found to be
\begin{eqnarray}
C=-\frac{9\sqrt{2}\pi^{3/2}}{32\Lambda^4}\frac{N_c!}{(2\pi)^7}\,m_{\tilde q}^8\left(\frac{T}{m_{\tilde q}}\right)^{9/2}\,e^{-3m_{ \tilde{q}}/T}\left[1+O\left(\frac{T}{m_{ \tilde{q}}},\frac{m_X}{m_{\tilde{q}}}\right)\right].
\end{eqnarray}
where the powers of $T$ have the same origin as the ones in~(\ref{C}), and we assumed that the squarks are degenerate for simplicity.
The analysis can now proceed similarly to the lepton-number violating transfer operator. The result can again be summarized with a plot analogous to Fig.~\ref{figHuLXX}, which for brevity we do not show. The comments in Section~\ref{example1} regarding s--wave annihilation and direct detection via the superpotential operator $ X \overline{X} H_u H_d$ also apply in this case.
\section{Conclusions}
\label{conc}
In this paper we studied the implication of a nontrivial DM primordial asymmetry $\eta$ on the DM relic abundance. For any $\eta$, the physical quantity controlling the effective asymmetry of the species is the fractional asymmetry $r(x)=Y^-/Y^+$. By solving the coupled Boltzmann equations for $Y^\pm$ we found that $r(x)$ depends \emph{exponentially} on both the annihilation cross section and $\eta$, see~(\ref{estimate1})~\cite{GS}. The present anti-particle population of an asymmetric DM species is therefore extremely sensitive to the strength of the dynamics.
The parameter space in a DM model with a primordial asymmetry $\eta$ is defined as follows. The DM mass $m$ is bounded from above by the mass $m_{ADM}$ of a totally asymmetric, $r_\infty=0$, DM species, see~(\ref{ADMm}):
\begin{eqnarray}
m\leq m_{ADM}\equiv\frac{\Omega_{DM}}{\Omega_B}\frac{\eta_B}{\eta}m_p.
\end{eqnarray}
\emph{For any $\eta$} one finds $m< m_{ADM}$ as soon as one departs from the condition $r_\infty=0$. In particular, very low masses $m\ll m_p$ are always allowed provided $r_\infty\rightarrow1$.
The thermally averaged annihilation cross section for an asymmetric DM candidate is instead bounded from below by the value found for a totally symmetric species of the same mass (thermal WIMP value), see~(\ref{WIMP}):
\begin{eqnarray}
\langle\sigma_{\textrm{\small{ann}}}v\rangle\geq\langle\sigma_{\textrm{\small{ann}}}v\rangle_{WIMP}.
\end{eqnarray}
\emph{For any $\eta$} the cross section approaches its lowest value when $r_\infty\rightarrow1$, while a larger cross section can still be accommodated in the regime $r_\infty<1$. Notice that because of the exponential dependence of $r_\infty$ on the cross section, an effectively asymmetric DM population, say with $r_\infty\lesssim10^{-(2-3)}$, is already obtained with an annihilation cross section of just a factor $\sim(2-3)$ bigger than the thermal WIMP cross section. This latter point has interesting consequences for model-building, as it implies that new light states, often introduced in previous studies of asymmetric DM scenarios to induce large cross section
, are not necessary.
A generic weakly coupled UV completion for the standard model will typically have weak scale annihilation cross sections. If the DM is stabilized by a symmetry distinguishing particles from anti-particles it is also likely that the DM has a nonzero primordial asymmetry $\eta$. As a result, asymmetric WIMP scenarios are expected to be generic. These scenarios typically lie between the extremely asymmetric and symmetric limits. Surprisingly, the indirect signals from present-day annihilation can be much larger than what one would naively expect from an asymmetric scenario.
To illustrate the generality of our conclusions we applied our results to two existing models of asymmetric DM~\cite{ADM}. We saw that the parameter space consistent with the present-day DM abundance and characterizing the asymmetric WIMP regime is sizable. This property stems from the fact that the present-day fractional asymmetry $r_\infty$ acts as new, low energy parameter.
Moreover direct detection bounds impose interesting limits on the Higgs portal operator $\overline{X}XH_{u}H_{d}$. Although this operator was introduced to yield a sufficiently large annihilation cross section, the same interaction gives the dominant interaction with nuclei. We found that the current bounds from CDMS-II and XENON100 constrain the model, but upcoming direct detection experiments will be sensitive to a wide range of the parameters relevant for the asymmetric WIMP through the Higgs portal.
\acknowledgments
We would like to thank Matthew McCullough for pointing out a minor error in our chemical potential analysis in the first version of this work. This work has been supported by the U.S. Department of Energy at Los
Alamos National Laboratory under Contract No. DE-AC52-06NA25396. The preprint number for this manuscript is LA-UR 11-00565.
\section{Introduction}
Andrei Sakharov~\cite{Sak} pointed out that a primordial particle number asymmetry is produced given three ingredients: symmetry violation, C and CP violation, as well as a departure from thermal equilibrium. The observed baryon asymmetry suggests that these conditions have been met in the history of the Universe. Since global symmetry violation is expected to occur in the early Universe, we expect the existence of primordial asymmetries to be rather generic.
The dominant contributions to the observed matter density in the Universe are due to dark matter and baryons. Their density fractions are remarkably well determined from cosmic microwave background data from WMAP7~\cite{WMAP7}:
\begin{eqnarray}\label{Omegas}
\Omega_{DM} h^{2} =0.1109 \pm 0.0056\quad\quad\quad\Omega_{B}h^{2} =0.02258^{+0.00057}_{-0.00056}.
\end{eqnarray}
Yet, very little remains known about the nature of dark matter (DM). In most theoretical efforts, the problems of dark matter and baryogenesis are treated separately, with the implicit assumption that the comparable densities of these two types of matter is simply a coincidence. However, the fact that the abundances of baryons and DM are just a factor $\sim5$ apart may be an indication of a common underlying origin. The paradigm of asymmetric dark matter (ADM) has been suggested as a way of linking the asymmetries, and thus the abundances in the dark and visible sectors; see~\cite{ADM} and references therein~\footnote{There are also many other models seeking to link the abundances of baryons and DM~\cite{Dodelson,Kaplan,Thomas,enqvist,Kitano,GravQ,Hylogenesis,Allahverdi,mcd,gu}.}.
In ADM two distinct scenarios are typically considered. In one case, a primordial asymmetry in one sector is transferred to the other sector~\cite{Barr:1990ca,Darkogenesis, adm16,adm15,adm14,adm13,adm12,adm11,adm10,adm9,adm8,adm7,adm6,adm5,adm4,adm3,adm2,adm1,Falko,Xogenesis}. Here the primordial DM ($\eta_{DM}$) and baryonic ($\eta_B$) asymmetries generally satisfy either $\eta_{DM}\sim\eta_B$ or $\eta_{DM}\ll\eta_B$ depending on whether the transfer mechanism decouples when the DM is relativistic or non-relativistic~\cite{Barr:1990ca,adm9,adm15,Xogenesis}, respectively. Whereas if both asymmetries are generated by the same physical process then $\eta_{DM}\gg\eta_B$ is also possible. Now, if, in analogy with the visible sector, the present-day DM population is totally asymmetric, one can explain the coincidence expressed in~(\ref{Omegas}) by \emph{tuning} the DM mass to the value
\begin{eqnarray}\label{mDM}
m_{ADM}=\frac{\Omega_{DM}}{\Omega_B}\frac{\eta_B}{\eta_{DM}}m_p,
\end{eqnarray}
with $m_p$ denoting the proton mass.
The final DM abundance is rendered asymmetric by removing the conventional symmetric component that arises from thermal freeze-out. In the existing literature, this is typically achieved by the introduction of either a strong coupling, in analogy with QCD for the visible sector, or new light states. Both ingredients effectively induce an annihilation cross section sufficiently large to suppress the thermal symmetric component. Our results alleviate this hurdle to ADM model building and bolsters the case for their further study.
Here we consider the coupled evolution of the symmetric and asymmetric populations via the Boltzmann equations and quantify how large the annihilation cross section needs to be in order for the present DM densities to be asymmetric. We will see that the symmetric population depends \emph{exponentially} on the annihilation cross section, and hence that a weak scale force typically suffices to remove the symmetric component. In this sense, the weakly interacting massive particle (WIMP) thermal freeze-out scenario is not incompatible with the paradigm of ADM.
We then discuss the implications for ADM and point out that the spectrum of possible ADM scenarios is much richer than previously thought. In particular, we claim that most weakly coupled extensions of the standard model are likely to interpolate between the extreme cases of purely asymmetric DM and symmetric DM.
In this {\it asymmetric WIMP} scenario the present-day DM abundance is determined by a combination of its thermal annihilation cross section, its mass, and the primordial asymmetry. This relation among the short-distance and primordial parameters generalizes what is required in the extremely asymmetric or symmetric scenarios. In fact, for an asymmetric WIMP the total dark matter abundance is allowed to depend on a new observable, namely the ratio of the anti-dark matter to dark matter number densities. In contrast, in the asymmetric scenarios considered so far this ratio is fixed to be zero, and consequently the present abundance is set by fixing the mass as in~(\ref{mDM}). Likewise, in the symmetric scenario the dark matter and anti-dark matter abundances are assumed to be exactly equal, and the present abundance is then set by fixing the annihilation cross section. For asymmetric WIMP dark matter, these two scenarios are recovered as limiting cases of either small or large present-day anti-DM particle abundances.
We then apply our results to some existing models of ADM transfer operators \cite{ADM}, and investigate more generally the utility of weak-scale suppressed higher dimensional operators. This is motivated by our findings that
an asymmetric WIMP needs an annihilation cross-section only a few times larger than a picobarn to obtain the correct dark matter abundance. As an illustration, for the Higgs portal we find that requiring the correct annihilation cross-section leads to spin-independent direct detection rates at currently observable levels. The next round of direct detection experiments can cover a substantial part of the parameter space.
The outline of the paper is as follows. In Sec.~\ref{origin} we solve the Boltzmann equations for a generic species in the presence of a primordial particle/anti-particle asymmetry. In Sec.~\ref{adm} we specialize our results to the case in which the species is the DM, and discuss the implications of the asymmetric WIMP. In Sec.~\ref{adm:app} we apply our results to existing scenarios of ADM and determine the limits imposed by direct detection experiments. We summarize our results in Sec.~\ref{conc}.
In Appendix~\ref{CT} we present accurate results for $2\rightarrow n$ collision terms in the approximation that the incoming particles are non-relativistic on average and the final state threshold is much larger than the masses of the incoming particles. In Appendix~\ref{colapps} we apply the results of the previous Appendix to the specific transfer operators considered in the text. Finally in Appendix~\ref{appDD} we compute the elastic scattering cross section for direct detection arising from the Higgs portal operator.
\section{On the origin of asymmetric species}
\label{origin}
We begin with an analysis of the effect of a primordial particle-antiparticle asymmetry on a generic species $X$ of mass $m$. Previous work on the relic abundance in the presence of an asymmetry appears in \cite{GS,ST}.
We assume that the particle $X$ is \emph{not} self-conjugate~\footnote{This is certainly the case if $X$ carries a $U(1)_X$ global number, but our results apply more generally.}, and that a particle/anti-particle asymmetry in the $X$-number is generated at high temperatures. As the Universe expands, the number violating effects decouple at a temperature $T_{D}$ and the asymmetry is frozen in for $T< T_{D}$. At this stage the number density of particles ($n^+$) and antiparticles ($n^-$) is controlled by a set of coupled Boltzmann equations. Under reasonable assumptions~\footnote{In writing~(\ref{Boltz}) we assumed that i) our particle species is in a bath of particles in thermal equilibrium, ii) that no mass degeneracy between the two sectors is present, iii) that the dominant process changing the $n^\pm$ densities is annihilation, and iv) that the annihilation process occurs far from a resonant threshold~\cite{GS1}.} these reduce to:
\begin{eqnarray}\label{Boltz}
\frac{dn^\pm}{dt}+3Hn^\pm=-\langle\sigma_{\textrm{ann}}v\rangle\left(n^+n^--n_{eq}^+n_{eq}^-\right),
\end{eqnarray}
where $\langle\sigma_{\textrm{\small{ann}}}v\rangle\equiv\sigma_0 (T/m)^n$ is the thermally averaged annihilation cross section times the relative particle velocity. In the absence of significant entropy production it is convenient to introduce the quantity $Y^\pm\equiv n^\pm/s$, where $s$ denotes the total entropy density. In terms of this quantity, the particle-antiparticle asymmetry can be expressed by
\begin{eqnarray}
\eta\equiv Y^+-Y^-.
\end{eqnarray}
Without loss of generality, we conventionally define \emph{particles} to be more abundant than \emph{anti-particles} such that $\eta \ge 0$. Notice that, as anticipated, $\eta$ stays constant in the thermal evolution described by~(\ref{Boltz}).
The present densities $Y^\pm_\infty$ turn out to be strong functions of the primordial asymmetry $\eta$ \emph{and} the annihilation cross section $\sigma_0$. We distinguish between two limiting scenarios:
\begin{itemize}
\item In the first regime the dynamics can be considered ``strong" and the final abundance is dominated by particles over antiparticles. In this case $Y^-_\infty\ll Y^+_\infty\simeq\eta$. This is typical of most baryogenesis schemes if $X$ is identified with the ordinary baryons.
\item In the second regime the dynamics is ``weak" and the final abundances of particles and antiparticles are comparable, and separately much bigger than the asymmetry. In this case $Y^-_\infty\simeq Y^+_\infty\gg\eta$. This is generally assumed in the standard thermal WIMP scenario if $X$ is identified with the dark matter.
\end{itemize}
In the remainder of this section we will present a careful study of the Boltzmann equations~(\ref{Boltz}) in the presence of a primordial asymmetry, and provide a quantitative assessment of what ``strong" and ``weak" dynamics mean (see Sec.~\ref{analyticsol}). We will see that there is a broad range in parameter space between the two extremes discussed above in which the final abundances are comparable, namely
\begin{eqnarray}\label{intermediate}
Y^-_\infty\sim Y^+_\infty\sim\eta.
\end{eqnarray}
\subsection{The fractional asymmetry}
In order to solve the coupled system~(\ref{Boltz}) it is useful to introduce the quantity
\begin{eqnarray}
r=\frac{n^-}{n^+}=\frac{Y^-}{Y^+}.
\end{eqnarray}
Because the asymmetry $\eta$ is conserved, knowledge of the above quantity
suffices to determine the relative abundances at any time via~\footnote{As anticipated, the following discussion does not depend on our convention that particles are more abundant that anti-particles. In particular, the physics is invariant under the transformation $\eta\rightarrow-\eta$, $r\rightarrow1/r$.}
\begin{eqnarray}\label{YY}
Y^+=\eta\,\frac{1}{1-r}\quad\quad Y^-=\eta\,\frac{r}{1-r}.
\end{eqnarray}
Notice that by definition $r$ satisfies $0\leq r\leq1$.
We refer to $r$ as the \emph{fractional asymmetry} because it gives a measure of the effective particle-antiparticle asymmetry of a given species at any time. This should be compared to the primordial asymmetry $\eta$, which provides such information \emph{only} in the ultra-relativistic limit. The functional dependence of $r$ on the primordial asymmetry $\eta$ and the strength of the particle interactions, $\langle\sigma_{\textrm{ann}} v\rangle$, will be given below. The parameter $r$ is in principle an observable quantity as it controls the expected indirect signal coming from particle annihilation. Small $r$ corresponds to an extremely asymmetric case in which indirect annihilation signals are small or even absent, whereas in the large $r$ regime such signals are unsuppressed.
We assume that the Universe is radiation dominated in the epoch of interest, in which case the Hubble parameter is
\begin{eqnarray}\label{H}
H(T)=\left(\frac{8 \pi^{3}}{90}\right)^{1/2} g_{\textrm{\small{eff}}}^{1/2}(T)\frac{T^2}{M_{Pl}}=\frac{1}{2t},
\end{eqnarray}
with $M_{Pl}\approx 1.22\times10^{19}$ GeV.
If the annihilation cross section for the species $X$ is not too small compared to a typical weak scale process, the dynamical effects encoded in~(\ref{Boltz}) become relevant when the particle is non-relativistic. In this case the equilibrium distributions can be taken to be
\begin{eqnarray}
n_{eq}^\pm=n_{eq}e^{\pm\xi}\quad\quad\quad n_{eq}=g\left(\frac{mT}{2\pi}\right)^{3/2}e^{-m/T}.
\end{eqnarray}
In the above expression $g$ counts the internal degrees of freedom and $\mu=\xi T$ is the chemical potential. The explicit expression of the entropy density per comoving volume
\begin{eqnarray}
s(T)=\frac{2\pi^2}{45}h_{\textrm{\small{eff}}}(T)T^3,
\end{eqnarray}
allows us to introduce the convenient notation
\begin{eqnarray}
Y_{eq}\equiv\frac{n_{eq}}{s}=ax^{3/2}e^{-x},~~~~~ x=\frac{m}{T},
\end{eqnarray}
with $a\equiv 45g/(4 \sqrt{2} \pi^{7/2}h_{\textrm{\small{eff}}}) $.
With these definitions the dynamical equation for $r(x)$ following from~(\ref{Boltz}) reads
\begin{eqnarray}\label{Boltz'}
\frac{dr}{dx}&=&-\lambda\eta g_*^{1/2}\, x^{-n-2}\left[r-\frac{Y_{eq}^2}{\eta^2}(1-r)^2\right] \nonumber \\
&=& -\lambda\eta g_*^{1/2}\, x^{-n-2} \left[ r - r_{eq}\left( \frac{1-r}{1-r_{eq}}\right)^{2} \right], \end{eqnarray}
where
\begin{eqnarray}
\label{lambda}
\lambda&=&\left(\frac{\pi}{45}\right)^{1/2}M_{Pl}m\sigma_0,
\end{eqnarray}
and~\footnote{By requiring that $sR^3$ stays constant in time, taking into account the temperature dependence of $h_{\textrm{\small{eff}}}$ and using the relation $t=t(T)$ given by the definition of $H$, one finds that $d\log h_{\textrm{\small{eff}}}=\frac{3}{4}d\log g_\textrm{\small{eff}}$. With the help of this result one verifies that~(\ref{gstar}) coincides with the expression used in~\cite{darksusy}.}
\begin{eqnarray}
\label{gstar}
g_*^{1/2}=\frac{h_\textrm{\small{eff}}}{g_\textrm{\small{eff}}^{1/2}}\left(1-\frac{1}{4}\frac{x}{g_\textrm{\small{eff}}}\frac{dg_\textrm{\small{eff}}}{dx}\right).
\end{eqnarray}
In the above we have also defined $r_{eq} \equiv e^{-2\xi(x)}$, where $\xi$ is determined by
\begin{eqnarray}\label{xi}
2\sinh\xi=\frac{\eta}{Y_{eq}}.
\end{eqnarray}
Notice from (\ref{gstar}) that we have taken into account the temperature dependence in $h_\textrm{\small{eff}}$ and $g_\textrm{\small{eff}}$. Because $g_\textrm{\small{eff}}$ is monotonically increasing with $T$, we see that the parentheses in the definition of $g^{1/2}_*$ is positive definite. In the numerical results that follow we use the data table from DarkSUSY~\cite{darksusy} for the temperature dependence of $g_{*}(T)$ and $h_{\textrm{\small{eff}}}(T)$.~\footnote{Note that we use the notation of DarkSUSY for the massless degrees of freedom parameters. To translate our notation to that of Kolb and Turner~\cite{KT} one should make the substitutions $g_\textrm{\small{eff}} \rightarrow g_{*}$ and $h_\textrm{\small{eff}} \rightarrow g_{*S}$.}
Eq.~(\ref{Boltz'}) reproduces the well known case $\eta=0$ for which one finds that $r=1$ for any $x$. We will instead focus on scenarios with $\eta\neq0$ in the following. As shown in Fig.~\ref{fig0}, the effect of nonzero $\eta$ is to deplete the less abundant species more efficiently compared to $\eta = 0$ for the same annihilation cross section and mass.
\begin{figure
\begin{center}
\includegraphics[width=4.5in]{yx2.pdf}
\caption{\small Evolution of $Y^{\pm}(x)$ illustrating the effect of the asymmetry $\eta$. After freeze-out both $Y^-$ and $Y^+$ continue to evolve as the anti-particles find the particles and annihilate. The $Y^\pm_{\eta =0}$ curve shows the abundance for $\eta=0$, a mass $m=10$ GeV and annihilation cross-section $\sigma_0=2$ pb. In contrast, with a non-zero asymmetry $\eta=\eta_B = 0.88 \times 10^{-10}$ and same mass and cross-section, the more abundant species (here $Y^+$) is depleted less than when $\eta=0$. Also shown is the equilibrium solution $Y_{eq}(x)$.
\label{fig0}}
\end{center}
\end{figure}
\subsection{The relic abundance of asymmetric species}
\label{analyticsol}
Equation~(\ref{Boltz'}) can be solved by numerical methods and imposing an appropriate initial condition at a scale $x=x_i\geq10$, where the non-relativistic approximation works very well. Although we have chosen $x_i=10$, we have checked that larger values ($10<x_i<x_f$, where $x_f$ is defined below) do not alter the final result. From (\ref{Boltz'}) one sees that in the early Universe $r = r_{eq}$, provided the cross-section is not too small. In our numerical solutions $r(x_i)$ is chosen to equal its equilibrium value $r(x_i)\equiv r_{eq}(x_i)=e^{-2\xi(x_i)}$.
We now present a very accurate analytic approximation of the numerical results. The numerical, exact solution will be compared to it shortly.
While there are no analytic solutions of~(\ref{Boltz'}), it is not difficult to guess the qualitative behavior of $r(x)$. As already mentioned, for small $x$ the solution is very close to the equilibrium solution $r_{eq}(x)$.
As the Universe expands, $r$ eventually has difficulty tracking the equilibrium expression which is exponentially decreasing. The easiest way to see this is to inspect the first equation in~(\ref{Boltz'}) and observe that the second term on the right-side is proportional to $Y_{eq} \sim e^{-x}$. The deviation of $r$ from the equilibrium solution begins to grow.
The \textit{freeze-out} scale $x\equiv x_f$ is then defined by the condition that the two terms on the right hand side of~(\ref{Boltz'}) no longer balance each other, and $d r /dx \approx d r_{eq}/dx $ becomes comparable to the terms on the right-side of the Boltzmann equation.
The behavior for $x>x_f$ is then dominated by the first term on the right hand side of~(\ref{Boltz'}) whatever the asymmetry is because of the exponential dependance of $Y_{eq}$ on $x$. We can therefore write
\begin{eqnarray}\label{rx}
r(x)=r_{eq,f}\, e^{-\lambda\eta\,\Phi(x,m)}
\label{Exponentialr}
\end{eqnarray}
with
\begin{eqnarray}\label{Phi}
\Phi(x,m)&\equiv& \int_{x_f}^x dx'\, x'^{-n-2}g_*^{1/2}\,\\\nonumber
&=& \int_{m/x}^{m/x_f} \frac{dT }{m}\left(\frac{T}{m}\right)^ng_*^{1/2} \\\nonumber
\end{eqnarray}
and $r_{eq,f} \equiv r_{eq}(x_f) $.
The present value of the fractional asymmetry is thus controlled by the quantities $r_{eq,f}$, $x_f$, $m$, and ultimately by $\xi_{f}$. Importantly, observe from (\ref{Exponentialr}) that the fractional asymmetry is \emph{exponentially} sensitive to the annihilation cross-section.
It remains to find an estimate for the freeze-out scale. An analytic expression for $x_f$ is obtained by the condition that $d r /dx \approx d r_{eq}/dx$ is comparable to either the first or second term in eq.~(\ref{Boltz'}).
This leads to
\begin{eqnarray}
\frac{d r_{eq}}{d x } \approx -\delta\, \lambda \eta g^{1/2}_* x^{-n-2} r_{eq},
\end{eqnarray}
or equivalently using the approximate relation~(\ref{xi}),
\begin{eqnarray}\label{FO}
\left(1-\frac{3}{2x_f}\right)\tanh\xi_f=\left(1-\frac{3}{2x_f}\right)\frac{1-r_{eq,f}}{1+r_{eq,f}}\approx\delta\frac{\lambda\eta g_{*,f}^{1/2}}{2 x_f^{n+2}},
\end{eqnarray}
with $\delta$ a number to be determined by fitting the numerical solution. Because $x_f$ turns out to be $\gtrsim20$ for the scenarios of interest (see the end of this section), we can ignore the $3/2x_f$ factor in the above expression.
As long as $g_{*,f}^{1/2}\lambda\eta<2x_f^{n+2}$, the condition~(\ref{FO}) can be solved iteratively. After one iteration, and with the help of equation~(\ref{xi}), we find
\begin{eqnarray}\label{xf}
x_f\approx \log(\delta g_{*,f}^{1/2}a_f\lambda)+\frac{1}{2}\log\frac{\log^3(\delta g_{*,f}^{1/2}a_f\lambda)}{\log^{2n+4}(\delta g_{*,f}^{1/2}a_f\lambda)-g_{*,f}\left(\delta \frac{\lambda\eta}{2}\right)^2}+\dots.
\end{eqnarray}
This expression is accurate at the percent level, so we will use it in our applications. Notice that the effect of a nonzero asymmetry $\eta$ on the freeze-out temperature is very mild (see for example Fig.~\ref{fig0}). In fact, for all $g_{*,f}^{1/2}\lambda\eta<2x_f^{n+2}$ the scale $x_f$ is very close to the result found in the symmetric limit, in which case $\delta$ is typically taken to be $\delta=n+1$~\cite{ST}. As already emphasized, however, the effect of $\eta$ on the anti-particle population is considerable.
The asymmetry can only be considered small if $g_{*,f}^{1/2}\lambda\eta\ll2x_f^{n+2}$, in which case one has $Y^+_\infty\simeq Y^-_\infty\gg\eta$, as anticipated in the introduction. More generally, in the ``weak" coupling regime $g_{*,f}^{1/2}\lambda\eta<2x_f^{n+2}$ we find that to a high accuracy $r_{eq,f}\simeq1$. The final estimate for the present fractional asymmetry in this limit thus reads
\begin{eqnarray}\label{estimate1}
r_\infty= e^{-\lambda\eta\,\Phi}\quad\quad \left[\textrm{if}\quad\frac{\lambda\eta g_{*,f}^{1/2}}{2 x_f^{n+2}}<
\right],
\end{eqnarray}
where $\Phi$ is understood to be $\Phi=\Phi(\infty,m)$. This result agrees with~\cite{GS}.
While it is clear that Eq.~(\ref{estimate1}) describes the regime where $r_{\infty} \lesssim 1$, it is less clear how far this approximation extends to the ``strong" coupling regime in which $r_{\infty} \ll1$. In order to appreciate the range of validity of~(\ref{estimate1}), we estimate the integral~(\ref{Phi}) assuming that $g_*^{1/2}$ is approximately constant in the relevant range of temperatures. With this approximation one finds that $\Phi\approx g_{*,f}^{1/2}/[(n+1)x_f^{n+1}]$, and~(\ref{estimate1}) becomes
\begin{eqnarray}\label{gg}
r_\infty\simeq\exp\left(-\frac{\lambda\eta g_{*,f}^{1/2}}{(n+1)x_f^{n+1}}\right)=\exp\left(-\frac{\lambda\eta g_{*,f}^{1/2}}{2x_f^{n+2}}\,\frac{2x_f}{(n+1)}\right).
\end{eqnarray}
To see how far the approximation extends, first note that for $r_{\infty} \gtrsim O(10^{-4})$, the exponent in~(\ref{gg}) should be at most $O(10)$. Since for a weak-scale cross section $x_f\sim20$, the upper bound on the exponent translates into an upper bound of $\lambda\eta g_{*,f}^{1/2}/[2x_f^{n+2}]<O(0.2)$, in agreement with our assumption~(\ref{estimate1}). We thus conclude that~(\ref{estimate1}) not only incorporates the intermediate limit anticipated in~(\ref{intermediate}), but also the ``strong" limit in which $r_{\infty} \gtrsim 10^{-4}$.
As $2x_f^{n+2}\rightarrow \lambda\eta g_{*,f}^{1/2}$ the fractional asymmetry gets further suppressed since now $r_{eq,f}$ can be much smaller than 1. Because in this regime $r_\infty$ is essentially vanishing for all practical purposes, we will be mainly concerned with the regime~(\ref{estimate1}) in the following sections.
Finally, the estimate~(\ref{estimate1}) turns out to be a very good approximation of the exact (numerical) result. We scanned a large range of parameter space ranging from $\eta=O(10^{\pm4})\eta_B$, $m=O(10^{\pm4})m_p$, and $\sigma_0$ several orders of magnitude above and below the ``thermal WIMP" value $\sigma_{0,WIMP}=O(1)$ pb for both s-wave and p-wave processes. Based on this analysis, we have found that our analytic expression departs by at most $5\%$ from the numerical result. Consistent with the derivation presented above, within the scanned range we have $\lambda\eta g_{*,f}^{1/2}/[2x_f^{n+2}]<O(0.1)$.
\section{Asymmetric WIMP dark matter}
\label{adm}
In asymmetric dark matter (ADM) scenarios the dark matter is assumed to have a primordial asymmetry $\eta$. For definiteness we will measure the DM asymmetry in units of the baryon asymmetry via the relation
\begin{eqnarray}\label{eta}
\eta=\epsilon\eta_B
\end{eqnarray}
where $\eta_B = (0.88 \pm 0.021)\times10^{-10}$~\cite{WMAPetab} \footnote{$\eta$ and $\eta_B$ are normalized relative to the total entropy density (i.e., neutrinos $+$ photons).}. The present-day abundance of dark matter and baryons is given by
\begin{equation} \label{presconst} \rho_{DM} = m~s \left(Y^{+} + Y^{-} \right) , ~~~\rho_{B} = m_{p}~s~\eta_{B}.
\end{equation}
Taking advantage of the above definitions, and using~(\ref{YY}) to express the present DM density in terms of $\eta$ and $r_\infty$, we write the observational constraint~(\ref{Omegas}) as
\begin{eqnarray}\label{const}
\frac{m}{m_p}\epsilon = \left(\frac{1-r_{\infty}}{1+r_{\infty}}\right)\frac{\Omega_{DM}}{\Omega_B},
\end{eqnarray}
with $m_p$ and $m$ the proton and DM masses, respectively, and $r_\infty\equiv Y^-_\infty/Y^+_\infty$ the present fractional asymmetry. The two present-day constraints (\ref{presconst}) are equivalent to satisfying (\ref{const}), generating the correct baryon asymmetry $\eta_B$, and having a theory for (\ref{eta}).
The literature so far has focused on the following two extreme scenarios. In the first scenario (ADM) the DM is totally asymmetric, namely $r_\infty=0$. The observed relic abundance is then explained if the DM mass is \emph{tuned} to the value
\begin{eqnarray}\label{ADMm}
\textrm{ADM:}\quad m_{ADM}= \frac{\Omega_{DM}}{\Omega_B}\frac{m_p}{\epsilon},
\end{eqnarray}
and there is no strong constraint on the cross section. The cross section simply needs to be large enough (see below) to suppress $r_\infty$.
In the second extreme scenario (thermal WIMP) the DM is totally symmetric, namely $r_\infty=1$. From our analytic expression~(\ref{estimate1}) we find that in the limit $\eta\rightarrow0$ the relation~(\ref{const}) becomes
\begin{eqnarray}\label{WIMP}
\textrm {Thermal WIMP:}\quad\quad\sigma_{0,WIMP}=\frac{1}{\eta_Bm_p}\frac{\Omega_B}{\Omega_{DM}}\frac{1}{M_{Pl}\Phi_{WIMP}}\sqrt{\frac{180}{\pi}}.
\end{eqnarray}
If $g_*$ is assumed to be constant, $\Phi=g_{*,f}^{1/2}/[(n+1)x_f^{n+1}]$, and the above formula becomes the one commonly used in the literature. The DM abundance is now determined by \emph{tuning} the annihilation cross section, the dependence on $m$ being very mild.
For the intermediate scenarios in which $0<r_\infty<1$ the physics interpolates between the above extremes. In Fig.~\ref{fig1} we plot the allowed parameter space in the $m-\sigma_0$ plane of a DM candidate satisfying the condition~(\ref{const}) for different choices of the asymmetry parameter $\epsilon$ (see Eq.~(\ref{eta})). The two figures refer to purely s-wave and p-wave processes, respectively. For a given asymmetry $\epsilon$, we see that the allowed region is bounded from below by the WIMP cross section $\sigma_{0,WIMP}$~(\ref{WIMP}), and bounded from the right as a maximum on the mass at the ADM mass value~(\ref{ADMm}).
For $\epsilon\rightarrow0$ and $m\gtrsim20$ GeV the lines asymptote to a horizontal line representing the thermal WIMP scenario, where the cross section is fixed but the mass can vary. The regime $m\lesssim20$ GeV is instead strongly sensitive to the abrupt change in the number of degrees of freedom around the QCD phase transition. The latter appears as a large gradient in $g_*^{1/2}$ at $T^*\lesssim1$ GeV~\cite{darksusy}. The effect on the relic abundance of the species is relevant if $T_f=m/x_f\lesssim T^*$, and can be seen in Fig.~\ref{fig1} as the bump at $m\lesssim x_fT^*\simeq20$ GeV. This change in the degrees of freedom also tends to reduce the exponent in~(\ref{rx}) for light species.
In Fig.~\ref{fig1} the fractional asymmetry varies from $r_\infty=0$ in the upper part of the curves to $r_\infty=1$ when the curves overlap with the thermal WIMP scenario ($\epsilon=0$). The same information is presented differently in Fig. \ref{fig2b}, where we trade $\epsilon$ for $r_{\infty}$.
In this figure the exponential sensitivity of $r_{\infty}$ to the annihilation cross-section is evident.
\begin{figure}
\begin{center}
\includegraphics[width=3.5in]{SigmaMn0.pdf}\\
\vspace{5mm}
\includegraphics[width=3.5in]{SigmaMn1.pdf}
\caption{\small Here we plot the annihilation cross section $\sigma_0$ required to reproduce the correct DM abundance $\Omega_{DM}$ via a s-wave process $n=0$ (above plot) and p-wave $n=1$ (bottom plot) for a given dark matter mass $m$, and for various values of the primordial asymmetry $\eta=\epsilon\eta_B$. The line for $\epsilon=0$ corresponds to the usual thermal WIMP scenario. Notice that the fractional asymmetry runs from $r_\infty=0$ in the upper part of the curves to $r_\infty=1$ when the lines converge on the standard thermal WIMP curve. The effect of the QCD phase transition appears as a bump at $m\lesssim20$ GeV, as anticipated in the text. Note that the bottom plot is basically enhanced by a factor $\Phi_{n=0}/\Phi_{n=1}\sim (n+1)x_f$ compared to the former. As a reference, recall that $1$ pb $\simeq2.6\times10^{-9}$ GeV$^{-2}$.
\label{fig1}}
\end{center}
\end{figure}
A generic model for beyond the SM physics is expected to lie in between the two extreme scenarios described above. In fact as emphasized in the introduction, a primordial DM asymmetry ($\epsilon\neq0$) is generally expected; meanwhile most models beyond the SM naturally lead to WIMP-like cross sections. The resulting \emph{asymmetric WIMP} scenario will typically have $\sigma_0\gtrsim\sigma_{0,WIMP}$ and $r_\infty\lesssim1$.
It is instructive to re-express the fractional asymmetry for an asymmetric DM species as a function of its thermally averaged cross section $\sigma_0$ and the quantity $\Phi=\Phi(\infty,m)$ defined in~(\ref{Phi}):
\begin{eqnarray}\label{rinf}
r_{\infty}= e^{-\lambda\eta\Phi}=\exp\left[-2\left(\frac{\sigma_0}{\sigma_{0,WIMP}}\right) \left(\frac{\Phi}{\Phi_{WIMP}}\right) \frac{1-r_{\infty}}{1+r_{\infty}}\right],
\end{eqnarray}
where $\sigma_{0,WIMP}$ and $\Phi_{WIMP}\approx\Phi$ refer to a symmetric DM candidate of the same mass. The first equality in~(\ref{rinf}) is a good analytic approximation to the numerical solution (see~(\ref{estimate1})), whereas the second is exact and follows from eliminating $\eta$, $\lambda$ and $m$ using (\ref{lambda}) and the empirical formul\ae~(\ref{const}) and~(\ref{WIMP}).
From (\ref{rinf}) we see explicitly that for WIMP-like cross-sections and larger, the dependence of $r_{\infty}$ on the cross section is very sensitive.
For a more quantitative evaluation, we show in Fig.~\ref{fig3} the solution of the above implicit equation. Indeed, we see that an $O(1)$ departure from the thermal WIMP cross section implies a significant change in $r_\infty$.
Indirect signals of annihilating asymmetric DM in the Universe are suppressed compared to those of a symmetric thermal candidate of the same mass by the quantity
\begin{eqnarray}\label{sup}
\frac{\sigma_0}{\sigma_{0,WIMP}}\times r_\infty\times \left(\frac{2}{1+r_\infty}\right)^2 \leq1.
\end{eqnarray}
One can understand the above formula as follows. The rate for indirect events is proportional to $\langle\sigma_{\textrm{\small{ann}}}v\rangle r_\infty(n^+_{\eta\neq0})^2$ for an asymmetric candidate, or similarly to $\langle\sigma_{\textrm{\small{ann}}}v\rangle (n^+_{\eta=0})^2$ for a symmetric (thermal WIMP) candidate. Notice, however, that the number densities $n^+_{\eta\neq0}$ and $n^+_{\eta=0}$ are not the same. In fact, for DM candidates of the same mass it is the total (particle plus antiparticle) densities that must be equal, namely $n_{\eta\neq0}^+(1+r_\infty)=2n^+_{\eta=0}$. Taking the ratio of these rates gives~(\ref{sup}).
We plot the quantity~(\ref{sup}) as a function of the cross section as a dashed line in Fig.~\ref{fig3}. Interestingly, we find that potentially important \emph{indirect} signals are still effective in the intermediate regime $\sigma_0\gtrsim\sigma_{0,WIMP}$ in which $r_\infty\lesssim1$. For example, a suppression of $\,\lesssim0.5$ is found for values of the fractional asymmetry below $\sim0.1$ (i.e. for $\sigma_0/\sigma_{0,WIMP}\gtrsim1.5$), when the DM could already be considered asymmetric.
\begin{figure
\begin{center}
\includegraphics[width=4.5in]{sigrinf.pdf}
\caption{\small Same as in Figure \ref{fig1} showing contours of constant $r_{\infty}$, for s--wave process. The completely asymmetric scenario $r_{\infty} \ll 1$ corresponds to the top region of the plot having cross-sections always larger than the thermal WIMP cross-section. For reference we show the $\epsilon=0$ line which corresponds to the usual thermal WIMP scenario ($r_{\infty}=0$). \label{fig2b}}
\end{center}
\end{figure}
\begin{figure
\begin{center}
\includegraphics[width=4.5in]{rinf.pdf}
\caption{\small Present fractional asymmetry $r_\infty=Y^-_\infty/Y^+_\infty$ (see (\ref{rinf})) for various values of the thermally averaged cross section $\sigma_0$ (defined as $\langle\sigma_{\textrm{\small{ann}}}v\rangle\equiv\sigma_0 (T/m)^n$ in the first section) in units of the value $\sigma_{0,WIMP}$ obtained for an exactly symmetric species of the same mass, see eq.(\ref{WIMP}). All the points in the curve, i.e. all solutions of~(\ref{rinf}), account for the observed DM density. The dashed line is the suppression factor~(\ref{sup}) for indirect detection. \label{fig3}}
\end{center}
\end{figure}
\section{Asymmetric WIMP dark matter: applications}\label{adm:app}
In the following subsections we apply our results to a few ADM models, choosing our examples from among those already existing in the literature~\cite{ADM,Xogenesis}. We demonstrate there is a large allowed region in these models in which the present-day fractional dark matter asymmetry is in the range $0<r_\infty<1$ and, {\em significantly}, the dark matter annihilation cross-section in the early Universe can be comparable to the typical WIMP cross-section. Our results generalize the conclusions found in the previous literature, which only discusses completely asymmetric dark matter, corresponding to a present fractional asymmetry of $r_\infty=0$ in our notation.
In the examples we discuss the MSSM is assumed to couple to dark matter superfields $X,\bar X$ via a heavy messenger sector. The dark matter chiral superfields $X, \overline{X}$ carry $\pm 1$ charge under a hidden sector global $U(1)$, are Standard Model gauge singlets, and have a SUSY invariant mass $m_X$. For definiteness we assume that after SUSY breaking the fermionic component $\psi_X$ is the dark matter. Similar results hold for scalar DM candidates.
The overall system is assumed to violate either lepton ($L$) number or baryon ($B$) number together with the dark ($X$) number, but to (perturbatively) preserve a linear combination of these symmetries. When the number-violating interactions are in thermal equilibrium, chemical equilibrium forces a relation between the asymmetries of the visible and dark sectors and therefore determine the quantity $\epsilon$ introduced earlier~(\ref{eta}). We will refer to this effect as a transfer mechanism.
\subsection{Example 1: Lepton-number violating transfer operator}
\label{example1}
In the first example we consider lepton number violation and take the leading transfer mechanism between the dark and leptonic sectors, after having integrated out the messengers, to be described by a nonrenormalizable term in the superpotential~\cite{ADM}:
\begin{eqnarray}\label{transf3}
\Delta W_{asym}= \frac{XXH_uL}{\Lambda},
\end{eqnarray}
with $\Lambda$ taken to be close to the TeV scale. The operator~(\ref{transf3}) preserves a global $L'=X-2L$ number, but violates both $L$ and $X$. The quantum numbers $L,X$ become good quantum numbers in the early Universe as soon as the transfer operator~(\ref{transf3}) freezes out, assumed to occur for simplicity at an instantaneous temperature $T_D$. One can show that for $\Lambda=O(1)$ TeV $T_D$ is below the scale $T_{sph}$ at which the sphalerons shut off, in which case the only source of lepton number violation at low temperatures is induced by~(\ref{transf3}).
This model has either one or two stable particles, depending on the relative mass difference between the $\psi_X$'s and the lightest superpartner of the SM. To see that note that the model has a $Z_4$ symmetry (see also~\cite{ADM}) which is the usual $Z_2$ $R-$parity in the visible sector, plus a $Z_4$ in the $X$ sector that acts as $+i$ on the fermionic component $\psi_X$ and $-i$ on the scalar component $\widetilde{X}$ (the $\overline{X}$ fields have the opposite charge). This symmetry remains unbroken by soft SUSY and electroweak symmetry breaking, and guarantees that at least one component of $X$ is always stable. The other SUSY component of $X$ generically decays through the transfer operator to the lightest component of $X$ and SM fields.
The transfer operator typically causes the lightest superpartner (LSP) of the SM to be unstable, if the channel is kinematically allowed. (The decay $\hbox{LSP} \rightarrow XX + SM ~particles$ is consistent with the $Z_4$ symmetry).
If $m_{LSP} < 2 m_X$ then this decay is forbidden and the model has two stable particles. In what follows we will assume $m_{LSP} > 2 m_X$ for simplicity.
\subsubsection{Chemical Potential Analysis}
To estimate the primordial DM asymmetry parameter $\epsilon$, see~(\ref{eta}), we proceed in two steps. For a useful review on chemical potential analysis see~\cite{HT}.
As a first step we impose chemical equilibrium and charge neutrality at the scale $T=T_{sph}>T_D$ at which the sphalerons decouple. At these temperatures the only dynamical particles are the DM and the Standard Model fields (including the top quark), and we have:
\begin{eqnarray}\label{chem}
Q&\propto&16\mu_u-10\mu_d-6\mu_e=0\\\nonumber
\mu_u-\mu_d&=&\mu_\nu-\mu_e\\\nonumber
\mu_u+2\mu_d+\mu_\nu&=&0\\\nonumber
2\mu_X+\mu_\nu&=&0,
\end{eqnarray}
where the chemical potentials $\mu_u, \mu_d, \mu_\nu, \mu_e,$ and $\mu_X$ refer to the Standard Model up, down quarks, the (purely left handed) neutrinos, the charged leptons, and the DM respectively. The first equation in~(\ref{chem}) imposes electrical neutrality, the second accounts for the $W^\pm$ exchange, the third for the sphaleron process, and finally the last equation follows from the transfer operator~(\ref{transf3}). The solution of the system~(\ref{chem}) can be used to derive the primordial baryon and $L'$ asymmetries:
\begin{eqnarray}\label{BL'}
\eta_B&= &-\frac{16}{3}\mu_e\\\nonumber
\eta_{L'}&=&\frac{1}{16}\left(66+5 \frac{f(m/T_{sph})}{f(0)}\right)\eta_B,
\end{eqnarray}
where the function $f(x)$ is defined by
\begin{eqnarray}\label{boltz}
f(x)=\frac{1}{4\pi^2}\int_0^\infty dy \frac{y^2}{\cosh^2\left(\frac{1}{2}\sqrt{x^2+y^2}\right)},
\end{eqnarray}
and accounts for the possibility that the DM be non-relativistic at the relevant temperature~\cite{Barr:1990ca}. If the DM is instead bosonic the $\cosh(x)$ function should be replaced by $\sinh(x)$.
Below the sphaleron temperature the asymmetries for $B$ and $L'$ given above are conserved, while $L$ and $X$ are still violated by~(\ref{transf3}). In the second step of our chemical potential analysis we thus compute the present asymmetry for the $X$ number at the temperature $T_D$ at which the transfer operator decouples.
Chemical equilibrium at $T_D$ imposes a number of conditions similar to~(\ref{chem}). The two main differences are that at this scale the sphaleron process (see third line in~(\ref{chem})) is no more effective, and that the top quark is not included. The system of equations for the chemical potentials is now solved by replacing the sphaleron process constraint with the initial conditions provided by~(\ref{BL'}), and not including the top quark contribution. The result is finally
\begin{eqnarray}\label{epsilonHuLXX}
\epsilon\equiv\frac{\eta_X}{\eta_B}=\left(\frac{2934+215 \frac{f(m/T_{sph})}{f(0)}}{9504+688 \frac{f(m/T_D)}{f(0)}}\right)\frac{f(m/T_D)}{f(0)}.
\end{eqnarray}
Our result~(\ref{epsilonHuLXX}) reduces to the one found in~\cite{ADM} if the top quark is instead decoupled above the sphaleron scale, and if one assumes $f(m/T_D)=f(m/T_{sph})=f(0)$. We also checked that the result changes only mildly if one assumes that the massive vector bosons are decoupled at the scale $T_D$. In all these cases one finds that~(\ref{epsilonHuLXX}) can be well approximated as $\epsilon\approx0.3 f(m/T_D)/f(0)$.
\subsubsection{Transfer Operator Decoupling}
In order for the analysis of the previous sections to be applicable, the freeze-out temperature $T_f$ must be below the scale $T_D$ at which the transfer operator decouples~\footnote{We will comment on the limiting case $T_D=T_f$ in Section~\ref{discuss}. For the moment, notice that the case $T_D<T_f$ is clearly meaningless: freeze-out always occurs when the strongest interaction decouples.}: only under this condition there exists a temperature range $T_f\leq T\leq T_D$ in which Eq.~(\ref{Boltz}), and the subsequent analysis, is fully reliable. We will see in subsection~\ref{anndd} that, in typical UV completions of the model~(\ref{transf3}), this condition is ensured by higher dimensional operators that preserve the dark number and $L$, and which are therefore not involved in transferring asymmetries.
The Boltzmann equation governing the evolution of the $X$ number at $T \gtrsim T_{D}$ is given by Eq.~(\ref{Boltz}), but with additional, $X$-number changing terms induced by~(\ref{transf3}). These latter involve the DM and the SM particles as final states, with typical reactions of the form
\begin{eqnarray}\label{XXnu'}
\psi_X \psi _X \leftrightarrow \overline{\psi}_X \overline{\psi}_X \overline{\nu} \overline{ \nu}.
\end{eqnarray}
The decoupling temperature $T_D$ is then estimated as the scale at which the strongest interaction mediated by the transfer operator -- formally described by a ``collision term" $C$ on the right hand side of~(\ref{Boltz}) -- becomes comparable to the Hubble expansion rate, namely when
\begin{eqnarray}\label{TD}
C=\,H(T_D)n_X,
\end{eqnarray}
with $n_X$ the DM equilibrium number density.
If the transfer operator decouples when the DM is relativistic, the Boltzmann suppression in~(\ref{epsilonHuLXX}) is not effective and the parameter $\epsilon$ does not depend on the DM mass. In this case the details of the transfer mechanism are irrelevant.
The Boltzmann suppression in~(\ref{epsilonHuLXX}) is effective if the transfer operator decouples when the DM is non-relativistic. In this case only the precise value of $T_D$, and hence the details of the transfer operator matter. In the specific case in which the neutralino is the LSP, we find that the dominant number-violating interaction mediated by~(\ref{transf3}) in the non-relativistic limit is associated to the process~(\ref{XXnu'}), and occurs via the production of an on-shell intermediate neutralino which eventually decays with a branching ratio BR $=1/2$ into an $X$- and $L$-number violating final state $\psi_X \psi _X \rightarrow \tilde{\chi} \overline{\nu}\rightarrow\overline{\psi}_X \overline{\psi}_X \overline{\nu} \overline{ \nu}$.
If we neglect the DM mass, then at leading order in $T/m_{ \tilde{\chi}}$ the collision term $C$ reduces to (see Appendix \ref{lhu})
\begin{eqnarray}\label{C}
C=-\frac{\sqrt{2\pi}}{4(2\pi)^5}\left(\frac{g_w v_u}{m_{\tilde\nu}^2\Lambda}\right)^2\, m_{ \tilde{\chi}}^8\left(\frac{T}{m_{ \tilde{\chi}}}\right)^{9/2}\,e^{-m_{ \tilde{\chi}}/T}.
\end{eqnarray}
In the above expression, $g_w$ is a weak gauge coupling, $v_u=v\sin\beta$ is the vacuum expectation value of the up-type Higgs, and $m_{\tilde\nu}$ is the sneutrino mass. The exponential suppression in~(\ref{C}) reflects the fact that the process occurs far in the Boltzmann tail of the thermal distribution of the incoming DM particles.
The factors of $T$ in (\ref{C}) arise from a partial integration in the process of taking the thermal average, and reflect the fact that the phase space integral vanishes when the energy of the DM particles is at threshold. Another way to understand these factors of $T$ is to note that for the inverse process $\nu \tilde{\chi} \rightarrow \psi_{X} \psi_{X}$ the powers of $T$ arise solely from the number densities of the incoming particles simply because there is no threshold and the matrix element is set by the mass of the LSP. The inverse collision term is therefore proportional to $n_{\tilde{\chi}} n_{\nu}\sim T^{3/2} T^{3} e^{-m_{\tilde{\chi}}/T}$.
The decoupling temperature is finally estimated using~(\ref{TD}), with $C$ given by~(\ref{C}) and $n_X$ the non-relativistic DM equilibrium density. We find that, up to negligible logarithmic corrections, $T_D$ is given by
\begin{eqnarray}
T_D\approx \frac{m_{ \tilde{\chi}}-m_X}{\log\gamma},~~\quad\gamma=\frac{1}{4(2\pi)^3}\left(\frac{g_wv_u}{m_{\tilde\nu}^2\Lambda}\right)^2\frac{(m_{ \tilde{\chi}}-m_X) M_{Pl}m_{ \tilde{\chi}}^{7/2}}{\left(\frac{8 \pi^{3}}{90}\right)^{1/2} g_{\textrm{\small{eff,D}}}^{1/2}\,m_X^{3/2}}
\end{eqnarray}
which approximately reads $T_D\sim(m_{ \tilde{\chi}}-m_X)/20$ if the mass scales are of $O($GeV-TeV$)$. We now have all the ingredients required to estimate the primordial asymmetry parameter $\epsilon$, see eq.~(\ref{epsilonHuLXX}), as a function of the DM mass and the LSP mass $m_{ \tilde{\chi}}$. This estimate has been found here working under the assumption that the transfer operator decouples when the DM is already non-relativistic. Yet, because the Boltzmann suppression in~(\ref{epsilonHuLXX}) ceases to be effective when the non-relativistic limit is still a good approximation, it is clear that our estimate of $\epsilon$ should also apply to the case in which the transfer operator decouples when the DM is relativistic. As emphasized above, indeed, in this latter case the details of the transfer mechanism do not matter.
\subsubsection{\label{discuss}Discussion}
The main purpose of the present subsection is to scan the parameter space of the model~(\ref{transf3}) and show that there is a sizeable region in the allowed parameter space with a present-day fractional asymmetry in the range $0.1<r_\infty<1$. This is the asymmetric WIMP regime.
The relevant parameters of the model~(\ref{transf3}) at low temperatures reduce to the decoupling temperature $T_D$, the DM mass $m_X$, and the annihilation cross section (encoded in $r_\infty$). The requirement that the particle $\psi_X$ saturates the observed DM density is implemented by~(\ref{const}) and it amounts to a constraint on these parameters. In Fig.~\ref{figHuLXX} we show the resulting contour plot for the quantity $r_\infty$ \emph{defined} by~(\ref{const}) as a function of $m_X$ and $m_{ \tilde{\chi}}$.
\begin{figure
\begin{center}
\includegraphics[width=4.5in]{nicerplot.pdf}
\caption{\small Contour plot for $r_\infty$ in the allowed parameter space for the model~(\ref{transf3}). Below the lower line and above the upper line the DM density is smaller and higher than experimentally observed, respectively. In the dark blue area $0.9\leq r_\infty<1$, in the blue area $0.1\leq r_\infty\leq0.9$, and in the light blue area $0\leq r_\infty\leq0.1$. See the text for more details. Here we have taken $m_{\tilde{\nu}} = \Lambda = 1$ TeV and $\tan \beta = 20$. \label{figHuLXX}}
\end{center}
\end{figure}
The upper curve shows the line~(\ref{ADMm}) at which $r_\infty=0$. In the area above this curve the DM abundance exceeds the observed value. This region of the parameter space is therefore excluded. The lowest curve is determined by the condition $T_D= T_f\sim m_X/20$, which approximately reads $m_{ \tilde{\chi}}\sim2 m_X$ in our case, at which $r_\infty\simeq1$ (and $\epsilon\ll1$). In the area below this latter curve the relic abundance of the species $\psi_X$ is smaller than the observed value. This area is excluded here because $\psi_X$ is assumed for simplicity to be the only DM candidate. The intermediate lines in Fig.~\ref{figHuLXX} separate regions with different present-day fractional asymmetry $r_\infty$.
The plot can be understood as follows. For any $r_\infty$, a given value for the DM mass completely determines $\epsilon$ with~(\ref{const}). This fixes $T_D$ and in turn it determines a function $m_{ \tilde{\chi}}(m_X)$. As long as the Boltzmann suppression in~(\ref{epsilonHuLXX}) is active the curve is approximately a straight line determined by $m_X/T_D\sim const\gtrsim1$. This behavior is visible in the large DM mass regime of Fig.~\ref{figHuLXX}. The Boltzmann suppression shuts off when the previous lines hit the boundary of the non-relativistic regime $m_X/T_D\sim1$. For DM masses below this point the primordial asymmetry parameter is given by $\epsilon\sim0.3$ and $m_X$ is fully determined by $r_\infty$. This effect is seen as a vertical asymptote in the low DM mass regime of Fig.~\ref{figHuLXX}.
It is evident from Figure~\ref{figHuLXX} that a large region in parameter space is associated to the asymmetric WIMP scenario $0.1\lesssim r_\infty\lesssim1$. Here indirect signals are unsuppressed.
\subsubsection{Annihilation and direct detection}
\label{anndd}
In all the cases discussed above -- either the strongly asymmetric $r_\infty\ll1$ or strongly symmetric $r_\infty\simeq1$ limit -- we see from Fig.~\ref{fig3} that the annihilation cross section can be of thermal WIMP magnitude. Such a relatively low cross section may be induced by annihilation into SM particles mediated by non-renormalizable operators suppressed by the very same scale $\Lambda$ suppressing~(\ref{transf3}). From an effective field theory perspective in fact we expect that a generic UV completion of~(\ref{transf3}) also leads to direct couplings between the DM and the SM particles. For example, the following corrections to the superpotential and the K\"{a}hler are allowed by all the symmetries
\begin{eqnarray}\label{Wother}
\Delta W_{sym}&=&{\lambda}^2\frac{X\overline{X}H_uH_d}{\Lambda}+\dots\\
\Delta K_{sym} &=& y^2\frac{X^\dagger XL^\dagger L}{\Lambda^2}+\dots, \label{Kother}
\end{eqnarray}
and are therefore generic
We now show that both operators~(\ref{Wother}) and~(\ref{Kother}) provide annihilation cross sections of the right strength if $\Lambda=O( \hbox{TeV})$. These operators can therefore be responsible for keeping the DM in thermal equilibrium below $T_D$, and in particular for accounting for the values of $r_\infty$ required to generate the correct relic abundance. As we will see, they can also provide direct detection signals. Let us now discuss a few annihilation modes in more detail.
The correction to the K\"{a}hler triggers the process $\psi_X\overline{\psi_X}\rightarrow l\bar l$, where $l$ are SM leptons and $\psi_X$ is a Dirac DM fermion. In order to avoid large flavor violating effects that might arise from~(\ref{Kother}), we follow~\cite{ADM} and assume that the only relevant coupling is with the third SM lepton generation. Neglecting the $\tau$ and $\nu_\tau$ masses we find~\cite{ADM}:
\begin{eqnarray}\label{annl}
\langle\sigma_{\textrm{ann}}v\rangle=\frac{y^4}{16\pi}\frac{m_X^2}{\Lambda^4}.
\end{eqnarray}
It is easy to see that this channel can provide an annihilation cross section of the correct order for several DM masses. For example, for a light DM candidate, say with $m_X\sim20$ GeV one should require $\Lambda/y\lesssim200$ GeV, whereas for weak-scale DM masses, say $m_X\sim200$ GeV one sees that $\Lambda/y\lesssim700$ GeV suffices. If~(\ref{Wother}) provides the dominant annihilation mode, one finds no constraining direct detection bounds~\cite{ADM}.
The dominant annihilation mode mediated by the operator $X\overline{X}H_uH_d$, on the other hand, occurs via the s-wave annihilation either to final state fermions -- through the exchange of the (virtual) CP odd scalar $A^0$ -- or to a $h A^0$ final state. For the first channel the largest rate occurs if the final state fermions are top quarks, if kinematically allowed. Specifically, for $\overline{\psi_X}\psi_X\rightarrow {A^0}^{*}\rightarrow\bar t t$ and $ m_X > m_t$, we have
\begin{eqnarray}\label{XXtt}
\langle\sigma_{\textrm{ann}}v\rangle=\frac{\lambda^4}{\tan^2\beta}\frac{N_c}{8\pi}\frac{m_t^2}{\Lambda^2}\frac{m_X^2}{(4m_X^2-m_{A^0}^2)^2} \sqrt{1-\frac{m_t^2}{m_X^2}}.
\end{eqnarray}
This can be of the correct magnitude to account for the DM abundance if $A^0$ is not too heavy, $m_X$ is not too small, and $\tan\beta$ not too large.
For example, if we take $m_X=200$ GeV $>m_{A^0}/2$ and $\tan^2\beta\approx1$
the averaged cross section is close to the WIMP value; for these values $\langle\sigma_{\textrm{ann}}v\rangle\approx\lambda^4\times4\times 10^{-9}$ GeV$^{-2} (1.2 \hbox{ TeV}/\Lambda)^2$. If this channel is kinematically closed the other channel may be open and provide a large enough rate. For $2m_X > m_{h} + m_{A^0}$, we have for the s-wave annihilation $\overline{X}X \rightarrow h A^0$,
\begin{eqnarray}\label{XXAh}
\langle\sigma_{\textrm{ann}}v\rangle=\lambda^4 \frac{\sin^2 (\beta -\alpha) }{64 \pi \Lambda^2}
\sqrt{1-\frac{(m_{A^0}+m_{h})^2}{4 m^2_X} } \sqrt{1-\frac{(m_{A^0}-m_{h})^2}{4 m^2_X} } .
\end{eqnarray}
Unlike the previous cross-section, this one is not suppressed at large $\tan \beta$.
If we take $\sin^2 (\beta -\alpha) \approx 1$, the averaged cross-section is
$\langle\sigma_{\textrm{ann}}v\rangle\approx\lambda^4 \times4\times 10^{-9}$ GeV$^{-2} (1.5 \hbox{ TeV}/\Lambda)^2$.
If~(\ref{Wother}) provides the dominant annihilation mode, the DM direct detection bounds are not generally negligible. In the absence of light degrees of freedom other than the MSSM particles and the DM, the ``single nucleon'' cross section for spin-independent elastic DM scattering mediated by virtual exchanges of CP-even Higgses is
\begin{equation}
\label{sigxn}
\sigma_{Xn} = \left(\frac{1~{\rm GeV}}{\mu_{T}}\right)^{2} \frac{\sigma_{N}}{A^{2}},
\end{equation}
where the spin-independent dark matter-nucleus cross section is
\begin{equation} \sigma_{N} =\frac{\mu_{T}^{2}}{\pi} \left(Z f_{p} + (A-Z) f_{n}\right)^{2} \end{equation}
with $f_{p}$ and $f_{n}$ controlling the coupling between nucleons and the dark matter, and $\mu_{T}$ the DM-nucleus reduced mass. The nucleon coupling parameters $f_{p}$ and $f_{n}$ can be found in Appendix~\ref{appDD} and depend on the scale in Eq.~(\ref{Wother}) as well as parameters in the Higgs sector.
Requiring that the annihilation cross section is larger than the WIMP value implies a lower bound on DM-nucleon cross section. Similarly, the bounds from direct detection can be translated into an upper bound for the annihilation cross section and thus a lower bound for $r_\infty$.
\begin{figure
\begin{center}
\mbox{\subfigure{\includegraphics[width=2.8in]{ddbounds700.pdf}}~~~
\subfigure{ \includegraphics[width=2.8in]{ddbounds1000.pdf}}}
\caption{\small Here we show the impact of the CDMS-II bound~\cite{cdms} $\sigma_{Xn} = 5\times 10^{-44}~\rm{cm}^{2}$ and the projected XENON100 bound~\cite{dmtools} $\sigma_{Xn} = 7\times 10^{-45}~\rm{cm}^{2}$ on the parameter space of our model for $m_{X}=400$ GeV and $\Lambda = 700$ GeV (left plot) and $\Lambda = 1000$ GeV (right plot). To be consistent with the null results of CDMS-II, a parameter point must lie below the red curve, while to be consistent with projected XENON100 bounds a point must lie to the right of the dashed blue curve. Note that for these parameters, there is no allowed region consistent with the projected XENON100 bounds. Obtaining the correct relic abundance requires $\langle\sigma_{\textrm{ann}}v\rangle\gtrsim \sigma_{0,WIMP}$ which occurs for points to the left of the solid black curve. For reference we include the dashed black curve corresponding to $\langle\sigma_{\textrm{ann}}v\rangle= 2 ~\sigma_{0,WIMP}$, which is not present in the $\Lambda = 1000$ GeV plot. \label{figdd}}
\end{center}
\end{figure}
To see that, let us assume that the dominant annihilation mode is $\overline{X}X \rightarrow h A^0$ and that $m_X$ is at the weak scale. We also know from Fig.~\ref{fig1} that $\langle\sigma_{\textrm{ann}}v\rangle\gtrsim \sigma_{0,WIMP} = 4.5\times10^{-9}$ GeV$^{-2} $. The CDMS-II bound along with this constraint on the annihilation cross section is shown in Fig.~\ref{figdd}. We also include the projected XENON100 $\sigma_{Xn} = 7\times 10^{-45}~\rm{cm}^{2}$ bound assuming 6000 kg-days of data~\cite{dmtools}. In Fig.~\ref{figdd} we have taken a dark matter mass $m_{X} = 400$ GeV, a transfer operator scale $\Lambda = 700$ GeV (left plot) and $\Lambda = 1000$ GeV (right plot) and a light Higgs mass $m_{h} = 120$ GeV. We have also assumed $m_{A^{0}}^{2} \gg m_{Z}^{2}$ such that the mixing angle $\alpha$ is determined at tree-level by $\alpha \approx \beta - \pi/2$ and the heavy Higgs mass is $ m_{H}^{2} \approx m_{A^{0}}^{2}$. The relation between $\alpha$ and $\beta$ implies that the annihilation cross section contours appear as straight vertical lines. The interested reader can find additional details in Appendix~\ref{appDD}. This cross section~(\ref{sigxn}) is close to the present bounds from XENON 100~\cite{xenon100} and CDMS~\cite{cdms}, and for these parameters will be completely covered by the projected bounds for XENON 100.
Our analysis of the annihilation modes has been model-independent so far. We end this section by emphasizing that, once a UV completion for the operator~(\ref{transf3}) is known, alternative reactions, and constraints, might arise. As an instance, consider the UV completion proposed in~\cite{ADM}. One introduces two electro-weak doublets $D, \bar D$ with hypercharge $\mp\frac{1}{2}$ and generalized lepton number $L'=\mp1$, and two SM singlets $X, \bar X$ with $L'=\pm1$. One readily sees that integrating out the $D$ field one gets the operator~(\ref{transf3}) as well as~(\ref{Wother}) and~(\ref{Kother}). A typical feature of this UV completion is the occurrence of a mixing between the neutral component of the doublet $D$ and the DM after electro-weak symmetry breaking. This mixing induces a coupling of the DM $\psi_X$ to the $Z^0$ of order $\sim g\lambda_{u,d}v_{u,d}/\Lambda$, and thus implies new annihilation and direct detection modes mediated by a virtual gauge boson. Given the lower bound on the annihilation cross section, one finds that the direct detection signal is always too large if the mixing-induced $Z^0$ exchange is the dominant channel. One should then suppress this channel by taking $\lambda_{u,d}<O(10^{-1})$. At this point the dominant annihilation mode would be given by~(\ref{annl}). Alternatively one could consider the NMSSM, where a large contribution to~(\ref{Wother}) might be obtained integrating out the singlet $S$, also coupled to the DM via $\lambda'S\bar X X$. In this latter case the coupling of~(\ref{Wother}) would be naturally at the TeV scale, and freeze-out would thus be controlled by either $\overline{\psi_X}\psi_X\rightarrow {A^0}^{*}\rightarrow\bar t t$ or $\overline{X}X \rightarrow h A^0$. Notice that in all of the cases discussed above the DM is assumed to be the lightest state beyond the MSSM.
\subsection{Example 2: Baryon-number violating transfer operator}
We now consider an example in which the transfer operator violates $B$ and $X$, but not $L$. The model is again an extension of the MSSM, and it is described by the following nonrenormalizable term~\cite{ADM}:
\begin{eqnarray}\label{op}
\Delta W_{asym}=\frac{XXu^c d^c d^c}{\Lambda^2}.
\end{eqnarray}
The superpotential operator~(\ref{op}) preserves a global $B'=X+2B$ number, but violates $B$ and $X$ separately. The physics of this model is very similar to the lepton violating operator~(\ref{transf3}), see for example the subsection~\ref{example1}.
A chemical potential analysis similar to the one presented for the $L$-violating operator in this case gives:
\begin{eqnarray}\label{epsilonXXqqq}
\epsilon\equiv\frac{\eta_X}{\eta_B}=\left(\frac{64-5 \frac{f(m/T_{sph})}{f(0)}}{192-16 \frac{f(m/T_D)}{f(0)}}\right)\frac{f(m/T_D)}{f(0)}.
\end{eqnarray}
Similarly to~(\ref{epsilonHuLXX}), this result is very robust and approximately reads $\epsilon\approx0.3 f(m/T_D)/f(0)$.
The typical processes transferring the asymmetry between the visible and the dark sectors
can occur through on-shell or off-shell neutralinos or squarks. The processes involving neutralinos are always strongly suppressed by phase space, and we find that on-shell production of squarks dominates for typical masses; see Appendix~\ref{appudd} for details. Using again our approximate expression~(\ref{Capp}) the collision term for the process $\psi_X \psi_X \rightarrow\tilde{u}\tilde{d}\tilde{d}$ is found to be
\begin{eqnarray}
C=-\frac{9\sqrt{2}\pi^{3/2}}{32\Lambda^4}\frac{N_c!}{(2\pi)^7}\,m_{\tilde q}^8\left(\frac{T}{m_{\tilde q}}\right)^{9/2}\,e^{-3m_{ \tilde{q}}/T}\left[1+O\left(\frac{T}{m_{ \tilde{q}}},\frac{m_X}{m_{\tilde{q}}}\right)\right].
\end{eqnarray}
where the powers of $T$ have the same origin as the ones in~(\ref{C}), and we assumed that the squarks are degenerate for simplicity.
The analysis can now proceed similarly to the lepton-number violating transfer operator. The result can again be summarized with a plot analogous to Fig.~\ref{figHuLXX}, which for brevity we do not show. The comments in Section~\ref{example1} regarding s--wave annihilation and direct detection via the superpotential operator $ X \overline{X} H_u H_d$ also apply in this case.
\section{Conclusions}
\label{conc}
In this paper we studied the implication of a nontrivial DM primordial asymmetry $\eta$ on the DM relic abundance. For any $\eta$, the physical quantity controlling the effective asymmetry of the species is the fractional asymmetry $r(x)=Y^-/Y^+$. By solving the coupled Boltzmann equations for $Y^\pm$ we found that $r(x)$ depends \emph{exponentially} on both the annihilation cross section and $\eta$, see~(\ref{estimate1})~\cite{GS}. The present anti-particle population of an asymmetric DM species is therefore extremely sensitive to the strength of the dynamics.
The parameter space in a DM model with a primordial asymmetry $\eta$ is defined as follows. The DM mass $m$ is bounded from above by the mass $m_{ADM}$ of a totally asymmetric, $r_\infty=0$, DM species, see~(\ref{ADMm}):
\begin{eqnarray}
m\leq m_{ADM}\equiv\frac{\Omega_{DM}}{\Omega_B}\frac{\eta_B}{\eta}m_p.
\end{eqnarray}
\emph{For any $\eta$} one finds $m< m_{ADM}$ as soon as one departs from the condition $r_\infty=0$. In particular, very low masses $m\ll m_p$ are always allowed provided $r_\infty\rightarrow1$.
The thermally averaged annihilation cross section for an asymmetric DM candidate is instead bounded from below by the value found for a totally symmetric species of the same mass (thermal WIMP value), see~(\ref{WIMP}):
\begin{eqnarray}
\langle\sigma_{\textrm{\small{ann}}}v\rangle\geq\langle\sigma_{\textrm{\small{ann}}}v\rangle_{WIMP}.
\end{eqnarray}
\emph{For any $\eta$} the cross section approaches its lowest value when $r_\infty\rightarrow1$, while a larger cross section can still be accommodated in the regime $r_\infty<1$. Notice that because of the exponential dependence of $r_\infty$ on the cross section, an effectively asymmetric DM population, say with $r_\infty\lesssim10^{-(2-3)}$, is already obtained with an annihilation cross section of just a factor $\sim(2-3)$ bigger than the thermal WIMP cross section. This latter point has interesting consequences for model-building, as it implies that new light states, often introduced in previous studies of asymmetric DM scenarios to induce large cross section
, are not necessary.
A generic weakly coupled UV completion for the standard model will typically have weak scale annihilation cross sections. If the DM is stabilized by a symmetry distinguishing particles from anti-particles it is also likely that the DM has a nonzero primordial asymmetry $\eta$. As a result, asymmetric WIMP scenarios are expected to be generic. These scenarios typically lie between the extremely asymmetric and symmetric limits. Surprisingly, the indirect signals from present-day annihilation can be much larger than what one would naively expect from an asymmetric scenario.
To illustrate the generality of our conclusions we applied our results to two existing models of asymmetric DM~\cite{ADM}. We saw that the parameter space consistent with the present-day DM abundance and characterizing the asymmetric WIMP regime is sizable. This property stems from the fact that the present-day fractional asymmetry $r_\infty$ acts as new, low energy parameter.
Moreover direct detection bounds impose interesting limits on the Higgs portal operator $\overline{X}XH_{u}H_{d}$. Although this operator was introduced to yield a sufficiently large annihilation cross section, the same interaction gives the dominant interaction with nuclei. We found that the current bounds from CDMS-II and XENON100 constrain the model, but upcoming direct detection experiments will be sensitive to a wide range of the parameters relevant for the asymmetric WIMP through the Higgs portal.
\acknowledgments
This work has been supported by the U.S. Department of Energy at Los
Alamos National Laboratory under Contract No. DE-AC52-06NA25396. The preprint number for this manuscript is LA-UR 11-00565.
|
\section{Introduction}
Consider a set $P$ of points in the plane and a set of closed
polygonal obstacles whose vertices together with the points in $P$
are in {\em general position}, that is, no {\em three} of them are
on a line. The corresponding {\em visibility graph} has $P$ as its
vertex set, two points $p,q\in P$ being connected by an edge if
and only if the segment $pq$ does not meet any of the obstacles.
Visibility graphs are extensively studied and used in
computational geometry, robot motion planning, computer vision,
sensor networks, etc.; see \cite{BKOS00}, \cite{G07}, \cite{OR97},
\cite{O99}, \cite{Ur00}.
Alpert, Koch, and Laison \cite{AKL09} introduced an interesting
new parameter of graphs, closely related to visibility graphs.
Given a graph $G$, we say that a set of points and a set of
polygonal obstacles as above constitute an {\em obstacle
representation} of $G$, if the corresponding visibility graph is
isomorphic to $G$. A representation with $h$ obstacles is also
called an $h$-obstacle representation. The smallest number of
obstacles in an obstacle representation of $G$ is called the {\em
obstacle number} of $G$ and is denoted by ${\rm obs}(G)$. Alpert
et al. \cite{AKL09} proved that there exist graphs with
arbitrarily large obstacle numbers.
Using tools from extremal graph theory, it was shown in \cite{PS}
that for any fixed $h$, the number of graphs with obstacle number
at most $h$ is $2^{o(n^2)}$. Notice that this immediately implies
the existence of graphs with arbitrarily large obstacle numbers.
In the present note, we establish some more precise estimates.
\begin{theorem*}\label{enumeration}
(i) For any positive integer $h$, the number of graphs on $n$
(labeled) vertices that admit a representation with $h$ obstacles
is at most $$2^{O(hn log^2 n)}.$$
(ii) Moreover, the number of graphs on $n$ (labeled)
vertices that admit a representation with a set of obstacles
having a total of $s$ sides, is at most $$2^{O(n log n + s log
s)}.$$
\end{theorem*}
In the above bounds, it makes no difference whether we count
labeled or unlabeled graphs, because the number of labeled graphs
is at most $n!=2^{O(n log n)}$ times the number of unlabeled
ones.
It follows from Theorem~\ref{enumeration} (i) that for every $n$,
there exists a graph $G$ on $n$ vertices with obstacle number
$${\rm obs}(G)\geq \Omega\left({n}/{\log^2n}\right).$$
Indeed, as long as $2^{O(hn\log^2n)}$ is smaller than
$2^{\Omega(n^2)}$, the total number of (labeled) graphs with $n$
vertices, we can find at least one graph on $n$ vertices with
obstacle number $h$.
Here we show the following slightly stronger bound.
\begin{theorem*}\label{concave}
For every $n$, there exists a graph $G$ on $n$ vertices with
obstacle number
$${\rm obs}(G)\geq \Omega\left({n}/{\log n}\right).$$
\end{theorem*}
This comes close to answering the question in \cite{AKL09} whether
there exist graphs with $n$ vertices and obstacle number at least
$n$. However, we have no upper bound on the maximum obstacle
number of $n$-vertex graphs, better than $O(n^2)$.
Our next theorem answers another question from \cite{AKL09}.
\begin{theorem*}\label{exactly}
For every $h$, there exists a graph with obstacle number exactly
$h$.
\end{theorem*}
A special instance of the obstacle problem has received a lot of
attention, due to its connection to the Szemer\'edi-Trotter
theorem on incidences between points and lines~\cite{ST83a},
\cite{ST83b}, and other classical problems in incidence
geometry~\cite{PA95}. This is to decide whether the obstacle
number of ${\overline{K}}_n$, the empty graph on $n$ vertices, is
$O(n)$ if the obstacles must be {\em points}. The best known upper
bound is $n2^{O(\sqrt{\log n})}$ is due to Pach \cite{Pach03}; see
also Dumitrescu et al.~\cite{DPT09}, Matou\v sek~\cite{M09}, and
Aloupis et al. \cite{A+10conf}.
It is an interesting open problem to decide whether the obstacle
number of planar graphs can be bounded from above by a constant. For
outerplanar graphs, this has been verified by Fulek, Saeedi, and
Sar{\i}\"oz~\cite{FSS}, who proved that every outerplanar graph has
obstacle number at most $5$.
Theorem $i$ is proved in Section $i+1,\; 1\le i\le 3$.
\section{Proof of Theorem \ref{enumeration}}\label{sec:convex}
We will prove the theorem by a simple counting method.
Before turning to the proof, we
introduce some terminology. Given any placement (embedding) of the
vertices of $G$ in general position in the plane, a {\em straight-line drawing} or, in short, a {\em drawing}
of $G$ consists of the image of the embedding and the set of {\em
open line segments} connecting all pairs of points that correspond to the edges of $G$. If there is no danger of confusion, we make no
notational difference between the vertices of $G$ and the
corresponding points, and between the pairs $uv$ and the
corresponding open segments. The complement of the set of all
points that correspond to a vertex or belong to at least one edge
of $G$ falls into connected components. These components are
called the {\em faces} of the drawing. Notice that if $G$ has an
obstacle representation with a particular placement of its vertex
set, then
(1) each obstacle must lie entirely in one face of the drawing,
and
(2) each non-edge of $G$ must be blocked by at least one of the
obstacles.\\
We start by proving a result about the {\em convex obstacle number}
(a special case of Theorem \ref{concave}), as the arguments are
simpler here. Then we tackle Theorem \ref{enumeration} using similar methods.
Following Alpert et al., we define the {\em convex
obstacle number} ${\rm obs}_c(G)$ of a
graph $G$ as the minimal number of obstacles in an obstacle
representation of $G$, in which each obstacle is convex.
\begin{claim*}\label{convex}
For every $n$, there exists a graph $G$ on $n$ vertices with
convex obstacle number
$${\rm obs}_c(G)\geq \Omega\left({n}/{\log n}\right).$$
\end{claim*}
The idea is to find a short encoding of the obstacle
representations of graphs, and to use this to give an upper bound
on the number of graphs with low obstacle number. The proof uses
the concept of order types. Two sets of points, $P_1$ and $P_2$,
in general position in the plane are said to have the same {\em
order type} if there is a one to one correspondence between them
with the property that the orientation of any triple in $P_1$ is
the same as the orientation of the corresponding triple in $P_2$.
Counting the number of different order types is a classical task.
\medskip
\noindent {\bf Theorem A.} [Goodman, Pollack~\cite{GP86}] {\em
The number of different order types of $n$ points in general
position in the plane is $2^{O(n\log n)}$. }\medskip
\noindent Observe that asymptotically the same upper bound holds for the number
of different order types of $n$ {\em labeled} points, because the
number of different permutations of $n$ points is $n!=2^{O(n\log
n)}$.
\begin{proof}[Proof of Claim~\ref{convex}]
We will give an upper bound for the number of graphs that admit a
representation with at most $h$ convex obstacles. Let us fix such
a graph $G$, together with a representation. Let $V$ be the set of
points representing the vertices, and let $O_1,\ldots, O_h$ be the
convex obstacles. For any obstacle $O_i$, rotate an oriented
tangent line $\ell$ along its boundary in the clockwise direction.
We can assume without loss of generality that $\ell$ never passes
through two points of $V$. Let us record the sequence of points
met by $\ell$. If $v\in V$ is met at the right side of $\ell$, we
add the symbol $v^+$ to the sequence, otherwise we add $v^-$. (See
Figure 1.
\input{convexObstacle.tex}
When $\ell$ returns to its initial position, we stop. The
resulting sequence consists of $2n$ characters. From this
sequence, it is easy to reconstruct which pairs of vertices are
visible in the presence of the single obstacle $O_i$. Hence,
knowing these sequences for every obstacle $O_i$, completely
determines the visibility graph $G$. The number of distinct
sequences assigned to a single obstacle is at most $(2n)!$, so the
number of graphs with convex obstacle number at most $h$ cannot
exceed $((2n)!)^h/h!<(2n)^{2hn}$. As long as this number is
smaller than $2^{n\choose 2}$, there is a graph with convex
obstacle number larger than $h$.
\end{proof}
To prove Theorem \ref{enumeration}, we will need one more result.
Given a drawing of a graph, the \emph{complexity} of a face is the number of line-segment sides bordering it. The following result was
proved by Arkin, Halperin, Kedem, Mitchell, and Naor (see Matou\v
sek, Valtr~\cite{MV97} for its sharpness).
\medskip
\noindent {\bf Theorem B.} [Arkin et al.~\cite{AHK95}] {\em
The complexity of a single face in a drawing of a graph with $n$
vertices is at most $O(n\log n)$. }
\medskip
\noindent Note that this bound does not depend on the number of
edges of the graph.
\begin{proof}[Proof of Theorem~\ref{enumeration}] First we show how to reduce part (i) of the theorem to part (ii). For each graph $G$ with $n$ vertices that admits a representation with at most $h$ obstacles, fix such a representation. Consider the visibility graph $G$ of the vertices in this representation. As explained at the beginning of this section, each obstacle belongs to a single face in this drawing. In view of Theorem~B, the complexity of every face is $O(n\log n)$. Replacing each obstacle by a slightly shrunken copy of the face containing it, we can achieve that every obstacle {\em is} a polygonal region with $O(n\log n)$ sides.
Now we prove part (ii). Notice that the order type of the sequence
$S$ starting with the vertices of $G$, followed by the vertices of
the obstacles (listed one by one, in cyclic order, and properly
separated from one another), completely determines $G$. That is, we
have a sequence of length $N$ with $N\le n + s$. According to
Theorem~A (and the comment following it), the number of different
order types with this many points is at most
$$2^{O(N\log N)}<2^{c(n+s)\log (n+s)},$$
for a suitable constant $c>0$. This is a very generous upper
bound: most of the above sequences do not correspond to any
visibility graph $G$.
\end{proof}
Following Alpert et al., we define the {\em segment obstacle number} ${\rm obs}_s(G)$ of a graph $G$ as the minimal number of obstacles in an obstacle representation of $G$, in which each obstacle is a
{\em (straight-line) segment}. If we only allow segment obstacles, we have $s=2n$, and thus Theorem \ref{enumeration} (ii) implies the following bound.
\begin{corollary*}\label{segment}
For every $n$, there exists a graph $G$ on $n$ vertices with
segment obstacle number
$${\rm obs}_s(G)\geq \Omega\left({n^2}/{\log n}\right).$$
\end{corollary*}
In general, as long as the sum of the sides of the obstacles, $s$,
satisfies $s\log s=o({n\choose 2})$, we can argue that there is a
graph that cannot be represented with such obstacles.
\section{Proof of Theorem \ref{concave}}\label{sec:concave}
Before turning to the proof, we need a simple property of obstacle representations.
\begin{lemma*}\label{lemma}
Let $k>0$ be an integer and let $G$ be a graph with $n$ vertices that has an obstacle representation with fewer than $\frac{n}{2k}$ obstacles. Then $G$ has at least $\lfloor\frac{n}{2k}\rfloor$ vertex disjoint induced subgraphs of $k$ vertices with obstacle number at most one.
\end{lemma*}
\begin{proof}
Fix an obstacle representation of $G$ with fewer than $\frac{n}{2k}$
obstacles. Suppose without loss of generality that in this representation no two vertices have the same $x$-coordinate. Using vertical lines, divide the vertices of $G$ into $\lfloor\frac{n}{k}\rfloor$ groups of size $k$ and possibly a last group that contains fewer than $k$ vertices. Let $G_1, G_2, \ldots$ denote the subgraphs of $G$ induced by these groups. Notice that if the convex hull of the vertices of $G_i$ does not entirely contain an obstacle, then ${\rm obs}(G_i)\le 1$. Therefore, the number of subgraphs $G_i$ that have $k$ points and obstacle number at most one is larger than $\lfloor\frac{n}{k}\rfloor - \frac{n}{2k}$, and the lemma is true.
\end{proof}
We prove Theorem \ref{concave} by a probabilistic argument.
\begin{proof}[Proof of Theorem \ref{concave}]
Let $G$ be a random graph on $n$ labeled vertices, whose edges are chosen independently with probability 1/2. Let $k$ be a positive integer to be specified later. According to Lemma~\ref{lemma},
$${\mbox{\rm Prob}}[{\rm obs}(G)<n/(2k)]$$
can be estimated from above by the probability that $G$ has at least
$\lfloor\frac{n}{2k}\rfloor$ vertex disjoint induced subgraphs of $k$ vertices such that each of them has obstacle number at most one. Let $p(n,k)$ denote the probability that $G$ satisfies this latter condition.
Suppose that $G$ has $\lfloor n/(2k)\rfloor$ vertex disjoint induced subgraphs $G_1, G_2,\ldots$ with $|V(G_i)|=k$ and ${\rm obs}(G_i)\le 1$. The vertices of $G_1, G_2,\ldots$ can be chosen in
at most ${n\choose k}^{\lfloor n/(2k)\rfloor}$ different ways. It follows from Theorem~\ref{enumeration}(i) that the probability that a fixed $k$-tuple of vertices in $G$ induces a subgraph with obstacle number at most one is at most
$$2^{O(k\log^2k)-{k\choose 2}}.$$
For disjoint $k$-tuples of vertices, the events that the obstacle numbers of their induced subgraphs do not exceed one are independent.
Therefore, we have
$$p(n,k)\le {n\choose k}^{\lfloor n/(2k)\rfloor}\cdot
(2^{O(k\log^2k)-{k\choose 2}})^{\lfloor n/(2k)\rfloor}\le
2^{n\log n - nk/4 + O(n\log^2k)}.$$
Setting $k=\lfloor 5\log n\rfloor)$, the right-hand side of the last inequality tends to zero. In this case, almost all graphs on $n$ vertices have obstacle number at least $\frac{n}{2k}>\frac{n}{10\log n}$, which completes the proof.
\end{proof}
\section{Proof of Theorem \ref{exactly}}\label{sec:exactly}
Alpert, Koch, and Laison \cite{AKL09} asked whether for every natural number $h$ there exists a graph whose obstacle number is exactly $h$. Here we answer this question in the affirmative.
\begin{proof}[Proof of Theorem \ref{exactly}]
Pick a graph $G$ with obstacle number $h'>h$. (The existence of
such a graph was first proved in \cite{AKL09}, but it also follows
from Theorem \ref{concave}.) Let $n$ denote the number of
vertices of $G$. Consider the complete graph $K_n$ on $V(G)$. Clearly, ${\rm obs}(K_n)=0$, and $G$ can be obtained from $K_n$
by successively deleting edges. Observe that as we delete an edge
from a graph $G'$, its obstacle number cannot increase by more
than {\em one}. Indeed, if we block the deleted edge $e$ by adding a very small obstacle that does not intersect any other edge of $G'$, we obtain a valid obstacle representation of $G'-e$. (Of course, the obstacle number of a graph can also {\em decrease} by the removal of an edge.) At the beginning of the process, $K_n$ has obstacle number {\em zero}, at the end $G$ has obstacle number $h'>h$, and whenever it increases, the increase is {\em one}. We can conclude that at some stage we obtain a graph with obstacle number precisely $h$.
\end{proof}
The same argument applies to the convex obstacle number, to the
segment obstacle number, and many similar parameters.
\subsection*{Acknowledgement.}
We are indebted to Den{\.{i}}z Sar{\i}{\"{o}}z for many valuable
discussions on the subject. A conference version containing some
results from this paper and some from \cite{PS} appeared in
\cite{MPS}.
\bibliographystyle{plain}
|
\section{Introduction}
In recent years, the experimental accessibility of the single and multi
layered graphene
samples~\cite{Novoselov,Yzhang,Bunch,Berger,Novoselovpnas}
has attracted considerable theoretical and experimental attention due
to its unusual electronic structure described by the Dirac equation,
namely electrons in monolayer graphene have linear dispersion thus
behave like massless Dirac fermions at the corners of the Brillouin
zone (BZ)~\cite{Wallace}. In the presence of magnetic field
perpendicular to the graphene plane, the system shows anomalous
integer quantum hall
effect~\cite{Novoselovnature,Yzhangnature,Novoselovscience,Abanin} (IQHE)
which is different from that of the conventional two-dimensional
electron system in semiconductor heterostructures.
The Hall conductivity in the IQHE of graphene shows half integer
plateaus instead of integer ones in the IQHE of a normal
semiconductor due to the different degeneracy at $N=0$ Landau level.
Edge states in graphene have been the focus of much theoretical study
because of the important role they play in transport.~\cite{castrormp}
It is well known that there are two basic types of edges in graphene,
namely, the armchair and zigzag edges. Some theoretical
work~\cite{stein1987,Tanaka,Fujita,Fujitapsj,Nakada,zhenghx,brey,
Peres, kohmoto, Ryu, sasaki, sasaki07, castro, Niimi, kobayashi} on
the electronic structure
of finite-sized systems, either as molecules or as one-dimensional
systems, has shown that graphene with zigzag edge has localized edge
states at the Fermi energy, but those with armchair edges do not
have such state. Therefore, the zigzag edge states is essential to the
transport properties in graphene since the localized edge state has
contribution to the conductivity. However, most of the work were focused on
the uniform monolayer and bilayer graphene. On the other hand the
hybrid edge structure composed of partial monolayer and partial
bilayer graphene , which is quit general in reality, has received
not so much attention. Experimentally,~\cite{yliu} the anomalous
quantum oscillations in magnetoconductance was observed due to the
peculiar physics along the interface. Some of the transport properties
and the presence of interface Landau levels was explored within an
effective-mass approximation.~\cite{Koshino, Takanishi}
In this work\cite{wenxin}, we study the electronic properties of the
hybrid interface systematically via both tight binding model and
its effective theory in the continuum limit - the Dirac
equation. The
edge states in graphene can be studied experimentally by using a local
probe such as scanning tunneling microscopy
(STM). The STM experiments measure the differential conductance which
is proportional to the density of states. Thus we study the local
density of states (LDOS) in different hybridized zigzag edge graphene. It is defined as:
\beq
N(r, eV) = |\Psi_\alpha(r)|^2 \delta(eV - E_\alpha),
\eeq
where $\Psi_\alpha(r)$ is the eigenfunction with energy
$E_\alpha$. The LDOS shows the strength of the local electronic
density which is related to the strength of the signal in STM
experimental data. Therefore, the LDOS can show us the signals of the
different edge state as in our previous work.~\cite{hu} In this
paper, our study of the step edge graphene shows that there are two
different zigzag step edge, and there always exists
zero energy states localized near the zigzag step edge but the
distribution of LDOS of these edge states strongly depends on the
details of how the edge stacks together. We find the energy dispersion around the Dirac cones also
presents different characters for different edge arrangements either within or without a magnetic field.
This paper is arranged as follows: In section II, we set up the model
Hamiltonian in different geometries. The zero energy solutions in
zero field and in magnetic field are obtained by solving the Dirac
equations and numerical diagonalization is implemented in a finite
system in section III and IV, and some discussions and conclusions are
in the section V.
\section{Model and Different Geometries}
We consider the bilayer graphene with $AB$ stacking, as shown in
Fig. \ref{fig:ab_stacking}; its tight-binding Hamiltonian can be written as:
\beqarr
H = &-&t \sum_{i=1}^2 \sum_{m,n} a^\dagger_{i;m,n} (b_{i;m,n} +
b_{i;m-1,n} + b_{i;m,n-1}) \nonumber\\
&-& t_{\perp} \sum_{m,n} a^\dagger_{2;m,n}
b_{1;m,n} + h.c.,
\eeqarr
where $a_{i;m,n}$ ($b_{i;m,n}$) is the annihilation operator at
position ($m,\ n$) in sublattice $A_i$ ($B_i$), and $i=1,\ 2$, indicating
the two layers.
\begin{figure}
\includegraphics[width=4cm]{./leb.eps}
\includegraphics[width=4cm]{./heb.eps}
\caption{\label{fig:ab_stacking}(Color online)The schematic pictures for two kinds
of monolayer-bilayer interface. The atoms of the extended bottom
layer (layer 1) are indicated by black dots while the green dots
represent the top layer which terminates at the interface. In the left plot, the top layer
is ended with $B_2$ sites ({\it l.e.b}) while on the right it is
ended with $A_2$ sites ({\it h.e.b}).}
\end{figure}
The first term is the Hamiltonian within each layer,
and the second term describes the interlayer coupling
in which we only consider the hopping between the two atoms stacked
right on top of each other. Let us label the bottom
(extended) layer as layer $1$, the half plane upper layer as layer
$2$. In this work we will only consider the $A-B$ (Bernal)
stacking. One can expand the effective Hamiltonian near the two Dirac
points $K$ and $K'$ which are time reversal symmetric partners. In
momentum space, the hamiltonian near $K$ can be written as:
\beq
H = \sum_k \Psi^\dagger_k \cdot H_k \cdot \Psi_k,
\eeq
where
\beq \label{hambilayer}
H_k = \begin{pmatrix} 0 & v_F k & 0 & 0 \\ v_F k^* & 0 & t_{\perp} &
0 \\ 0 & t_\perp & 0 & v_F k \\ 0 & 0 & v_F k^* & 0 \end{pmatrix}
= v_F \begin{pmatrix} 0 & k & 0 & 0 \\ k^* & 0 & \gamma &
0 \\ 0 & \gamma & 0 & k \\ 0 & 0 & k^* & 0 \end{pmatrix},
\eeq
in which $k = k_x + i k_y$, $\gamma = t_\perp / v_F$, and $\Psi_k =
(a_{1;k},b_{1;k},a_{2;k},b_{2;k})$. For the other Dirac point, as
stated before, $H_{K'} = H_{K}^*$. Here we only consider the zigzag type interface(or edge) to explore the localized edge
state. Without magnetic field and the interface, it is sufficient
to discuss just one Dirac cone in the continuum model due to the
symmetry. But with the interface breaking the inversion symmetry and
the magnetic field breaking the time reversal symmetry, the two Dirac
points are not equal to each other, and both must be studied.
According the lattice orientation we
adopt which is shown in Fig.~\ref{fig:ab_stacking}, the zigzag
interface is along the $x$ direction. For simplicity, we consider an
infinite stripe along the $x$ direction, therefore the system has
translational symmetry along the $x$ direction thus $k_x$ remains a good
quantum number. We then do Fourier transformation in the $x$ direction
and reduces the 2D problem to 1D. There are actually TWO distinct
geometries which are physically different. i) The outmost sites
of the upper layer are the $B_2$ sites which do not stack directly on
the lower layer atoms as showed in
Fig.~\ref{fig:ab_stacking}(left). The $B_2$ sites are the low energy
degrees of freedom (along with $A_1$) which are kept if one further
considers an $2\times2$ effective theory on energy scale $\epsilon \ll
t_{\bot}$. We label it the low energy sites boundary ({\it l.e.b.});
ii) the $A_2$ sites are the outmost sites on the upper layer as showed
in Fig.~\ref{fig:ab_stacking}(right). The $A_2$ sites, together with
the $B_1$ sites which they stack right on top of, form the dimer
sites. In the $2\times2$ low energy effective theory the wavefunction
have almost zero weight on those dimes sites when $\epsilon \ll
t_{\bot}$. As a result these dimers are ignored in this limit. In
other words, they are occupied considerably only at high energy
(comparing to $t_{\bot}$). Therefore, we shall
address such interface as the high energy sites boundary ({\it h.e.b.}).
\section{Interface properties in Zero Field}
For a semi-infinite sheet, it is well known that the existence of the zero energy
edge modes in both monolayer and bilayer graphene.
Presumably, such modes are also expected at the interface between
them. Let us consider the following geometry: a
half plane of monolayer graphene and a half plane of bilayer graphene
joined along the zigzag edge; and look for solution(s) with zero
eigenenergy by using the Dirac equation.
Before we present the results, let us discuss the boundary conditions
at the interface firstly as in Ref.~\onlinecite{Nilsson}. Let
\beq \Psi_{\text{mono}}(x,y) = \binom{\psi_A(x,y)}{\psi_B(x,y)} \eeq and
\beq \Psi_{\text{bi}}(x,y) = \begin{pmatrix} \psi_{A1}(x,y) \\ \psi_{B1}(x,y)
\\ \psi_{A2}(x,y) \\ \psi_{B2}(x,y)\end{pmatrix} \eeq
be the wavefunctions at Dirac point in the monolayer and bilayer respectively. Suppose the
interface locate at $y = 0$, the boundary condition for monolayer is then
straightforward: both components of the wavefunction must be
continuous. For the upper layer, it terminates at $y=0$ and therefore
satisfies the open boundary condition. Note that the term `terminates'
indicates the last row of lattice sites are the high/low energy sites,
however, the boundary condition is not the wavefunctions being zero
on these sites. They should be the wavefunctions on sites one unit
cell `outside' the boundary being zero. To sum up, we have:
\beqarr
&&\psi_A(x,0) = \psi_{A1}(x,0),\qquad \psi_B(x,0) = \psi_{B1}(x,0), \nonumber\\
&&\begin{cases} \psi_{A2}(x,0) = 0 & l.e.b. \\ \psi_{B2}(x,0) = 0 &
h.e.b. \end{cases}.
\eeqarr
In the monolayer region with zigzag interface,
we do the substitution
$k\rightarrow k_x+\partial _y$ in the Dirac hamiltonian $H_{\text{mono}} =v_F \left( \begin{array}{cc}
0 & k \\
k^* & 0
\end{array}
\right)$.~\cite{castrormp} The zero energy solution is
\beq
\left\{
\begin{array}{c}
\phi_A(y)=e^{y k_x} C_2 \\
\phi_B(y)=e^{-y k_x} C_1
\end{array}
\right.
\eeq
In analogy, we do the same substitution in the bilayer Hamiltonian (Eq. ~\ref{hambilayer}) and obtain its zero energy solution:
\beq
\begin{cases}
\phi_{B1}(y)=e^{-y k_x} A_1 \\
\phi_{B2}(y)=-e^{-y k_x} y \gamma A_1+e^{-y k_x} A_2 \\
\phi_{A1}(y)=e^{y k_x} A_3 + e^{y k_x} y \gamma A_4\\
\phi_{A2}(y)=e^{y k_x} A_4
\end{cases}.
\eeq
Applying the requirement that the wavefunctions remain finite at
infinity, for the {\it l.e.b.} interface (so
$\phi_{A2}(y=0)=0$), one easily find that nonzero solutions only exist
for $k_x > 0$. The solution is
\beq
C_1 = C_2 = A_1 = A_3 = A_4 = 0, \qquad A_2 = \text{const}.
\eeq
For the {\it h.e.b.} interface, one finds that for $k_x > 0$
\beq
C_1 = C_2 = A_1 = A_2 = A_3 = 0, \qquad A_4 = \text{const}.
\eeq
For the other Dirac cone, the solutions remain the same but only exist
for $k_x < 0$. The only nonzero constant is to be determined by
normalization. Both results are in agreement with
tight-binding analysis\cite{castro-2008-84}.
Even though the zero energy states exist for both types of interface,
there is an important difference between them. In the $l.e.b.$ case, the
wavefunction is only none-zero in the upper layer, which is trivial
as such mode is expected when a graphene sheet is terminated at a
zigzag edge. However, the $h.e.b.$ is less trivial. The wavefunction
also lives on the extended layer at the interface where no cut is
present. We interpret this as, when the dimer sites are the boundary,
the interlayer coupling $t_\bot$ imposes an energy cost for electrons
to go through the interface in the extended layer which can be
considered as an effective potential barrier. The potential barrier can localize the electron states along the interface.
\begin{figure}
\includegraphics[width=8cm]{dispersion.eps}
\caption{\label{fig:disp_step}The dispersion around Dirac cone for
the step graphene with (a)l.e.b and (b)h.e.b respectively.}
\end{figure}
We numerically diagonalize a system with a finite width up
to 600 unit cells in the $y$ direction. The intralayer hopping strength
$t$ is set to identity and the interlayer hopping strength
$t_\perp = 0.2t$. The dispersion relation is shown in Fig.~\ref{fig:disp_step}.
Compare the dispersions in different geometries, the {\it h.e.b.} edge
has an obviously stronger level anti-crossing feature than the other. The
reason is that for {\it l.e.b.} edge, the zero energy edge state just
locates on the bilayer graphene which has quadratic dispersion and
has nothing to do with the monolayer part with a linear
dispersion. However, in the case of {\it h.e.b.}, the zero energy edge
state also has component on the monolayer graphene at the interface,
the linear dispersion part for monolayer should also connect to the edge state which
induces more energy level anticrossing around the Dirac points.
Fig.\ref{nofieldldos} shows the LDOS for the two kinds of step
graphene. We label the coorindinate of the extended layer by $d \in [0,600)$ and the
$d \in [900, 1200)$ for upper layer. Therefore, the interface locates at
$d=300$ and $d=900$ for layer 1 and 2 respectively.
One can notice that the localized edge state shows as a peak at zero energy
along the interface. The prominent difference is that the peak just appears on the top layer of bilayer part in
the {\it l.e.b.} case; but for the {\it h.e.b.} case, the zero energy
peak appears on both two layers although still
locates at the bilayer side. The numerical results are in
agreement with the analysis by the Dirac equations
and we conclude that the two kinds of edge arrangements should have
different consequences in experiments since the different distributions of the zero mode. Here we notice that the other
peaks in LDOS at the end of the finite system is the signal of the
general monolayer or bilayer zigzag edge graphene as discussed in many others
work.~\cite{stein1987,Tanaka,Fujita,Fujitapsj,Nakada,zhenghx,brey, Peres, kohmoto, Ryu, sasaki, sasaki07, castro, Niimi, kobayashi}
\begin{figure}
\includegraphics[width=4.cm]{./LDOSnofieldA.eps}
\includegraphics[width=4.cm]{./LDOSnofieldB.eps}
\caption{\label{nofieldldos}(color online) The LDOS of the
step bilayer graphene with l.e.b.(left) and h.e.b.(right) respectively.
The width of of monolayer $L = 600$. In plot, the region $d \in
[0:600)$ stands for the lower extended layer and $d \in [600,1200)$ is
the upper layer. }
\end{figure}
\section{Interface properties in Magnetic Fields}
In the presence of magnetic field, by minimal coupling the Dirac
equation should be modified by doing the substitution ${\bf k}
\rightarrow {\bf k} + \frac{e {\bf A}}{c}$. Assume the magnetic field
is along the direction of perpendicular to the plane $B = B
\hat{z},\ B>0$. We adopt the Landau gauge here $\vec{A} = (A_x,
A_y) = (- y B, 0)$ due to the translational invariant along $x$
direction. Thus the Hamiltonian of monolayer becomes:
\beq
H_{\text{mono}} = \begin{pmatrix} 0 & k_x + \partial_y -
\frac{eBy}{c}\\ k_x - \partial_y - \frac{eBy}{c} &
0\end{pmatrix}.
\eeq
and its zero energy solution is
\beq
\left\{
\begin{array}{c}
\phi_A(y)=C_1 e^{k_x y - \frac{eB}{2c}y^2} \\
\phi_B(y)=C_2 e^{-k_x y+ \frac{eB}{2c}y^2}
\end{array}
\right.
\eeq
Similarly, the zero energy solution for bilayer graphene becomes:
\beq
\begin{cases}
\phi_{A1}(y)= (A_3 \gamma y + A_2 )e^{k_x y - \frac{eB}{2c}y^2 }\\
\phi_{B1}(y)= A_1 e^{- k_x y + \frac{eB}{2c} y^2}\\
\phi_{A2}(y)=A_3 e^{k_x y-\frac{eB}{2c} y^2}\\
\phi_{B2}(y)=(- A_1 \gamma y + A_4 )e^{- k_x y + \frac{eB}{2c}y^2 }
\end{cases}.
\eeq
Apply the same boundary conditions as in the zero field case, we will
have the solutions. For $l.e.b.$, one gets
\beq
C_2 = A_1 = A_3 = A_4 =0, \qquad C_1 = A_2 = \text{const}.
\eeq
For $h.e.b.$ the solution is
\beq
C_2 = A_1 = A_4 =0, C_1 = A_2 = \text{const}_1, A_3 = \text{const}_2.
\eeq
The constants are to be determined by normalization conditions. One
immediately notices that the wavefunction lives only on the $A$ sublattice.
We should note that on the other Dirac point, the solutions remain the
same form, but resides on the $B$ sublattice. Another important
feature of the solution is that for $h.e.b.$ we actually have TWO
independent solutions here.
\subsection{Dispersion Relation}
The general problem in magnetic field can be written as
\beq
\frac{1}{\sqrt{2}}\left(
\begin{array}{cc}
0 & \partial _{\text{\textit{$\xi $}}}+\xi \\
-\partial _{\text{\textit{$\xi $}}}+\xi & 0
\end{array} \right) \Psi_{\text{mono}} =\epsilon \Psi_{\text{mono}},
\eeq
for the monolayer
while that for the bilayer is
\beq
\frac{1}{\sqrt{2}} \left( \begin{array}{cccc}
0 & \partial _{\xi}+\xi & 0 & 0 \\
-\partial _{\xi}+\xi & 0 & \tilde{\gamma} & 0
\\
0 & \tilde{\gamma} & 0 & \partial _{\xi}+\xi \\
0 & 0 & -\partial _{\xi}+\xi & 0 \\
\end{array} \right) \Psi_{\text{bi}} =\epsilon \Psi_{\text{bi}}
\eeq
where $
\xi =\frac{y}{l_B}-l_Bk_x
\text{, }l_B=\sqrt{\frac{c}{e B}}, \epsilon =\frac{E}{\omega _c},
\omega _c=\sqrt{2}\frac{v_F}{l_B}$, $\tilde{\gamma} =
\frac{\gamma}{v_F \omega_c}$.
It is known that in the bulk, the solution to the monolayer is
\beq
\Psi_{\text{mono}} = \begin{pmatrix} \frac{1}{\Gamma(\epsilon^2 )}
D_{\epsilon^2-1} (\sqrt{2} \xi)
\\ \pm \frac{1}{\Gamma(\epsilon^2 + 1)} D_{\epsilon^2} (\sqrt{2}
\xi) \end{pmatrix} = \begin{pmatrix} \psi_{\epsilon^2-1}(\xi)
\\ \pm \psi_{\epsilon^2}(\xi) \end{pmatrix} ,
\eeq
where the $\epsilon = \pm \sqrt{N}$, $N=0, 1, 2 \dots$, and $D_\nu$'s
are the parabolic cylinder functions, which combined with the factor
$1/\Gamma(v+1)$ give us the eigen wavefunctions of a harmonic
oscillator $\psi_{\nu}(\xi)$. The bulk solution to the bilayer
Hamiltonian can be written in a similar way:
\beq
\Phi _{\text{bi}} =\left(
\begin{array}{c}
\frac{\epsilon \left(\epsilon ^2-(j+1)-\tilde{\gamma}
^2\right)}{\tilde{\gamma} \sqrt{j (j+1)}} \psi _{j-1} \\
\frac{\epsilon ^2-(j+1)}{\tilde{\gamma} \sqrt{j+1}} \psi _j \\
\frac{\epsilon }{\sqrt{j+1}} \psi _j \\
\psi _{j+1}
\end{array}
\right),
\eeq
where $\epsilon =\pm \sqrt{\frac{1+\tilde{\gamma} ^2+2 j\pm
\sqrt{\left(1+\tilde{\gamma} ^2\right)^2+4\tilde{\gamma} ^2 j}}{2}}$ ($j = 0, 1,
2, \dots$) is the eigen energy.
However, for the interface problem, we need solutions on the
half-plane. In this case, we let the bilayer live on the $y>0$ side,
so the monolayer is on the $y<0$ part. Therefore, for the bilayers,
the solutions on $(0, \infty)$ take on the same form as in the bulk,
but $j$'s are no longer required to be integers. Instead, we now
should replace $j$'s by
\beq
j_{1,2} = \epsilon ^2-\frac{1}{2}\pm \sqrt{\epsilon ^2 \tilde{\gamma}
^2+\frac{1}{4}},
\eeq
and $\epsilon$ now varies continuously. For the monolayer, the
solution on $(-\infty, 0)$ can be chosen as
\beq
\Psi_{\text{mono}} = \begin{pmatrix} \psi_{\epsilon^2-1}(-\xi)
\\ \mp \psi_{\epsilon^2}(-\xi) \end{pmatrix} .
\eeq
With the above solutions, and combined with the boundary conditions
that are already discussed in the zero field cases, we obtain the
transcendental equations that dictates the dispersion relations. For
the $l.e.b.$ interface,
\beqarr
&&\hspace{-0.5cm}(\epsilon ^2-(j_1+1)) D_{j_1-1}(-\sqrt{2}k_x) (D_{\epsilon ^2}(\sqrt{2} k_x)
D_{j_2}(-\sqrt{2} k_x) \nonumber\\
&& + D_{\epsilon ^2-1}(\sqrt{2}k_x) D_{j_2+1}(-\sqrt{2} k_x)) \nonumber\\
&&\hspace{-0.5cm}= (\epsilon^2-(j_2+1)) D_{j_2-1}(-\sqrt{2} k_x)
(D_{\epsilon ^2}(\sqrt{2} k_x) D_{j_1}(-\sqrt{2}k_x) \nonumber\\
&& + D_{\epsilon ^2-1}(\sqrt{2} k_x)D_{j_1+1}(-\sqrt{2} k_x)),
\eeqarr
where we have set $l_B = 1$. For the $h.e.b.$, the dispersion equation
is
\beqarr
&&\hspace{-0.5cm} (\epsilon ^2-(j_1+1)-\gamma ^2)
D_{j_1}(-\sqrt{2} k_x) (D_{\epsilon ^2}(\sqrt{2}
k_x) D_{j_2}(-\sqrt{2} k_x) \nonumber\\
&& +D_{\epsilon ^2-1}(\sqrt{2} k_x) D_{j_2+1}(-\sqrt{2} k_x)) \nonumber\\
&&\hspace{-0.5cm} =(\epsilon ^2-(j_2+1)-\gamma ^2)D_{j_2}(-\sqrt{2} k_x) (D_{\epsilon ^2}(\sqrt{2}
k_x) D_{j_1}(-\sqrt{2} k_x) \nonumber\\
&&+D_{\epsilon ^2-1}(\sqrt{2} k_x) D_{j_1+1}(-\sqrt{2} k_x))
\eeqarr
However, one must note that in the presence of magnetic field, the
time reversal symmetry is broken, therefore, the other Dirac cone, the
time reversal partner, no long behaves the same way. So the dispersion
relation must be calculated separately. By the same approach, one can
get, for the $l.e.b.$,
\beqarr
&&\hspace{-0.5cm}(\epsilon ^2-j_1) D_{j_1} (-\sqrt{2} k_x) (j_2
D_{\epsilon ^2} (\sqrt{2} k_x)D_{j_{2-1}} (-\sqrt{2} k_x)\nonumber\\
&&+\epsilon ^2 D_{\epsilon ^2-1} (\sqrt{2} k_x) D_{j_2} (-\sqrt{2} k_x)) \nonumber\\
&&\hspace{-0.5cm} = (\epsilon ^2-j_2) D_{j_2} (-\sqrt{2} k_x) (j_1
D_{\epsilon ^2} (\sqrt{2} k_x)D_{j_1-1} (-\sqrt{2} k_x)\nonumber\\
&&+\epsilon ^2 D_{\epsilon ^2-1} (\sqrt{2} k_x) D_{j_1} (-\sqrt{2} k_x));
\eeqarr
for the $h.e.b.$,
\beqarr
&&\hspace{-0.5cm}(j_2+1)(\epsilon ^2-j_1-\gamma ^2)D_{j_1} (-\sqrt{2}
k_x) (j_2 D_{\epsilon ^2} (\sqrt{2} k_x) \nonumber\\
&&D_{j_{2-1}} (-\sqrt{2}k_x) +\epsilon ^2 D_{\epsilon ^2-1} (\sqrt{2} k_x)D_{j_2} (-\sqrt{2} k_x)) \nonumber\\
&&\hspace{-0.5cm}=(j_1+1) (\epsilon ^2-j_2-\gamma ^2) D_{j_2}
(-\sqrt{2}k_x) (j_1 D_{\epsilon ^2} (\sqrt{2} k_x)\nonumber\\
&&\hspace{-0.5cm}D_{j_1-1}(-\sqrt{2} k_x)+\epsilon ^2 D_{\epsilon^2-1}
(\sqrt{2} k_x) D_{j_1} (-\sqrt{2} k_x)).
\eeqarr
\begin{figure}
\includegraphics[width=7cm]{./disp_Bfield_K1.eps}
\includegraphics[width=7cm]{./disp_Bfield_K2.eps}
\caption{\label{disp:Bfield}(color online) The dispersion around Dirac cones in the
presence of magnetic field. The upper two figures are for the {\it
l.e.b.} and {\it h.e.b.} interfaces around one Dirac cone
respectively, and the lower two figures are that for the other
Dirac cone.}
\end{figure}
It is easy to see from the dispersion equation that when $\abs{k_x}$
is very large, the eigen-energy should restore to either the monolayer
value or the bilayer value as the parabolic cylinder functions $D_\nu$ only
converge to zero for integer $\nu$ at infinity. But what interests us
is how these bulk Landau levels connect with each other when crossing
the interface. To study that,we solve the
above equations for $\epsilon$ and $k_x$ near the interface $k_x =
0$ for a finite size system as shown in Fig.~\ref{disp:Bfield}. We see
that at the two Dirac cones, the Landau levels match in different ways.
\begin{figure}
\includegraphics[width=6cm, height=5cm]{./FieldAenergy.eps}
\includegraphics[width=7cm]{./FieldALDOS.eps}
\caption{\label{Blowenergy}The energy spectrum and LDOS of the
step bilayer graphene with {\it l.e.b.} in a
magnetic field. The strength of the magnetic field is expressed as
magnetic flux $\phi$ in each unit cell. Here we set the magnetic flux per each unit cell
$\phi=\phi_0/1315$ ($\phi_0$ is the magnetic flux quanta) which corresponds to $B \sim 60 T$.}
\end{figure}
\begin{figure}
\includegraphics[width=6cm, height=5cm]{./FieldBenergy.eps}
\includegraphics[width=7cm]{./FieldBLDOS.eps}
\caption{\label{Bhighenergy}The energy spectrum and LDOS of the
step bilayer graphene with {\it h.e.b.} within magnetic field. The
parameters are the same as in Fig.~\ref{Blowenergy}}
\end{figure}
Furthermore, we present the numerical result for directly diagonalizing a ribbon with width $L = 600$.
Our diagonalization results consistent with the analytic analysis also.
Fig.~\ref{Blowenergy} and Fig.~\ref{Bhighenergy} shows the energy
spectrum and LDOS for step bilayer graphene in a magnetic field
with {\it l.e.b.} and {\it h.e.b.} respectively. In magnetic
field the energy spectrum is split into a set of Landau levels(LLs)
and the Landau level splitting in bilayer is smaller than that of
monolayer graphene since the energy of LLs in bilayer system satisfies $E
\propto \sqrt{n(n-1)}$ and $E \propto \sqrt{n}$ in monolayer.
Therefore, LLs are not matched at the interface.
The major difference between the two sets of dispersion relation is
that the Landau levels in the two distinct regions connect in two
different ways. The dispersion at the interface with {\it l.e.b.} is more flat than that with
{\it h.e.b.}. These interface Landau levels was discussed by M. Koshino et.al.
within an effective-mass approximation~\cite{Koshino}. From the plot
of LDOS, the same as in the case of zero field, according to the different distribution of
the edge state, the presence of the peak at $E=0$
along the interface only on the upper layer for {\it
l.e.b.} and on both two layers for {\it h.e.b.}.
We also use the Dirac equation result to study how the dispersive
Landau levels near the interface changes when the magnetic field
varies. If we measure the energy in unit of $\hbar \omega_c$, the
monolayer bulk Landau levels remain unchanged as the magnetic field is
being tuned; but for the bilayer, the effective interlayer coupling
$\tilde{\gamma} \sim 1/\omega_c$ so the bulk Landau levels vary with
the magnetic field. In strong field $\tilde{\gamma} \sim 0$, so the
two layers behave as if they were decoupled, and the Landau levels'
energy becomes just like the monolayer's but with a two-fold
degeneracy. In weak field limit, $\tilde{\gamma} \sim \infty$, in this
case one gets $\epsilon \simeq \sqrt{n(n-1)}/\tilde{\gamma}$. The
dispersions for $\tilde{\gamma} \in {0.2, 1,2,5}$ of both $l.e.b.$ and
$h.e.b.$ at one Dirac point are shown in
Fig.~\ref{fig:disp_g1} and Fig.~\ref{fig:disp_g2} respectively. Here we only plot the
Landau levels up to $n=6$. The evolution of dispersion on the other Dirac point have the similar
behavior except different energy level connection as shown in Fig.~\ref{disp:Bfield}.
As the strength of the magnetic field is increased, the bilayer
LLs become more denser which induces more mediate states around the interface. The dispersion of the
$l.e.b.$ interface is always more flat than that of $h.e.b.$ interface while tuning the magnetic field.
\begin{figure}
\includegraphics[width=8cm]{gLK1.eps}\\
\caption{The dispersion around Dirac cones $K$ of $l.e.b.$ type
interface in different magnetic field with $\tilde{\gamma} =
0.2,1,2,5$ respectively.}\label{fig:disp_g1}
\end{figure}
\begin{figure}
\includegraphics[width=8cm]{gHK1.eps}\\
\caption{The dispersion around Dirac cones $K$ of $h.e.b.$ type
interface in different magnetic field with $\tilde{\gamma} =
0.2,1,2,5$ respectively.}\label{fig:disp_g2}
\end{figure}
\section{Discussion and Conclusion}
In this paper, we discuss the edge state at the interface between
monolayer and bilayer graphene. Physically, there are two types
of interface structures named by the $l.e.b.$ and $h.e.b.$ due to the different terminations at the interface.
By studying the system with both effective theory and numerical calculation, we
find that with or without magnetic field, zero energy edge states
exist for both types of interface, however, the LDOS shows different
features. For $l.e.b.$ the zero energy edge states only live in the
upper layer which terminates at the interface while the $h.e.b.$
induces an enhanced LDOS in the bottom layer. Another major difference
between the two different geometries is that the dispersion of
$h.e.b.$ system shows a stronger anticrossing feature in the zero field. Both
differences can be interpreted as the following. When
an electron goes through the $h.e.b.$ interface, it can choose to lower
its energy through the interlayer coupling. This can be viewed as the
$h.e.b.$ imposes an effective potential barrier at interface in the
extended layer. Since the localized zero energy edge states locates on both
two layers, the monolayer dispersion which is linear should smoothly connect
to the edge state in the $h.e.b.$ case. This can explain the strong energy
level anticrossing in the $h.e.b.$ interface. Similar differences between $l.e.b.$ and $h.e.b.$ interfaces are also
discussed for transimission coefficients across the
interface\cite{Nilsson,Takanishi,wenxin1}. For the $h.e.b.$ the
transmission probability is reduced significantly for incoming
electrons with energy $E \sim t_{\bot}$ which can be interpreted in a
similar way.
We also see that in presence of magnetic field the dispersion of Landau levels
continueously goes through the interface in different manners at the two Dirac
cones. At Dirac cone $K(K')$, the zero energy Landau levels in the bilayer
region split, one branch rises up and becomes the $n=1$ Landau levels
in the monolayer while the other branch remains as zero energy
states, and the other higher Landau levels continues from bilayer to
monolayer accordingly, i.e. $n=1 \big\vert_{\text{bi}}\rightarrow n
=2\big\vert_{\text{mono}}$ $\dots$. However, near the other Dirac
point $K'(K)$, the zero energy state remains intact, so the other higher Landau
levels connect as $n=1 \big\vert_{\text{bi}}\rightarrow n
=1\big\vert_{\text{mono}}$ $\dots$. We also see an effect of the
effective potential barrier imposed by the $h.e.b.$ in the dispersive
Landau levels near the interface. For a certain range of field
strength the dispersion develops a local maximum for $h.e.b.$ while
the $l.e.b.$ only develops a plateau feature.
In summary, the physical properties of hybrid graphene systems are
mostly dominated by both that of monolayer and bilayer. The Landau levels
remain the same away from the interface. But the existence of
dispersive Landau levels near the interface could be related to the
unexpected feature other than that of the monolayer and bilayer
graphene in the magneto transport experiment\cite{yliu}. Further study of the
hybrid structures for a more realistic setup, like including edge
disorder, gate voltage, etc., are needed to undertand the experimental
data.
\section{acknowlegement}
We wish to thank K. Yang and Y. Barlas for very helpful discussions.
This work is supported by Fundamental Research Funds for the Central Universities CDJRC10300007 (Z. X. H) and NSF under Grant No. DMR-1004545 (W.X.D.).
|
\section{Introduction}
Our research group has a particular interest in using open clusters in
our studies of the spiral structure of our Galaxy. Some interesting
results were obtained with data of these objects that are compiled in
our catalog (\cite{Dias2002}): the rotation speed of the spiral
pattern as $25 \pm 1$ $km^{-1} s^{-1} kpc^{-1}$ and the co-rotation
radius at $1.06 \pm 0.08$ $\times$ the solar Galactocentric distance
(\cite{Dias2005}) and \cite{Lepine2008} determined the epicyclic
frequency to be $43 \pm 5$ $km^{-1} s^{-1} kpc^{-1}$ at the solar
radius.
In this kind of study and in many others that involve open clusters it
is crucial to accurately determine the fundamental parameters
(distance, reddening, and age) of the objects. Traditionally this is
done using the color-magnitude diagrams via the main-sequence fitting
(MSF) technique, which in the majority of cases is fit to the data by
visual inspection. The work of \cite{PN06} (hereafter PN06) indicates
that the usual method applied up to now can determine distances of
open clusters with absolute errors of less than 20$\%$ (see all
details in the cited paper).
From the observational point of view, the determination of fundamental
parameters of open clusters requires accurate photometry (errors lower
than 0.03 in magnitude and lower than 0.05 in color index), deep
photometry (mainly for objects not previously studied) and also seems
to require data in the U-filter. The U-filter in principle allows a
better determination of the color excess and therefore leads to an
accurate determination of the distance modulus via MSF. The
determination of the color excess toward the cluster is performed
based on the observed photometric data, typically by visual
main-sequence fitting of the (B-V) vs. (U-B) diagram. Results on
determining the color excess without the use of the U-filter are
unreliable if done in the usual manner by visual fits because of the
near linearity of the color-color diagrams and the subjectivity of the
analysis.
We have addressed the subjectivity of the isochrone fits in
\cite{MDC10} (hereafter MDC10), where we developed a global
optimization tool using the cross-entropy optimization algorithm (CE)
to automatically fit open cluster UBV photometric data with well
defined criteria (see paper for details). We showed that the method
was robust and the results obtained for 10 open clusters agreed well
with previous studies found in the literature. Although the work used
only UBV photometry and the usual sequence to fit isochrones
(determine E(B-V) and then distance and age), the results obtained
with the global optimization indicated that it might be possible to
address the color excess determination without the use of the
U-filter. The advantages of such a procedure are clear, given that the
U-filter is possibly one of the most problematic ones to get good data
for, which typically requires longer exposure times.
In this work we investigate the possibility of determining
open cluster fundamental parameters from BVRI photometry alone. In
Sec. 2 we briefly describe the global optimization method developed in
MDC10 and the adaptations made for the present problem. In
Sec. 3 we describe benchmark tests using a set of synthetic open
clusters generated with known parameters. In Sec. 4 we apply the
method to the well studied open clusters Trumpler 1 and and Melotte
105 and discuss the results. In Sec. 5 we present our conclusions.
\section{The cross-entropy optimization method}
The CE procedure provides a simple adaptive way of estimating the
optimal model parameters. Basically, the CE method involves an
iterative procedure where each iteration consists of:
\begin{enumerate}
\item Random generation of the initial parameter sample, respecting
pre-defined criteria;
\item Selection of the best candidates based on some mathematical
criteria;
\item Random generation of updated parameter samples from the previous
best candidates to be evaluated in the next iteration;
\item Optimization process repeats steps (ii) and (iii) until a
pre-specified stopping criterion is fulfilled.
\end{enumerate}
In MDC10 we fully describe the algorithm and the fitting method
applied to open clusters and we refer the reader to that work and
references therein for more details. In short, the algorithm generates
a set of possible isochrone solutions (simulated open clusters) given
a pre-defined initial mass function, binary fraction, and
metallicity. Each generated solution is then compared to the observed
data through an objective function. The best solutions are selected, a
new set of solutions is generated from them and the process iterated
until convergence. In the present work we maintain the complete
structure of the previous algorithm discussed in detail in MDC10,
changing only the objective function to use the BVRI data instead of
only UBV.
In the case of isochrone fitting, the objective function used is the
likelihood of the data (in the present case BVRI photometry) for a
given model isochrone, which is obtained from the product of
probabilities of the $m$ stars observed and is given by:
\begin{eqnarray}
P(BVRI|{ISO}_{mod})= \sum_{m} \frac{1}{\sqrt{2\pi}\sigma_{B}\sigma_{V}\sigma_{R}\sigma_{I}} \times \nonumber\\
EXP-\frac{(B-B_{ISO})^2} {2\sigma_{B}} \times\nonumber\\
EXP-\left ( \frac{V-V_{ISO}} {2\sigma_{V}} \right )^2 \times \nonumber\\
EXP-\left ( \frac{R-R_{ISO}} {2\sigma_{R}} \right )^2 \times \nonumber\\
EXP-\left ( \frac{I-I_{ISO}} {2\sigma_{I}} \right )^2,
\end{eqnarray}
where B, V, R, and I are the observed magnitudes and $\sigma_{B}$,
$\sigma_{V}$, $\sigma_{R}$, and $\sigma_{I}$ are the photometric
uncertainties in each filter. Here ${ISO}_{mod}$ designates the model
isochrone for a set of parameters (E(B-V), distance and age) and
$B_{ISO}$, $V_{ISO}$, $R_{ISO}$, and $I_{ISO}$ are the model
magnitudes for each point generated from the model.
\begin{equation}
\mathcal L = \prod P(BVRI|ISO_{mod}),
\end{equation}
where $\mathcal L$ is the likelihood obtained in the usual manner.
As in MDC10, the tabulated isochrones are taken from \cite{GBBC00} and
\cite{MGB08} and are specified by four parameters, namely, age,
distance, extinction constant and metallicity. In the isochrone
fitting we defined the parameter space as follows:
\begin{enumerate}
\item {\bf Age}: from $log(age)=6.60$ to $log(age)=10.15$;
\item {\bf distance}: from 1 to 10000 parsecs;
\item ${\bf E(B-V)}$: from 0.0 to 3.0.
\end{enumerate}
To simplify the analysis we kept the metallicity constant at the solar
value. This should have no major impact on the evaluation of the
capability of the method to determine the fundamental parameters
because the changes due to metallicity are relatively small when
compared to the other parameters.
\section{Synthetic clusters}
\begin{figure}[!ht]
\begin{center}
\includegraphics[width=\columnwidth]{fit-comp-age.ps}
\caption{ Relative errors for fits of the synthetic clusters fitted
with our method and BVRI photometry only as a function of synthetic
cluster age. The synthetic clusters have fixed $E(B-V)=0.4$,
$3\sigma_{phot}=0.8\%$ and are at a distance of 1000~pc.}
\end{center}
\end{figure}
\begin{figure}[!ht]
\begin{center}
\includegraphics[width=\columnwidth]{fit-comp-nstars.ps}
\caption{ Relative errors for fits of the synthetic clusters fitted
with our method and BVRI photometry only as a function of number of
generated stars. The synthetic clusters have fixed $E(B-V)=0.4$,
$3\sigma_{phot}=0.8\%$, $log(age)=8.0$ and are at a distance of
1000~pc.}
\end{center}
\end{figure}
To test the capability of recovering fundamental parameters using the
CE method we constructed a series of synthetic open clusters. The
generated clusters were chosen to be representative of clusters
present in the catalog (\cite{Dias2002}) and were generated from the
isochrone tables of \cite{GBBC00} and \cite{MGB08}, which are the same
as those used in the fitting algorithm.
We also generated clusters with distinct numbers of stars from low to
high density to evaluate the capacity of the method to retrieve the
parameters in unfavorable conditions.
All clusters were generated adopting solar metallicities, sampled from
a Salpeter IMF and with a 50\% binary population at a distance of
1000~pc, $E(B-V)=0.40$, $3\sigma_{phot}=0.8\%$ and limiting $m_V$
magnitude of 18. All fits were executed 20 times to perform an error
analisis via bootstrap re-sampling.
We did not include simulated contamination from field stars, but this
evaluation is already under way. But as we show in MDC10, the
filtering method we developed removed a good part of the contamination
and we expect that the main effect would likely be an increase in the
estimated uncertainty for the parameters.
The results obtained for the synthetic clusters fitted with our method
and BVRI photometry only are summarized in Fig.1 and Fig. 2 where the
relative errors for fits as a function of synthetic cluster age and
relative errors for fits as a function of number of generated stars
are shown. The distance and $E(B-V)$ were not varied in the results
presented in Figs. 1 and 2 because the major changes were caused by
the variation in age and number of stars. Varying the distance only
affects the photometric errors, which we assumed here to be some
percentage of the observed magnitudes, and so would only increase the
final uncertainty. In Fig. 2 the age was also fixed at $log(age)=8.0$.
To check the reliability of the method under distinct reddening
values, we performed a set of tests with synthetic clusters generated
with $E(B-V)$ varying from 0.3 to 3.0, covering the observed parameter
range, and with $log(age)=8.70~yr$ at a distance of 2100 pc. We note
that an increase in the magnitude limit to $m_V=22$ was necessary for
the synthetic clusters with $E(B-V) > 2$ so that the main sequence
could be minimally sampled. The results presented in Fig. 3 indicate
that the uncertainty of the distance determination tends to increase
with increasing $E(B-V)$, which is to be expected. Interestingly, the
uncertainties obtained for the various reddening values studied were
essentially constant around 0.03, which shows up as declining relative
errors in Fig.2.
\begin{figure}[!ht]
\begin{center}
\includegraphics[width=\columnwidth]{fit-comp-ebv.ps}
\caption{ Relative errors for fits of the synthetic clusters fitted
with our method and BVRI photometry only, as a function of number of
generated $E(B-V)$ values. The synthetic clusters have fixed
$log(age)=8.70~yr$ at a distance of 2100 pc.}
\end{center}
\end{figure}
We also tested the fitting method on a synthetic cluster generated
with the same parameters as the one presented in Sec. 5.1 of MDC10,
i.e., 400 stars with photometric error $3\sigma_{phot}=1\%$ sampled
from an isochrone with $log(age)=8.70~yr$, at a distance of 2100 pc,
$E(B-V)$=0.40 and Z=0.019. The results obtained from the fitting
algorithm were
\begin{itemize}
\item $log(age)=(8.73 \pm 0.05)~yr$
\item $distance=(2100 \pm 70)~pc$
\item $E(B-V)=(0.39 \pm 0.03)$
\end{itemize}
The results indicate that the method using only BVRI is capable of
recovering the cluster parameters with about the same accuracy as the
UBV method presented in MDC10. The precision is also about the same
exept for the reddening, where the BVRI-fitted value shows larger
uncertainty.
\section{Application to observed clusters}
We applied this new approach of fitting isochrones to a couple of
well-studied open clusters. The selected clusters were Trumpler 1 and
Melotte 105, which have observational data in the BVRI filters obtained
by \cite{YS02} (hereafter YS02) and \cite{SMB01} (hereafter SMB01)
respectively. The main reason for choosing these particular clusters
is that they are the only ones that have BVRI data and were also
studied in our previous work (MDC10), where we applied the CE method
using UBV photometry only, following the traditional route of first
determining the reddening through the (U-B) versus (B-V) diagram and
then fitting the theoretical isochrones.
The results obtained for Trumpler 1 and Melotte 105 using our fitting
method and the BVRI photometry agrees very well with the results
obtained by YS02 and SMB01 and the results from our previous work
using only UBV photometry. The values also agree with those of
PN06. In Fig.~4 we show the final-fit results (isochrones) for
Trumpler 1 and in Fig.~5 for Melotte 105, overplotted on
color-magnitude diagrams where the symbol sizes are proportional to
the weight determined for each star based on the filtering scheme
described in MDC10.
For Trumpler 1 the results obtained from our BVRI fit agrees well with
the values obtained by YS02 (see Table 1). Some small but significant
differences are found when we consider the results of PN06 and MDC10,
but these are likely due to weights given to the bright stars as
discussed in MDC10. The weight of the bright stars are likely the main
cause for the age differences. Notice, however, that using the BVRI
only we obtain almost the same E(B-V) as in the fit made with the
(U-B) versus (B-V) diagram.
For Melotte 105 we find a similar situation where we found small
significant differences in E(B-V) and distance when comparing our
results with those of SMB01. The differences are again likely due to
distinct weights given to particular stars. The weight might be
particularly relevant when comparing the results of MDC10, where the
red giants received a lower weight compared to this work, which could
be the reason for the lower distance encountered in that work.
\begin{table}
\caption{CE fit and literature results for Trumpler 1}
\label{table:1}
\centering
\begin{tabular}{c c c c}
\hline\hline
Ref & $E(B-V)$ & Distance (pc) & log(Age(yr)) \\
\hline
PN06 & $0.57 \pm 0.04$ & $2356\pm 511$ & $7.5 \pm 0.1$ \\
MDC10 & $0.59 \pm 0.05$ & $2419\pm 185$ & $7.8 \pm 0.2$ \\
YR02 & $0.60 \pm 0.05$ & $2600\pm 100$ & $7.6 \pm 0.1$ \\
this work & $0.60 \pm 0.03$ & $2706\pm 120$ & $7.2 \pm 0.3$ \\
\hline
\end{tabular}
\end{table}
\begin{table}
\caption{CE fit and literature results for Melotte 105}
\label{table:1}
\centering
\begin{tabular}{c c c c}
\hline\hline
Ref & $E(B-V)$ & Distance (pc) & log(Age(yr)) \\
\hline
PN06 & $0.48 \pm 0.05$ & $2094\pm 159$ & $8.4 \pm 0.1$ \\
MDC10 & $0.47 \pm 0.02$ & $1750\pm 111$ & $8.4 \pm 0.1$ \\
SMB01 & $0.52 \pm 0.05$ & $2300\pm 200$ & $8.4 \pm 0.1$ \\
this work & $0.42 \pm 0.03$ & $1932\pm 95$ & $8.5 \pm 0.1$ \\
\hline
\end{tabular}
\end{table}
\begin{figure*}[!h]
\begin{center}
\includegraphics[scale=0.55]{trumpler1-bvri.eps}
\caption{ Final-fit results for Trumpler 1 overplotted on
color-magnitude diagrams where the symbol sizes are proportional to
the weight determined for each star as described in MDC10.}
\end{center}
\end{figure*}
\begin{figure*}[!h]
\begin{center}
\includegraphics[scale=0.55]{melote105-bvri.eps}
\caption{Same as Fig. 3 for Melotte 105.}
\end{center}
\end{figure*}
\section{Conclusions}
We investigated the possibility of fitting theoretical tabulated
isochrones to open cluster BVRI photometry with a global optimization
method developed in MDC10. In the context of upcoming ground-based
telescopes and space missions that will produce massive amounts of
data, this tool, wich additionally has the possibility of full
automation, is of great value.
The results obtained with synthetic open clusters that are generated
with parameters that cover typical values found in the literature show
that even without the use of the U-filter, fundamental parameters can
be determined with a relative precision of about 10\% or better in
most cases in the studied parameter space, also with an accuracy
similar to the method described in MDC10 using UBV photometry only.
In particular we showed that the reddening can be determined for a wide
range of cluster ages, despite the lack of visual structure in the
color-color diagrams of the clusters. Even when considering objetcs
with few stars the method was able to recover the original parameters
within the uncertainties, showing some systematic behavior for
clusters with a low number of stars as shown in Fig. 2.
It is important to point out that the relative uncertainties discussed
previously are to be taken as lower limits to the real expected
uncertainties that are attainable with real data. The errors reported
are only due to the adopted error in the photometry and Poisson
statistics that arise from drawing a finite number of stars from model
isochrones. Real clusters will have structural variations because of
binary fraction or chemical composition among many other factors, will
affect the final uncertainty of the method. Because it is difficult to
quantify the nature of these factors and their effect on the final
uncertainty, a definitive error analysis should be carried out in a
case-by-case basis.
Applying the method to two observed clusters also studied with our
optimization method and UBV photometry (see MDC10 for details) showed
that the use of BVRI filters only also produced consistent
results. The results agree well with those obtained by previous works,
in particular when considering the results of PN06, where the authors
included a large number of independent literature
determinations. Assuming that the results of PN06 are the best
estimates of the real cluster parameters, we can estimate a better
value of the final uncertainty of the method for these objetcs. For
Trumpler 1 the highest relative uncertainty comes from the distance
and is about 15\%. For Mellote 105 the largest relative error is due
to the reddening being approximately 13\%. These values are likely
more realistic estimates of the real uncertainty of the method for
these two clusters.
Our results show that it is indeed possible to reliably determine
open-cluster fundamental parameters from BVRI photometry only despite
the lack of visual features in the color-color diagrams.
\section{Acknowledgements}
H. Monteiro would like to thank CNPq (grant number
470135/2010-7). W. S. Dias thanks CNPq (grant number 302762/2007-8)
and FAPEMIG (process number APQ-00090-08). We also thank the referee
Dr. Bailer-Jones for extremely useful comments that greatly improved
the paper.
|
\section{Introduction}
The main theorem of the Bass-Serre theory of groups acting on
trees states that a group $G$ acting on a tree $T$ is the
fundamental group of a graph of groups whose vertex and edge
groups are the stabilizers of certain vertices and edges of $T$.
This tells that $G$ can be obtained by successively forming
amalgamated free products and ${\rm HNN}$-extensions.
The pro-$p$ version of this theorem does not hold in general
({\it cf.} Example \ref{ex-4.3}),
namely a pro-$p$ group acting on a pro-$p$ tree does not have
to be isomorphic to the fundamental pro-$p$ group of
a graph of finite $p$-groups (coming from the stabilizers).
Moreover, the fundamental pro-$p$ group of a profinite graph
of pro-$p$ groups does not have to split as
an amalgamated free pro-$p$ product or as
a pro-$p$ ${\rm HNN}$-extension over some edge stabilizer
(the reason is that by deleting an edge
of the profinite graph one may destroy its compactness).
These two facts are usually the main obstacles for proving
subgroup theorems of free constructions in the category of
pro-$p$ groups.
We show that the two Bass-Serre theory principal results
mentioned above hold for finitely generated infinite
pro-$p$ groups acting {\it virtually freely} on pro-$p$ trees,
{\it i.e.} such that the restriction of the action
on some open subgroup is free.
Such a group is then virtually free pro-$p$.
\begin{thm}\label{t-treeacting_intro}
Let $G$ be a finitely generated infinite pro-$p$ group acting
virtually freely on a pro-$p$ tree $T$.
Then
\begin{enumerate}
\item[\rm (a)] $G$ splits either as
an amalgamated free pro-$p$ product or as
a pro-$p$ ${\rm HNN}$-extension over some edge stabilizer;
\item[\rm (b)] $G$ is isomorphic to
the fundamental pro-$p$ group $\Pi_1(\cG,\Gamma)$ of
a finite graph of finite $p$-groups.
\end{enumerate}
\end{thm}
One should say that in contrast to the classical theorem from
Bass-Serre theory our $\Gamma$ in item (b) is not $T/G$.
The graph $\Gamma$ is constructed in a special way by first modifying $T$
without loosing the essential information of the action.
As a corollary we deduce the following subgroup theorem.
\begin{thm}\label{t-subgrouptheorem_intro}
Let $H$ be a finitely generated subgroup of a
fundamental pro-$p$ group $G$ of a finite graph of
finite $p$-groups.
Then $H$ is the fundamental pro-$p$ group of a finite graph of
finite $p$-groups which are intersections of $H$
with some conjugates of vertex and edge groups of $G$.
\end{thm}
Moreover, as an application of Theorem \ref{t-treeacting_intro},
we obtain the following result.
It is a pro-$p$ analogue of a classical result of
G. Baumslag \cite[Thm. 2]{Baumslag:62}
that gave an impulse to the theory known now as the theory of limit groups. Note that our theorem also generalizes
the pro-$p$ {\it ipsis litteris} version of
\cite{BBaumslag:68}, as well as \cite[Thm. 7.3]{KZ:11}.
\begin{thm} \label{t:freeorabelian_intro}
Let $G=A\amalg_{C} B$ be a free pro-$p$ product of
$A$ and $B$ with procyclic amalgamating subgroup $C$.
Suppose that
the centralizer in $G$ of each non-trivial closed subgroup of $C$ is
a free abelian pro-$p$ group and contains $C$ as a direct factor.
If each $2$-generated pro-$p$ subgroup of $A$
and each $2$-generated pro-$p$ subgroup of $B$
is either a free pro-$p$ group or a free abelian pro-$p$ group
then so is each $2$-generated pro-$p$ subgroup of $G$.
\end{thm}
The method of proof is to consider the standard pro-$p$ tree $T$
on which $G$ acts naturally; so $A$ and $B$ are stabilizers of
vertices $v$ and $w$, and $C$ is the stabilizer of the edge
connecting $v$ and $w$.
Then we decompose the pair $(G, T)$ as a
inverse limit of $(G_U,T_U)$ satisfying the hypothesis of
Theorem \ref{t-treeacting_intro}.
\
\noindent {\bf Notation.}
Throughout this paper, $p$ is a fixed but arbitrary prime number.
The additive group of the ring of $p$-adic integers is ${\mathbb Z}_p$;
the natural numbers, ${\mathbb N}$.
For $x$, $y$ in a group we shall write $y^x \!:= x\inv y x$.
All groups are pro-$p$, subgroups are closed and
homomorphisms are continuous.
For $A\subseteq G$ we denote by $\gp A$ the subgroup of $G$
(topologically) generated by $A$
and by $A^G$ the normal closure of $A$ in $G$, {\it i.e.},
the smallest closed normal subgroup of $G$ containing $A$.
By $d(G)$ we denote the smallest cardinality of a
generating subset of $G$.
Recall that a cyclic profinite group is always finite.
The Frattini subgroup of $G$ will be denoted by $\Phi(G)$.
By $\tor G$ we mean the set of all torsion elements of $G$.
For a pro-$p$ group $G$ acting continuously
on a space $X$ we denote the set of fixed points of $G$ by $X^G$
and for each $x\in X$ the {\em point stabilizer} by $G_x$.
We define $\widetilde{G}=\gp{ G_x\,|\, x\in X}$.
The rest of our notation is very standard and basically follows
\cite{RZ:00a}
and
\cite{RZ:00b}.
\section{Preliminary Results}
In this section we collect properties of amalgamated free pro-$p$
products, pro-$p$ {\rm HNN}-extensions and pro-$p$ groups acting
on pro-$p$ trees to be used in the paper.
Further information on this subject can be found in
\cite{RZ:00a} and \cite{RZ:00b}.
Recall the following two notions.
First, an amalgamated free pro-$p$ product
$G\!:=A\amalg_CB$ is {\em non-fictitious}
if $C$ is a proper subgroup of both, $A$ and $B$.
Unless differently stated we shall consider exclusively
non-fictitious free amalgamated products
and we shall make use of the fact from
\cite{Ribes:71} that a free pro-$p$ product
with either procyclic or finite amalgamating subgroup
is always {\em proper}, {\it i.e.},
the factors $A$ and $B$ embed in $G$ via the natural maps.
Second,
a pro-$p$ {\rm HNN}-extension $G={\rm HNN}(H,A,f,t)$
is {\em proper} if the natural map from $H$ to $G$ is injective.
Only such free pro-$p$ products and pro-$p$ {\rm HNN}-extensions
will be used in this paper and they are therefore always proper.
We start with a simple general lemma.
\begin{lemma}\label{l:invsys}
Let $G\!:=\varprojlim G_i$ be the inverse limit of
an inverse system $\{G_i,\varphi_{ij},I\}$ of
pro-$p$ groups and $H_i\le G_i$ so that $\varphi_{ij}(H_i)\le H_j$ holds whenever $j\le i$.
Suppose that there is a constant $d$ with $d(G_i)=d$ for all $i\in I$.
The following statements hold:
\begin{itemize}
\item[{\rm (a)}]
If $d(G)=d$,
then there exists $j\in I$ such that
the projection $G\to G_j$ is surjective.
\item[{\rm (b)}]
For the induced inverse limit $H\!:=\varprojlim H_i\le G$,
we have equality $H^G=\varprojlim H_i^{G_i}$.
\end{itemize}
\end{lemma}
\begin{proof}
For each $i\in I$, let $\varphi_i\colon G\to G_i$ be the
projection.
\noindent (a)
There is an induced inverse system of Frattini
quotients with $G/\Phi(G)=\varprojlim_iG_i/\Phi(G_i)$.
If $\varphi_{ij}(G_i)\Phi(G_j)/\Phi(G_j)$ is a proper subgroup of
$G_j/\Phi(G_j)$, for all $i,j$ belonging to a cofinal subset of $I$,
then $G/\Phi(G)=\varprojlim_i \varphi_{j}(G)\Phi(G_j)/\Phi(G_j)$
and so $G$ can be generated by $d-1$ elements.
Otherwise $\varphi_{ij}$ must be surjective for $i,j$
belonging to a cofinal subset of $I$, and so is $\varphi_j$
({\it cf.} \cite[Prop. 1.1.10]{RZ:00b}).
\medskip
\noindent (b)
Set, for the moment, $K\!:=\varprojlim H_i^{G_i}$.
Since $H\le K$, we have $H^G\le K$, as $K\triangleleft G$.
So, it suffices to establish $K\le H$.
This is certainly true when there is a bound on the orders of the $G_i$.
Fix $n\in\mathbb N$.
Then, as $d(G_i)=d$, there is a bound on the
orders of all $G_i/\Phi^n(G_i)$.
Then the statement reads
$\varprojlim H_i^{G_i}\Phi^n(G_i)\le H^G\Phi^n(G)$
and therefore $K\le H^G\Phi^n(G)$.
Since, $d(G)\le d$, $G$ is finitely generated, and so the set $(\Phi^n(G))_{n\ge1}$
is a fundamental system of neighbourhoods of the identity in $G$
({\it cf.} \cite[Prop. 2.8.13]{RZ:00b}).
Hence $K\le H$, as needed.
\end{proof}
We recollect the following fundamental results
from the theory of pro-$p$ groups acting on pro-$p$ trees
and their consequences for
an amalgamated free pro-$p$ product or
a pro-$p$ {\rm HNN}-extension.
Recall that for a pro-$p$ group $G$ acting on a pro-$p$ tree $T$, the closed subgroup generated by all vertex stabilizers
is denoted by $\widetilde{G}$; also,
the (unique) smallest pro-$p$ subtree of $T$ containing
two vertices $v$ and $w$ of $T$ is denoted by $[v,w]$
and called the geodesic connecting $v$ to $w$ in $T$
({\it cf.} \cite[p. 83]{RZ:00a}).
\begin{thm}
\label{t:trees1}
Let $G$ be a pro-$p$ group acting on a pro-$p$ tree $T$.
\begin{itemize}
\item[{\rm (a)}] {\rm (\protect{\cite[Prop. 3.5]{RZ:00a}})}
$T/\widetilde{G}$ is a pro-$p$ tree.
\item[{\rm (b)}] {\rm (\protect{\cite[Cor. 3.6]{RZ:00a}})}
$G/\widetilde{G}$ is a free pro-$p$ group.
\item[{\rm (c)}] {\rm (\protect{\cite[Cor. 3.8]{RZ:00a}})}
If $v$ and $w$ are two different vertices of $T$,
then $E([v,w])\neq\emptyset$ and
$(G_v\cap G_w)\leq G_e$ for every $e\in E([v,w])$.
\item[{\rm (d)}] {\rm (\protect{\cite[Thm. 3.9]{RZ:00a}})}
If $G$ is finite, then $G=G_v$, for some $v\in V(T)$.
\end{itemize}
\end{thm}
\begin{thm}
\label{t:afpproperties}
Let $G=G_1\coprod_H G_2$ be a proper amalgamated free pro-$p$
product of pro-$p$ groups.
\begin{itemize}
\item[{\rm (a)}] {\rm (\protect{\cite[Thm. 4.2(b)]{RZ:00a}})}
Let $K$ be a finite subgroup of $G$.
Then $K\leq G_i^{g}$ for some $g\in G$ and for some $i=1$ or $2$.
\item[{\rm (b)}] {\rm (\protect{\cite[Thm. 4.3(b)]{RZ:00a}})}
Let $g\in G$. Then
$G_i\cap G_j^{g}\leq H^{b}$
for some $b\in G_i$, whenever $1\leq i\neq j\leq 2$ or $g\not\in G_i$.
\end{itemize}
\end{thm}
\begin{thm}
\label{t:hnnproperties}
Let $G={\rm HNN}(H,A,f)$ be a proper pro-$p$ {\rm HNN}-extension.
\begin{itemize}
\item[{\rm (a)}] {\rm (\protect{\cite[Thm. 4.2(c)]{RZ:00a}})}
Let $K$ be a finite subgroup of $G$.
Then $K\leq H^{g}$ for some $g\in G$.
\item[{\rm (b)}] {\rm (\protect{\cite[Thm. 4.3(c)]{RZ:00a}})}
Let $g\in G $. Then
\begin{equation*}
H\cap H^{g}\leq A^{b}
\end{equation*}
for some $b\in H\cup tH$, whenever $g\not\in H$.
\end{itemize}
\end{thm}
The next technical results concern inverse systems that will play
an essential role during the proof of
Theorem \ref{t:freeorabelian_intro} in section \ref{s:2-generated}.
Until the end of this section the directed set $I$
will be assumed to be order isomorphic to $\mathbb N$.
\begin{prop} \label{p:decomphnn}
Let $G$ be the inverse limit of a surjective inverse system
$\{G_i,\varphi_{ij},I\}$ of pro-$p$ groups.
Suppose that each $G_i={\rm HNN}(H_i,A_i,t_i)$
is an
{\rm HNN}-extension with
$H_i$ finite and $\varphi_{ij}(H_i)\cong H_j$.
Then
there are inverse systems of groups
$\{H_i',\varphi_{ij},I\}$ and
$\{A_i'',\varphi_{ij},I\}$ such that
$G=\text{\rm HNN}(H,A,t)$ with
$H\!:=\varprojlim H_i'$, $A\!:=\varprojlim A_i''$
where each $H_i'$ (resp. $A_i''$)
is a conjugate of $H_i$ (resp. $A_i$)
by an element of $G_i$.
\end{prop}
\begin{proof} \
Fix $k\in I$.
By Theorem \ref{t:hnnproperties}(a)
there are $g_k\in G_k$ with $\varphi_{jk}(H_j)= H_k^{g_k}$
(remember they are isomorphic by the hypothesis).
Pick $g_j\in\varphi_{jk}^{-1}(g_k)$ and define
$H_j' \!:= H_j^{g_j\inv}$, $A_j' \!:= A_j^{g_j\inv}$, and,
$t_j' \!:= t_j^{g_j\inv}$; clearly $\varphi_{jk}(H_j')= H_k$.
Since $A_j = H_j \cap H_j^{t_j}$ and
$\varphi_{jk}$ is surjective, we have
\begin{equation*}
\varphi_{jk}(A_j')
\le H_k \cap H_k^{\varphi_{jk}(t_j')}
\le A_k^{t_k^{\epsilon_k}h_k} \, ,
\end{equation*}
for suitable $h_k\in H_k$ and $\epsilon_k=0$ or $1$,
by Theorem \ref{t:hnnproperties}(b).
Choose $h_j\in\varphi_{jk}^{-1}(h_k)\cap H_j'$ and
$x_j\in\varphi_{jk}^{-1}(t_k)$.
Defining
$A_j'' \!:= {A_j'}^{(x_j^{\epsilon_k}h_j)\inv}$
we obtain
$\varphi_{jk}(A_j'') \le A_k$
in both cases.
Continuing inductively
we obtain the desired inverse systems
$\{H_i',\varphi_{ij},I\}$ and
$\{A_i'',\varphi_{ij},I\}$.
It is straightforward to check that the other associated subgroup
also ``fits'' into the inverse system, that is
$\varphi_{jk}({A_j''}^{t_j''}) \le A_k ^{t_k}$ where $t_j'' \!:=
{t_j'}^{(x_j^{\epsilon_k}h_j)\inv}$.
Now, let $H\!:=\varprojlim H_i'$,
$A\!:=\varprojlim A_i''$ and $B\!:=\varprojlim {A_i''}^{t_i''}$.
For each $i\in I$ let us consider the subset
\begin{equation*}
X_i\!:=\{\tau_i\in G\,|\, A^{\tau_i}\!=\!B
{ \, \ and \, \ }
G_i\!:=\langle H_i, \tau_i\rangle \}\,.
\end{equation*}
Clearly every $X_i$ is compact, and
since $X_{i+1}\subseteq X_i$ for all $i\in I$,
there exists $t\in \bigcap_i X_i$ so that $B=A^t$.
The desired isomorphism from
$\text{\rm HNN}(H,A,t)$ onto $G$ follows now from
the universal property of {\rm HNN}-extensions.
\end{proof}
\begin{prop} \label{p:decompfp}
Let $G$ be the inverse limit
of a surjective inverse system
$\{G_i,\varphi_{ij},I\}$ of pro-$p$ groups
$G_i$ each a free pro-$p$ product
$G_i=A_i\amalg B_i$ whith $A_i$ cyclic and $B_{i}$ procyclic.
Then
there are inverse systems of pro-$p$ groups,
$\{A_{i}',\varphi_{ij},I\}$ and
$\{B_{i}',\varphi_{ij},I\}$,
where each $A_{i}'$
is a conjugate of $A_{i}$
by an element of $G_i$, and $B_{i}'\le G_i$, and,
$G\cong \left(\varprojlim A_{i}'\right)
\amalg \left(\varprojlim B_{i}'\right)$.
\end{prop}
\begin{proof}
Suppose first that there exists $i_0\in I$ such that
$B_{i_0}\cong {\mathbb Z}_p$.
Then, since each $\varphi_{ij}$ is surjective,
the induced homomorphism between the continuous abelianizations
$A_i\times {\mathbb Z}_p\cong
G_i/[G_i,G_i]\to G_{j}/[G_{j},G_{j}]\cong A_j\times {\mathbb Z}_p$
is surjective for $i_0\le j\le i$.
Therefore $\varphi_{ij}(A_i)\le A_j\Phi(G_j)$ and
by Theorem \ref{t:afpproperties}(a)
there is $g_{j}\in G_j$ with $\varphi_{ij}(A_i)\le A_j^{g_j}$
showing that $A_i$ maps onto a conjugate of $A_j$.
Now, observing that $G=\varprojlim G_i$ with
$G_i\cong {\rm HNN}(A_{i},1,t_i)$,
where $t_i$ generates $B_{i}$, we can apply
Proposition \ref{p:decomphnn}
to obtain the result.
Suppose that each $B_{i}$ is finite.
Since $\varphi_{ij}$ are surjective, from
Theorem \ref{t:afpproperties}(a),
we obtain that distinct free factors of $G_i$ are mapped,
up to conjugation, to distinct free factors of $G_j$.
So, there is $k_0$ in $I$ so that
for all $i,j$ we have
\begin{equation*}
\varphi_{ij}(A_{i})= A_{j}^{x_{j}}
{ \ \, and \, \ }
\varphi_{ij}(B_{i})= B_{j}^{y_{j}} \, ,
\end{equation*}
for some $x_{j},y_{j}\in G_j$.
Then inductively the desired inverse systems $\{A_{i}',\varphi_{ij},I\}$ and
$\{B_{i}',\varphi_{ij},I\}$,
can be exhibited.
The result follows now from
\cite[Lemma 9.1.5]{RZ:00b}.
\end{proof}
\begin{lemma} \label{l:2gen}
Let $G$ be a $2$-generated pro-$p$ group.
\begin{itemize}
\item[{\rm (a)}]
If $G$ is a free pro-$p$ product with
procyclic amalgamation,
then one of its free factors is procyclic.
\item[{\rm (b)}]
If $G$ is a proper {\rm HNN}-extension with
procyclic associated subgroups,
then its base subgroup $H$ is at most $2$-generated.
Moreover, if $d(H)=2$ then $H$ is generated by the associated subgroups.
\item[{\rm (c)}]
If $G$ is the fundamental pro-$p$ group of a finite tree of finite groups
such that all edge groups are cyclic,
then either $|G|< \infty$ or $G=K\amalg_{C} R$ with $K$ cyclic and $R$ finite,
or $G=K\amalg_CM\amalg_DN$,
with $K$ and $N$ cyclic and $M\le \Phi(G)$.
\end{itemize}
\end{lemma}
\begin{proof} \
\noindent (a) Suppose that $G=A\amalg_{C} B$ and let ``bar''
indicate passing to the Frattini quotient. We have an obvious
epimorphism from $G$ to the induced pushout
$P\!:= \bar A\amalg_{\bar C} \bar B$. Let
$n\!:=d(A)+d(B)$. Since $C$ is procyclic, the image $M$ of the
kernel of the canonical map $\bar A\amalg \bar B \to \bar G$ via
the cartesian map $\bar A\amalg\bar B\to \bar A\times \bar B$ is
also procyclic. The latter map induces an epimorphism from $\bar
G$ to the at least $(n-1)$-generated elementary abelian pro-$p$
group $(\bar A\times \bar B)/M$. Therefore, $n\leq 3$ and the
result follows.
\noindent (b) Suppose that $G={\rm HNN}(H,C,f,t)$ with $C=\gp c$.
If $d(H)\ge3$ then $d(G)\ge3$ as can be seen by using the obvious
epimorphism $G\to (H\times {\langle t\rangle})/{\langle tct^
{-1}f(c)^ {-1}\rangle}$. Thus $d(H)\le2$.
Finally suppose that $d(H)=2$. Now $G$ is the quotient of
$Q\!:=H\amalg \gp t$ modulo the relation
$f(c)\inv c^{t}$. Since $d(Q)=3$ we can conclude that $c\not\in
\Phi(G)$ and $f(c)\not\in\Phi(G)$. Therefore neither $c\in\Phi(H)$
nor $f(c)\in\Phi(H)$. So we cannot have $f(c)\inv c\in\Phi(H)$
else $d(G/\Phi(G))$ turns out to be 3. Hence $H=\langle C,
C^{t}\rangle$.
\noindent (c) Let $G=\Pi_1({\mathcal G},\Gamma)$ with finite
vertex groups ${\mathcal G}(v)$ and cyclic edge groups ${\mathcal
G}(e)$. We claim that $|V(\Gamma)|\le 3$. By assumption
$|V(\Gamma)|\ge2$, and therefore it has an edge $e$. Splitting $G$
over $e$, we can assume that ${\mathcal G}(d_0(e))$ is procyclic
by (a); hence $d_0(e)$ is a pending vertex of $\Gamma$. Suppose
now that $\Gamma$ has at least $3$ vertices, and let $a$ be an
arbitrary edge $\neq e$ of $\Gamma$ having initial or terminal
vertex $v=d_1(e)$. Without loss of generality, suppose that
$d_0(a)=v$. Then $d_1(a)$ is a pending vertex with procyclic
vertex group ${\mathcal G}(d_1(a))$; for, otherwise, by splitting
$G$ over the edge $a$ we would obtain that $d(G)>2$, a
contradiction. Now, if we have $r\geq 2$ edges with initial or
terminal vertex $v$ then it follows from the pro-$p$ presentation
of $G$ that it has a free pro-$p$ abelian group ${\mathbb Z}_p^r$
as a quotient; this implies $r=2$, whence $|V(\Gamma)|\le 3$.
If $|V(\Gamma)|=2$ then $G=K\amalg_CM$ with $K$ and $M$ finite,
and, by (a), we can assume that $K$ is cyclic.
Suppose now that $|V(\Gamma)|=3$. Then $G=K\amalg_CM\amalg_DN$
with $C$ and $D$ cyclic and $K$, $M$, and $N$ finite. By the
properness of our decomposition we have
$d(K\amalg_CM)=d(M\amalg_DN)=2$ and, making use of (a), we can
conclude that $K$ and $N$ must both be cyclic. Since $d(G)=2$ then
$M\le\Phi(G)$ follows.
\ignor{
}
\end{proof}
\begin{prop} \label{p:treeprod}
Let $G$ be the inverse limit of a surjective inverse system
$\{G_i,\varphi_{ij},I\}$ of pro-$p$ groups
$G_i$. Suppose $G_i$ decomposes as
an amalgamated free pro-$p$ product
$G_i=K_i\amalg_{C_i} R_i$ with $K_i$ cyclic and $R_i$ finite or
$G_i=K_i\amalg_{C_i}M_i\amalg_{D_i}N_i$,
with $K_i$ and $N_i$ cyclic and $M_i\le \Phi(G_i)$. Then,
passing to a cofinal subset of $I$, if necessary,
there are inverse systems
$\{K_i',\varphi_{ij},I\}$ and $\{C_i'',\varphi_{ij},I\}$ such
that $C_i''\le K_i'$,
$\varphi_{ij}(K_i')= K_j'$ and $\varphi_{ij}(C_i'')\le C_j''$
where each $K_i'$ (resp. $C_i''$) is a conjugate of $K_i$ (resp. $C_i$)
by an element of $G_i$.
\end{prop}
\begin{proof}
Using Theorem \ref{t:afpproperties}(a) in both cases we can pass to a cofinal
subset $J$ of $I$ such that for all $i\geq j$ in $J$ we have
$\varphi_{ij}(K_i)\leq K_j^{g_j}$, for some $g_j\in G_j$.
Indeed, in the first case $\varphi_{ij}$ sends factors to the
factors up to conjugation and in the second case $\varphi_{ij}$
sends cyclic factors to cyclic factors up to conjugation. Then in
fact, since $K_j$ is cyclic $\varphi_{ij}(K_i)=K_j^{g_j}$ (indeed,
otherwise $\varphi_{ij}(K_i)^{G_j}\neq K_j^{G_j}$ contradicting
the surjectivity of $\varphi_{ij}$). Now selecting
$g_i\in\varphi_{ij}^{-1}(g_j)$ and letting
$K_i'\!:=K_i^{g_i\inv}$, and using an induction argument, we
obtain the desired inverse system $\{K_i',\varphi_{ij},J\}$. Next,
letting $C_i'\!:=C_i^{g_i\inv}$ we have $C_i'\le K_i'\cap
M_i^{g_i}$; then, by Theorem \ref{t:afpproperties}(b),
$\varphi_{ij}(C_i')\le K_j\cap \varphi_{ij}(M_i)\le {C_j'}^{b_j}$,
for some $b_j\in K_j'$. Choosing $b_i\in\varphi_{ij}^{-1}(b_j)\cap
K_i'$ and letting $C_i''\!:={C_i'}^{b_i\inv}$ we obtain the other
inverse system $\{C_i'',\varphi_{ij},I\}$.
\end{proof}
\begin{lemma}\label{l:fp}
Let $X$ be a $G$-space and $(\widetilde U_n)_{n\ge1}$ be a subset
of normal subgroups of $G_n$ with $\bigcap \widetilde U_n=1$.
Write $X_n\!:=X/\widetilde U_n$ and $G_n\!:=G/\widetilde U_n$. Let
there be subgroups $S_n\le G_n$ so that $\varphi_{nm}(S_n)\le S_m$
and $S\!:=\varprojlim S_n$ be the inverse limit. If
$X_n^{S_n}\neq\emptyset$ for all $n\in\mathbb N$ then
$X^S\neq\emptyset$.
\end{lemma}
\begin{proof}
Let $\varphi_n$ denote the canonical projection from $X\cup G$
onto $X_n\cup G_n$. Then $X_n^{\varphi_n(S)}\supseteq
X_n^{S_n}\neq\emptyset$. Therefore
$Y_n:=\varphi_n\inv(X_n^{\varphi_n(S)})\neq\emptyset$. Now
$Y_n=\{x\in X\mid xS\subseteq x\widetilde U_n\}$ so that
$Y_{n+1}\subseteq Y_n$; by the compactness of $X$ we can deduce
that $\emptyset\neq \bigcap Y_n\subseteq X^S$.
\end{proof}
We end this section by quoting results to be used
in the next section.
\begin{prop}[\protect{\cite[Thm. 1.1]{Scheiderer:99}}]
\label{p:scheiderer}
Let $G$ be a finitely generated pro-$p$ group which
contains an open free pro-$p$ subgroup of index $p$.
Then $G$ is isomorphic to a free pro-$p$ product
\begin{equation*}
H_0 \amalg (S_1\times H_1 )\amalg\cdots\amalg (S_m\times H_m)
\end{equation*}
where $m\geq 0$, the $S_i$ are cyclic groups of order $p$
and the $H_i$ are free pro-$p$ groups of finite rank.
\end{prop}
\begin{cor}[\protect{\cite[Cor. 1.3(a)]{Scheiderer:99}}]
\label{c:finiteconjugacy}
Every pro-$p$ group which contains an open
free pro-$p$ subgroup of finite rank
has, up to conjugation, only a finite number
of finite subgroups.
\end{cor}
\begin{prop}\label{p-freeedgeaction}
A pro-$p$ group $G$ acting on a pro-$p$ tree $T$
with trivial edge stabilizers such that there
exists a continuous section $\sigma:V(T)/G\longrightarrow V(T)$ is
isomorphic to a free pro-$p$ product
\begin{equation*}
\left(\coprod_{\dot v\in V(T)/G} G_{\sigma(\dot v)}\right)\amalg \left(G/\gp{G_w\mid w\in V(T)}\right) \, .
\end{equation*}
\end{prop}
\begin{proof}
This follows from the proof of \protect{\cite[Thm. 3.6]{Z:96}}.
See also the last section of \cite{Melnikov:90}.
\end{proof}
\section{Groups acting virtually freely on trees}
\label{s-proofs}
If a pro-$p$ group $G$ acts on a profinite graph $\Gamma$ we shall
call sometimes $\Gamma$ a $G$-graph.
\begin{lemma}\label{l-actionfree}
Let $G$ be a non-trivial finitely generated pro-$p$ group,
and let $\Gamma$ be a connected $G$-graph.
Suppose that $\Delta$ is a connected subgraph of $\Gamma$
such that $\Delta G=\Gamma$.
Then there exists a minimal set of generators $X$ of $G$
such that $\Delta\cap \Delta x\neq \emptyset$ holds for each $x\in X$.
\end{lemma}
\begin{proof}
It is enough to prove the lemma under the
additional assumption that $G$ is elementary abelian.
Indeed, using ``bar'' to denote passing to the quotient modulo $\Phi(G)$,
making $Z$ a minimal generating set of $G$, suppose that for each
$z\in Z$ there exists a vertex $v_z\in \Delta$ with
$\overline{v_z}\in\ol \Delta\cap \ol \Delta\ol z\neq\emptyset$.
Then there exists $f_z\in\Phi(G)$ with $v_zz\in \Delta f_z$, so that
the set $X\!:=\{zf_z\inv\mid z\in Z\}$ is a minimal set of
generators of $G$ for which the assertion of the Lemma holds.
Suppose that the lemma is false for an elementary abelian group.
Then there is a counterexample $G$ with minimal $d(G)$.
Select a minimal generating subset $X$ of $G$.
If $d(G)=1$ then,
due to the connectedness of $\Gamma$,
there are $g_1,g_2\in G$ with $g_1\neq g_2$ such that
$\Delta g_1\cap \Delta g_2\neq\emptyset$.
Replacing $X$ by $\{g_1g_2\inv\}$ shows that
the conclusion of the lemma holds, a contradiction.
Hence $d(G)\ge2$.
Select an element $z\in X$ and let ``bar'' denote passing
to the quotient modulo $\gp z$.
Since $d(\ol G)=d(G)-1$, by the minimality assumption there is a
subset $\ol Y$ of $\ol G$ which is
a minimal set of generators of $\ol G$ such that
$\ol\Delta\cap \ol\Delta \ol y\neq\emptyset$ for all $\ol y\in \ol Y$.
Let $Y$ denote a transversal of $\ol Y$ in $G$.
Then there are elements $z_y\in \gp z$ such that
$\Delta\cap \Delta yz_y\neq\emptyset$ for all $y\in Y$.
Set $W=\{yz_y\mid y\in Y\}$. Since
$\Gamma$ can be viewed as a $\gp z$-graph,
$\Delta W \cap \Delta W z\neq\emptyset$ can be assumed by the
minimality assumption.
This means that there exist $w_1,w_2\in W$ such that
$\Delta w_1\cap \Delta w_2z\neq \emptyset$, so
$X=W\cup\{w_1^{-1}w_2z\}$ would satisfy the assertion
of the lemma, a contradiction.
\end{proof}
\begin{lemma}\label{l-action}
Let $G$ be a finitely generated infinite pro-$p$ group.
Suppose that $G$ acts on a pro-$p$ tree $T$ containing a
pro-$p$ subtree $D$ such that $DG=T$.
Then there exists a minimal set of generators $X$ of a retract $H$
of $G$ such that $D\cap Dx\neq \emptyset$ for each $x$ in $X$.
\end{lemma}
\begin{proof}
Let ``bar'' denote passing to the quotient modulo
$\widetilde G=\gp{G_v\mid v\in V(T)}$.
By Theorem \ref{t:trees1}(a)
the quotient graph $\ol T\!:=T/\widetilde G$ is a pro-$p$ tree.
Applying Lemma \ref{l-actionfree} to $\ol G$ acting on $\ol T$
yields a subset $Z$ of $G$ such that $\ol Z$ is
a minimal set of generators of $\ol G$ and
for each $z$ there exists a vertex $v_z\in D$ such that
$\overline{v_z}\ol z\in\ol D\cap \ol D\ol z\neq\emptyset$ holds.
Hence there exists $k_z\in \widetilde G$
with $v_zzk_z\in Dk_z\cap Dz$ and so $v_zz\in D\cap Dzk_z^{-1}$.
Now set $X\!:=\{zk_z^{-1}\mid z\in Z\}$ and $H\!:=\gp X$.
Finally observe that by
Theorem \ref{t:trees1}(b),
$\ol G$ is a free pro-$p$ group, so that $H$ is indeed a retract.
\end{proof}
\begin{lemma}\label{l-rankformula}
Let $G$ be a finitely generated pro-$p$ group
acting on a pro-$p$ tree $T$.
Suppose that all vertex stabilizers are finite and
all edge stabilizers are pairwise conjugated.
Assume further that there exist an edge $e\in T$ and
a finite subset $V\subseteq T^{G_e}$ such that:
\begin{itemize}
\item[\rm (i)] for every $v_1, v_2\in V$,
$v_1G=v_2G$ implies $v_1=v_2$,
\item[\rm (ii)] $G$ is generated by the $G_v$, $v\in V$.
\end{itemize}
If $F$ is a free pro-$p$ open normal subgroup of $G$, then
\begin{equation*}
{\rm rank}(F)-1=
[G:F]\left(\frac{|V|-1}{|G_e|}-\sum_{v\in V} \frac{1}{|G_v|}\right) \, .
\end{equation*}
\end{lemma}
\begin{proof}
We use induction on the index $[G:F]$.
Obviously $F\neq G$, from hypothesis (ii); so,
let us consider the preimage $N$ in $G$ of a central subgroup
of order $p$ of $G/F$.
\medskip
\noindent {\it Case 1}. $N\cap G_e=1$.
It follows that each non-trivial torsion element
$t$ of $N$ generates a self-centralized subgroup.
Indeed, by
Theorem \ref{t:trees1}(d)
$t$ stabilizes some vertex $w$, so if $g$ centralizes
$t$, the element $t$ also stabilizes $wg$.
But then by
Theorem \ref{t:trees1}(c)
$t$ stabilizes
the geodesic $[wg,w]$.
Since, however, $G_e\cap N=1$,
the element $t$ cannot stabilize any edge,
so $wg=w$, and therefore $g$ is a power of $t$.
Thus the decomposition of $N$ according to
Proposition \ref{p:scheiderer}
becomes $N=\left(\coprod_{i\in I}C_i\right)\amalg F_1$,
with $F_1$ a free pro-$p$ subgroup of $F$.
Taking into account that $G$ acts upon
the conjugacy classes of subgroups of order $p$ we have
\begin{equation}
\label{e-scheiderer}
N=\coprod_{v\in V}\left(\coprod_{r_v\in G/NG_v}
(N\cap G_{v})^{r_v}\right) \amalg F_1 \, .
\end{equation}
Set $V_1=\{v\in V\mid N\cap G_{v}\neq 1\}$.
Since $NG_v=FG_v$ for every $v\in V_1$ we can rewrite
Eq.~(\ref{e-scheiderer}) as
\begin{equation*}
N=\coprod_{v\in V_1}\left(\coprod_{r_v\in G/FG_v}
(N\cap G_{v})^{r_v}\right) \amalg F_1\, .
\end{equation*}
Using this free decomposition and comparing it with
Proposition \ref{p:scheiderer}
we find
\begin{equation}
|I|=\sum_{v\in V_1}|G/FG_v|=[G:F]\sum_{v\in V_1}\frac{1}{|G_v|} \, ,\label{e-AinLemma}
\end{equation}
and
\begin{equation}
{\rm rank}(F)-1=p\ {\rm rank}(F_1)+(p-1)(|I|-1)-1 \, .\label{e-rankF}
\end{equation}
If $N=G$ then $F_1=1$, since $G_v, v\in V$ generate $G$.
Then $G_e=1$ since otherwise $G$ is finite.
So $|V|=|I|$ and the last equation becomes exactly the needed one.
This gives the base of induction.
Suppose now that $N\not= G$.
Then the product $p\ {\rm rank}\of{F_1}$ can be computed
by observing that passing to the quotient modulo $\torgp{N}$
and indicating it by ``bar'' we have
${\rm rank}\of{\ol F}={\rm rank}\of{F_1}$,
so that using $[G:F]=p[\ol G :\ol F]$
the induction hypothesis yields
\begin{eqnarray*}
p\ {\rm rank}\of{F_1}&=&p\ {\rm rank}\of{\ol F}\\
&=&p[{\ol G}:{\ol F}]\left(\frac{|V|-1}{|\ol G_e|}-\sum_{v\in V} \frac{1}{|\ol G_v|}\right)+p\\
&=&[G:F] \left(\frac{|V|-1}{|G_e|}-\sum_{v\in V_1} \frac{1}{|\ol G_v|}-\sum_{v\in V\,-\, V_1} \frac{1}{|G_v|}\right)+p\\
&=&[G:F]\left(\frac{|V|-1}{|G_e|}-\sum_{v\in V_1} \frac{p}{|G_v|}-\sum_{v\in V\,-\, V_1} \frac{1}{|G_v|}\right)+p
\end{eqnarray*}
(we used $G_e\cap N=1=G_v\cap N$
for all $v\in V\,-\,V_1$ and $|G_v\cap N|=p$ for all $v\in V_1$
to obtain the last equality).
Inserting this expression and the expression for $|I|$
from Eq.~(\ref{e-AinLemma}) into Eq.~(\ref{e-rankF})
yields the claimed formula for ${\rm rank}\of F$.
\medskip
\noindent {\it Case 2.} $N\cap G_e \not= 1$.
Since for all $v\in V$ the edge group $G_e$ is contained in $G_v$
by the hypothesis, $G_v$ centralizes $N\cap G_e$.
But $G=\gp{G_{v} \mid v\in V}$
so $N\cap G_e$ is a central subgroup of $G$ of order $p$.
Then, using ``bar'' to pass to the quotient modulo $N\cap G_e$
and the inductive hypothesis, for $\ol G$ we have
\begin{eqnarray*}
{\rm rank}{(F)}-1&=& {\rm rank}{(\ol F)}-1 \\
&=&[\ol G:\ol F]\left(\frac{|V|-1}{|\ol G_e|}-\sum_{v\in V} \frac{1}{|\ol G_v|}\right) \\
&=&\frac{1}{p}[G:F]\left(\frac{p(|V|-1)}{|G_e|}-\sum_{v\in V} \frac{p}{|G_v|}\right) \\
&=&[G:F]\left(\frac{|V|-1}{|G_e|}-\sum_{v\in V} \frac{1}{|G_v|}\right)
\end{eqnarray*}
as needed.
\end{proof}
Recall that a pro-$p$ group $G$ acts
{\it faithfully} on a pro-$p$ tree $T$ if the kernel of the action is trivial;
and $G$ acts {\it irreducibly} on $T$
if $T$ does not contain a proper $G$-invariant pro-$p$ subtree.
\begin{lemma}\label{l-exchange1
Let a pro-$p$ group $G$ act faithfully and irreducibly
on a pro-$p$ tree $T$.
Suppose that $G_e$ is a minimal edge stabilizer and
the set of edges $E(T^{G_e})G$ is open in $T$.
Then $T'\!:=T\,-\, E(T^{G_e})G$ is a subgraph having each connected component a pro-$p$ tree.
Let $\ol T$ be the quotient graph obtained by collapsing distinct
connected components of $T'$ to distinct points.
Then $\ol T$ is a pro-$p$ tree on which
$G$ acts faithfully and irreducibly, and $\ol T=\ol T^{G_e}G$.
\end{lemma}
\begin{proof}
Since $T'$ is closed and contains $V(T)$, it is a subgraph of $T$;
hence its connected components
are pro-$p$ trees.
Moreover, $\ol T$ is a $G$-graph, and by
\cite[Proposition p. 486]{Zalesskii:89},
it is simply connected and hence a pro-$p$ tree.
Now, we have $m\in T'$ if and only if there exists a
subgroup $L\le G_m$, an edge stabilizer, so that $L^g$ is not
contained in $G_e$ for every $g\in G$.
Therefore, since $G_e$ is a minimal edge stabilizer,
we conclude that all edge stabilizers
of edges in $T^{G_e}G$ are conjugates of $G_e$.
Let us show that $G_{\ol e}$ is a conjugate of $G_e$ for every
$\ol e\in E(\ol T)$.
Let $f\in E(T)$ and $u,v$ be its end points which,
by construction, belong to $T'$.
Fix $g\in G_{\ol e}=G_{\ol u}\cap G_{\ol v}$.
Then $ug$ and $vg$ belong
respectively to the same connected components as $u$ and $v$.
The collapsing procedure induces a
canonical epimorphism which is injective when restricted to
$E(T^{G_e})G$.
Since $\ol T$ is a pro-$p$ tree we find that
after collapsing $e$ and $eg$ both map to $\ol e$, and
as edges of $eG$ under the collapsing procedure
are not identified, $e=eg$ must follow.
Hence $G_{\ol e}$ is a conjugate of $G_e$ indeed.
Suppose that $G$ does not act irreducibly on $\ol T$.
Since $\ol T$ is obtained by collapsing pro-$p$ subtrees,
the preimage of a proper $G$-invariant pro-$p$ subtree of $\ol T$
is a proper $G$-invariant pro-$p$ subtree of $T$; a contradiction.
Suppose that $g\in G$ acts trivially upon all of $\ol T$.
Then, in particular, $\ol eg=\ol e$ and,
as edges of $eG$ under the collapsing procedure are not identified,
we must have $eg=e$, {\it i.e.}, $g\in G_e$.
Therefore the kernel of the action of $G$ upon
$\ol T$ is contained in $G_e$ and so $G_e$ contains a normal
subgroup of $G$ which, by
\cite[Thm. 3.12]{RZ:00a}
must act trivially on $T$.
Hence $G$ acts faithfully on $\ol T$.
\end{proof}
Recall from the introduction that $G$ acts virtually freely
on a space $X$ if some open subgroup $H$ of $G$ acts freely on $X$.
\begin{lemma}\label{l-treetech
Let $G$ be a finitely generated pro-$p$ group acting
faithfully, irreducibly and virtually freely on a pro-$p$ tree $T$.
Then there are a pro-$p$ tree $D$, an edge $e\in E(D)$,
a finite subset $V\subseteq D^{G_e}$ and
a finite subset $X\subseteq G$ such that
\begin{itemize}
\item[{\rm (a)}]
$G$ acts faithfully upon $D$.
\item[{\rm (b)}]
All edge stabilizers are pairwise conjugate;
in particular, $D=D^{G_e}G$.
\item[{\rm (c)}]
for every $v_1,v_2\in V$, $v_1G=v_2G$ implies $v_1=v_2$.
\item[{\rm (d)}]
$X$ freely generates a free pro-$p$ subgroup $H$
such that for $G_0\!:=\gp{G_v\mid v\in V}$
we have $G=\gp{G_0, H}$ and $H\cap G_0^G=1$.
\item[{\rm (e)}] For each $x\in X$, we have
\begin{equation*}
G_e^x\subseteq\bigcup_{v\in V} G_v \, .
\end{equation*}
\end{itemize}
\end{lemma}
\begin{proof}
Let $e\in T$ be an edge with the stabilizer $G_e$
of minimal order.
Let $\Sigma$ denote the set of all non-trivial finite subgroups
$L$ of $G$ that are not conjugate to a subgroup of $G_e$.
Since $G$ is finitely generated,
Corollary \ref{c:finiteconjugacy}
says that there exist up to conjugation
only finitely many finite subgroups in $G$;
in particular there is a finite subset $S$ of
$\Sigma$ such that $\Sigma=\{L^g\mid L\in S, \ g\in G\}$.
Therefore the subset
$T_\Sigma\!:=\{m\in T\mid \exists L\in\Sigma, m\in T^L\}$,
which is the union of all subtrees of fixed points $T^L$ for
subgroups $L\in\Sigma$ can be represented in the form
$T_\Sigma=\bigcup_{L\in S}T^LG$ and is hence a closed
$G$-invariant subgraph of $T$.
Therefore
$E(T^{G_e})G=T\,-\, T_\Sigma$ is open and we can apply
Lemma \ref{l-exchange1}
to obtain, by collapsing the connected components of $T_\Sigma$,
a pro-$p$ tree $D$ on which $G$ acts irreducibly and faithfully with
all edge stabilizers conjugate to $G_e$.
Thus $D$ satisfies (a) and (b).
We come to proving (c),(d) and (e).
Set $N\!:=\gp{G_v\mid v\in V(D)}$.
By Lemma \ref{l-action}
there is finite minimal subset $X$ of generators of
a retract $H$ in $G$ of $G/N$ such that
$D^{G_e}\cap D^{G_e}x\neq\emptyset$ for every $x$ in $X$;
in fact, as $G/N$ is free pro-$p$ by
Theorem \ref{t:trees1}(b),
$X$ freely generates $H$.
Moreover, by the construction of $D$, there is only a finite subset
$V$ of vertices up to translation with stabilizers that
are not conjugates of $G_e$;
to see this we observe that if vertices $v,w$ are both
stabilized by $L\in\Sigma$, then $L$ stabilizes the geodesic $[v,w]$
(see Theorem \ref{t:trees1}(c))
and so $v,w$ belong to the same connected component of $T_\Sigma$.
It follows that $G=\gp{G_v, H\mid v\in V}$ and
$H\cap \gp{G_v\mid v\in V}=1$.
Moreover, since $G$ is pro-$p$ we can reduce $V$ such that
no two distinct vertices of it lie in the same orbit.
By construction, for every group element
$x\in X$ there is
a vertex $v_x\in D^{G_e}$ with $v_xx\inv\in D^{G_e}$.
When $f$ is any edge in the geodesic $[v_x,v_xx\inv]$ then
$G_e=G_f=G_{v_x}\cap G_{v_x}^{x\inv}$
(see Theorem \ref{t:trees1}(c)),
so that, in particular, $G_e^{x}\le G_{v_x}$.
Finally modify $V$ by replacing for every $x\in X$
a vertex $v\in V$ by the vertex $v_x$ whenever $vG=v_xG$.
Then we see that (c), (d), and (e) all hold.
\end{proof}
It is now convenient to introduce a notion of
{\em pro-$p$ {\rm HNN}-extension} as a
generalization of the construction described in
\cite[Sec. 4, p. 97]{RZ:00a}.
\begin{definition}\label{d:HNN-ext}
\rm Suppose that $G$ is a pro-$p$ group, and for a finite set
$X$, there are given monomorphisms $\varphi_x:A_x\to G$ for subgroups
$A_x$ of $G$.
The {\em {\rm HNN}-extension}
$\tilde G\!:={\rm HNN}(G,A_x,\varphi_x,x\in X)$
is defined to be the quotient of $G\amalg F(X)$ modulo the
equations $\varphi_x(a_x)=a_x^x$ for all $x\in X$.
We call $\tilde G$ an {\em {\rm HNN}-extension} and
term $G$ the {\em base group}, $X$ the set of {\em stable letters},
and the subgroups $A_x$ and $B_x\!:=\varphi_x(A_x)$
{\em associated}.
\end{definition}
One can see that every {\rm HNN}-extension in the sense of the
present definition can be obtained
by successively forming {\rm HNN}-extensions, as defined
in \cite{RZ:00a}, each time defining the base group
to be the just constructed group and then
selecting a pair of associated subgroups
and adding a new stable letter.
The {\rm HNN}-extension $\tilde G\!:={\rm HNN}(G,A_x,\varphi_x,X)$
can also be defined by a {\em universal property} as follows.
There are canonical maps
$\tilde f:G\to\tilde G$, $\tilde f_x:A_x\to\tilde G$, $\tilde g:X\to \tilde G$,
with $\tilde f_x(a_x)^{\tilde g(x)}=\tilde f\varphi_x(a_x)$
for all $a_x\in A_x$,
so that, given any pro-$p$ group $H$, any homomorphisms $f:G\to H$, $f_x:A_x\to H$ and a map $g:X\to H$
such that for all $x\in X$ and all $a_x\in A_x$
we have $f(\varphi_i(a_x))=f_x(a_x)^{g(x)}$ then there is a unique
homomorphism $\omega:\tilde G\to H$ with $f=\omega\tilde f$, $g=\omega\tilde g$, and, for all $x\in X$,
$f_x=\omega\tilde f_x$.
\begin{thm}\label{t-treeacting
Let $G$ be a finitely generated infinite pro-$p$ group acting
virtually freely on a pro-$p$ tree $T$.
Then $G$ splits either as an amalgamated free pro-$p$
product or as a proper pro-$p$ ${\rm HNN}$-extension
over some edge stabilizer.
\end{thm}
\begin{proof}\setcounter{claims}0
We consider $G$ to be a counterexample to the theorem
with minimal index $[G:F]$
where $F$ is an open subgroup of $G$ acting freely on $T$.
\bcl
$G$ does not have a non-trivial finite normal subgroup.
In particular, we can assume that
$G$ acts on $T$ faithfully and irreducibly.
\ecl
By \cite[Lemma 3.11]{RZ:00a} there
exists a unique minimal $G$-invariant subtree in $T$.
Replacing $T$ by this subtree we may assume that
the action of $G$ is irreducible.
Now, if $G$ contains a non-trivial finite normal subgroup,
it contains a central subgroup $C$ of order $p$.
By the minimality assumption on $[G:F]$
and as $[{G/C}:{FC/C}]<[G:F]$
the quotient group $\ol G\!:=G/C$ satisfies
the conclusion of the theorem, {\it i.e.} $\ol G$ is either
an amalgamated free pro-$p$ product
$\ol G=\ol G_1\amalg_{\ol H} \ol G_2$ with finite
amalgamating subgroup or it is
an ${\rm HNN}$-extension $\ol G={\rm HNN}(\ol G_1, \ol H, t)$
with finite associated subgroups.
Then $G$ is either a non-trivial amalgamated free pro-$p$ product
$G=G_1\amalg_H G_2$ or ${\rm HNN}(G_1, H, t)$
with $G_1,G_2, H$ being preimages of $\ol G_1,\ol G_2, \ol H$ in $G$,
respectively, as needed.
Hence $G$ does not possess a non-trivial
finite normal subgroup.
Since the vertex stabilizers are finite, the kernel of the
action of $G$ upon $T$ is finite, hence it is trivial.
\medskip
{\em Thus, there exist $D$, $e$, $V$, $X$ and $G_0$
having the properties {\rm (a)}-{\rm (e)} of
Lemma {\rm \ref{l-treetech}}.
Note that the stabilizers of vertices in $D$ may well be infinite.}
\bcl
The pro-$p$ subgroup $H$ of $G$
freely generated by $X$ must be trivial.
\ecl
Suppose that $H\neq1$.
Let $\tilde G={\rm HNN}(G_0, G_e, X)$ and
$\lambda:\tilde G\longrightarrow G$ be the epimorphism
given by the universal property.
By induction on ${\rm rank}\of F$ we show that
$\lambda$ is an isomorphism.
It suffices to show that the rank of $F$ equals the rank of
$\tilde F\!:=\lambda^{-1}(F)$.
If $F_0\!:=G_0\cap F\neq 1$ we can factor out
the normal closure of $F_0$ in $G$
(and, if necessary, the kernel of the action as well)
in order to obtain the quotient group $\ol G$
which acts on $D/F_0^G$ and satisfies
${\rm rank}\of{\ol F}<{\rm rank}\of F$.
Therefore the induced epimorphism
$\ol\lambda:{\rm HNN}(\ol G_0,\ol G_{\ol e},X)\to\ol G$
is an isomorphism, and it is not hard to see that
${\rm HNN}(\ol G_0,\ol G_{\ol e},X)$ is isomorphic to
$\tilde G/\tilde F_0^{\tilde G}$,
where $\tilde F_0\!:=G_0\cap \tilde F$.
This shows that the image $\ol F$ of $F$ in $\ol G$ is
isomorphic to $\tilde F/\tilde F_0^{\tilde G}$.
By Proposition \ref{p-freeedgeaction}
$F$ is a free pro-$p$ product $F\cong F_0\amalg \ol F$ and
$\tilde F\cong \tilde F_0\amalg \tilde F/\tilde F_0^{\tilde G}$, so $F\cong \tilde F$ and we are done.
Thus we may assume that $G_0$ is finite.
Now applying
Lemma \ref{l-rankformula}
to $\tilde G$ and $G$ we deduce that
${\rm rank}(\tilde F)={\rm rank}(F)$, so
$\lambda\restr{\tilde F}$ turns out to be an isomorphism,
contradicting $G$ being a counterexample.
This finishes the proof of the claim.
\bcl
The natural epimorphism
\begin{equation*}
\lambda:{\coprod_{v\in V}}\raisebox{-1.2ex}{\small $G_e$}
G_{v}\longrightarrow G
\end{equation*}
from the free pro-$p$ product of vertex stabilizers $G_v$
amalgamated along the single edge group $G_e$
onto $G$ is an isomorphism.
\ecl
Let us use induction on ${\rm rank}\of F$.
Since $F$ acts freely on $E(D)$
by Proposition \ref{p-freeedgeaction},
for each $v\in V$ the intersection $F\cap G_{v}$
is a free factor of $F$, so similarly as in the proof of Claim 2
we can use the induction hypothesis,
in order to achieve all $G_{v}$ to be finite.
Put $\tilde F=\lambda^{-1}(F)$.
Since $\lambda$ restricted to all $G_v$ is injective,
it suffices to prove that $\lambda\restr{\tilde F}$ is an isomorphism.
But by applying
Lemma \ref{l-rankformula} to $\tilde G$ and $G$
we get that $F$ and $\tilde F$ have the same rank
and therefore $\lambda$ is an isomorphism.
The result follows.
\medskip
Claim 3 shows that $G$ is not a counterexample, a final contradiction.
\end{proof}
\begin{thm}\label{t-fund
A finitely generated pro-$p$ group $G$ acting
virtually freely on a pro-$p$ tree $T$
is isomorphic to the
fundamental pro-$p$ group $\Pi_1(\cG,\Gamma)$ of a finite graph of
finite $p$-groups whose edge and vertex groups are isomorphic to
the stabilizers of some edges and vertices of $T$.
\end{thm}
\begin{proof}
By induction on the rank of a maximal normal free
pro-$p$ subgroup $F$ of $G$.
If ${\rm rank}(F)=0$, that is $G$ is finite,
take as graph of groups the single vertex $G$.
In the general case, we apply
Theorem \ref{t-treeacting}
to split $G$ as an amalgamated free pro-$p$ product
$G=G_1\amalg_K G_2$ or as a pro-$p$ {\rm HNN}-extension
$G={\rm HNN}(G_1,K,t)$ over a finite subgroup $K$.
Moreover, we are free to choose $K$ up to
conjugation in $G_1$.
Then every free factor, or the base group,
satisfies the induction hypothesis and
so exists the fundamental group of a finite graph of finite $p$-groups.
By \cite[Thm. 3.10]{ZM:89}
$K$ is conjugate to some vertex group of $G_1$ and so
we may assume that
$K$ is contained in a vertex group of $G_1$.
Now in the case of an amalgamated product
there is $g_2\in G_2$ such that $K^{g_2}$ is contained in
a vertex group of $G_2$, so $G$ admits a decomposition
$G=G_1^{g_2}\amalg_{K^{g_2}} G_2$.
Thus in both cases $G$ becomes
the fundamental group of a finite graph of finite $p$-groups.
\end{proof}
\begin{thm}\label{t:subgrouptheorem
Let $H$ be a finitely generated subgroup of the
fundamental pro-$p$ group $G$ of a finite graph of finite $p$-groups.
Then $H$ is the fundamental pro-$p$ group of
a finite graph of finite $p$-groups which are
intersections of $H$ with some conjugates of vertex and edge
groups of $G$.
\end{thm}
\begin{proof}
The fundamental pro-$p$ group $G=\Pi_1(\cG,\Gamma)$
acts naturally on the standard pro-$p$ tree $T$
({\it cf.} \cite[Sec. 3]{ZM:89})
and therefore so does $H$.
Moreover, since the graph $\Gamma$ is finite,
there exists an open normal subgroup $U$ of $G$
that intersects all vertex groups trivially and so acts freely on $T$.
Thus Theorem \ref{t-fund} can be applied.
\end{proof}
\begin{example}\label{ex-4.3}\rm
Let $A$ and $B$ be groups of order $2$ and
$G_0=\gp{ A\times B, t\mid A^{t}=B}$ be a
pro-$2$ {\rm HNN}-extension of $A\times B$ with
associated subgroups $A$ and $B$.
Note that $G_0$ admits an automorphism of order $2$ that
swaps $A$ and $B$ and inverts $t$.
Let $G=G_0\rtimes C$ be the holomorph.
Set $H_0=\torgp{G_0}$ and $H=H_0\rtimes C$.
Since $G_0$ acts on its standard pro-$2$ tree
with finite vertex stabilizers, so does $H$.
The main result in
\cite{HZ:10}
shows that $H$ does not decompose as the
fundamental group of a profinite graph of finite $2$-groups.
Its proof also shows that $H$ does not decompose as
an amalgamated free pro-$p$ product or
as a pro-$p$ {\rm HNN}-extension over a finite group.
\end{example}
\section{\texorpdfstring{$2$}{2}-generated subgroups}
\label{s:2-generated}
The final section is devoted to the proof of
Theorem \ref{t:freeorabelian_intro}.
So, henceforth, $G:=A\amalg_{C} B$ is a free pro-$p$ product of
$A$ and $B$ with procyclic amalgamating subgroup $C$
satisfying the following assumptions:
\begin{itemize}
\item[{\rm (i)}]
the centralizer in $G$ of each non-trivial closed subgroup of $C$ is
a free abelian pro-$p$ group and contains $C$ as a direct factor.
\item[{\rm (ii)}]
each $2$-generated pro-$p$ subgroup of $A$
and each $2$-generated pro-$p$ subgroup of $B$
is either a free pro-$p$ group or a free abelian pro-$p$ group.
\end{itemize}
\begin{lemma}\label{l:cent}
For every subgroup $D\le C$ we have $N_G(D)=C_G(D)$.
\end{lemma}
\begin{proof}
By the pro-$p$ version of \cite[Cor. 2.7(ii)]{RZ:96},
\begin{equation*}
N_{G}(D)=N_{A}(D)\amalg_{C}N_{B}(D)\, .
\end{equation*}
Since solvable $2$-generated subgroups of $A$ and $B$ are abelian,
$N_{A}(D)=C_{A}(D)$ and $N_{B}(D)=C_{B}(D)$;
hence $N_G(D)=\gp{C_{A}(D),C_{B}(D)}\subseteq C_G(D)$,
as needed.
\end{proof}
\begin{thm}\label{t:freeorabelian}
Let $G=A\amalg_{C} B$ be a free pro-$p$ product of
$A$ and $B$ with procyclic amalgamating subgroup $C$.
Suppose that
the centralizer in $G$ of each non-trivial closed subgroup of $C$ is
a free abelian pro-$p$ group and contains $C$ as a direct factor.
If each $2$-generated pro-$p$ subgroup of $A$
and each $2$-generated pro-$p$ subgroup of $B$
is either a free pro-$p$ group or a free abelian pro-$p$ group
then so is each $2$-generated pro-$p$ subgroup of $G$.
\end{thm}
\begin{proof}
Let $T$ be the standard pro-$p$ tree on which $G$ acts
({\it cf.} \cite[Sec. 4]{RZ:00a})
and let $L$ be a $2$-generated pro-$p$ subgroup of $G$.
It follows from the definition of $T$ that if
$L$ stabilizes a vertex of $T$, then $L$ is up to conjugation
in one of the free factors of $G$; hence $L$ is either free pro-$p$ or
free abelian pro-$p$, by hypothesis (ii).
Let us assume that $L$ fixes no vertex of $T$.
Since $L$ is finitely generated, we have
$L\cong \varprojlim L/U_n$ where
$\{U_n\mid n\in{\mathbb N}\}$ is a set of open normal subgroups of
$L$ with $\bigcap U_n =1$.
Recall our notation $\widetilde{U_n}$
for the closed subgroup of $U_n$ generated by all vertex
stabilizers with respect to the action of $U_n$ on $T$.
We consider the infinite set $I$ of integers $n$ such that
$U_n/\widetilde{U_n}$ is an infinite free pro-$p$ group
({\it cf.} Theorem \ref{t:trees1}(b)).
So, defining $L_n\!:=L/\widetilde{U_n}$ we see that
$L_n$ acts virtually freely on a pro-$p$ tree $T/\widetilde{U_n}$
({\it cf.} Theorem \ref{t:trees1}(a))
and so we are in position to apply
Theorem \ref{t-fund}
to each of them.
Thus $L_n=\Pi_1(\cL_n,\Gamma_n)$ is the fundamental pro-$p$ group of
a finite graph of finite $p$-groups whose edge and vertex groups
are stabilizers of certain edges and vertices of $T/\tilde U_n$.
Clearly we have $L\cong \varprojlim \{L_n, \varphi_{nm}, I\}$
where each $\varphi_{nm}$ is the canonical map.
Now, since $L/\widetilde{L}$ is a free pro-$p$ group
of rank at most $2$,
we need to consider only the two cases
$L=\widetilde{L}$ and $L/\widetilde{L}\cong {\mathbb Z}_p$;
in the remaining case, when $d(L/\widetilde L)=2$,
$L$ is itself free pro-$p$ of rank 2 -- by the Hopfian property.
We can assume that $\widetilde{L}\neq 1$, otherwise
there is nothing to prove.
\medskip
\noindent {\it Case 1}. $L=\widetilde{L}$.
\medskip
We claim that $\Gamma_n$ is a tree.
If not then there is an edge $e\in \Gamma_n$
so that $L_n={\rm HNN}(P_n,G(e),t)$ for $G(e)$ finite.
But then there is a homomorphism from $L_n$ onto ${\mathbb Z}_p$
contradicting $L_n=\torgp{L_n}$.
Then in light of
Lemma \ref{l:2gen}(c) and of Proposition \ref{p:treeprod},
we have inverse systems of conjugates of $K_n$ and $D_n$;
following the notation of the referred Proposition, we define
two procyclic groups $K\!:=\varprojlim K_n'$ and $D \!:=
\varprojlim D_n''$.
We claim that $D=1$.
Note that since each $D_n$ is an edge stabilizer
with respect to the $L_n$-action, we have
$D=L\cap C^g$, for some $g\in G$.
Since $C_L(D) = L\cap C_G(D)$, it follows from (i) that
$C^g$ is a direct factor of $C_G(D)$,
hence $D$ is a direct factor of $C_L(D)$.
Suppose on the contrary that $D\neq 1$.
Since the procyclic group $K$ contains $D$,
it follows that $D=K$.
Now, the projection $K\to K_{n_0}'$ is
surjective for some sufficiently large $n_0$, by
Lemma \ref{l:invsys}(a).
Hence $D_{n_0}=K_{n_0}$; a contradiction to the
non-fictitious decomposition of $L_{n_0}$.
Thus $D=1$, and so
$\varprojlim {D_n''}^{L_n} =1$, by
Lemma \ref{l:invsys}(b).
Then, writing $L_n\cong K_n'\amalg_{D_n''} R_n'$
we have
$L\cong \varprojlim L_{n}/{D_n''}^{L_n}
\cong \varprojlim (K_n'/D_n'' \amalg R_n'/{D_n''}^{R_n'})$.
Now, if $d(L_n/D_n''^{L_n})=1$ for every $n$, then
$L$ is procyclic; thus
without loss of generality we may and do assume that each
$L_{n}/{D_n''}^{L_n}$ is $2$-generated.
Since $K_n'/D_n''$ is $1$-generated,
so is $R_n'/{D_n''}^{R_n'}$.
Therefore $L\cong \mathbb{Z}_p\amalg \mathbb{Z}_p\,$,
by Proposition \ref{p:decompfp}.
Our proof is finished for {\it Case 1}.
\medskip
\noindent {\it Case 2}. $L/\widetilde{L}\cong {\mathbb Z}_p$.
\medskip
For $n\in \mathbb N$ we have ${\mathbb Z}_p\cong
L/\widetilde L\cong L_n/(\widetilde L/\widetilde U_n)$
and therefore $\Gamma_n$ cannot be a tree.
Then we can select a suitable edge $e_n$ of $\Gamma_n$,
set $\Delta_n:=\Gamma_n\,-\,\{e_n\}$, and present
$L_n={\rm HNN}(K_n,D_n,t_n)$
with cyclic edge group $D_n$ of $e_n$ and
$K_n=\Pi_1({{\mathcal G}_n}\restr{\Delta_n},\Delta_n)$.
Since $\widetilde{L}/\widetilde{U_n}$ is generated by torsion,
as a consequence of
Theorem \ref{t:hnnproperties}(a), it is contained in ${K_n}^{L_n}$;
so, ${\langle {\rm tor}(L_n) \rangle}={K_n}^{L_n}$.
By \cite[Prop. 1.7(ii)]{Zalesskii:04},
$K_n/{\langle {\rm tor}(K_n)\rangle}$ is a free pro-$p$ group,
whence ${\langle {\rm tor}(L_n) \rangle}$ has trivial image
in the quotient ${\rm HNN}(K_n/{\langle {\rm tor}(K_n)\rangle},1, t_n)$
of $L_n$.
Thus $K_n={\langle {\rm tor}(K_n)\rangle}$.
Since $K_n$ acts on the pro-$p$ tree $T/\widetilde{U_n}$
we have $K_n=\widetilde{K_n}$
({\it cf.} Theorem \ref{t:trees1}(d)),
so in particular, $\Delta_n$ must be a tree.
Passing now to a cofinal subset of $\mathbb N$,
if necessary, we may assume that for all $n$ either
$\Delta_n$ is a single vertex or $\Delta_n$ contains an edge.
We discuss the two subcases.
\medskip
\noindent {\it Subcase 2($\alpha$).
For each $n\in\mathbb N$ the tree $\Delta_n$ is a single vertex.}
\medskip
Then $K_n$ is finite.
Passing again to a cofinal subset of $\mathbb N$, if necessary,
we can, making use of
Theorem \ref{t:hnnproperties}(a)
and a projective limit argument, arrange that
$\varphi_{n+1n}(K_{n+1})\le K_n$ holds for all $n$.
Passing again to a cofinal subset of $\mathbb N$, if necessary,
and making use of
Lemma \ref{l:2gen}(b)
we can arrange that for all $n$ either
$K_n$ is cyclic or that $d(K_n)=2$.
We shall discuss the situations when $K\!:=\varprojlim K_n$
is procyclic and when $d(K)=2$.
If $K$ is procyclic, then for every $m$ there exists $n>m$
such that the $\varphi_{nm}(K_n)$ is cyclic and so
$\varphi_{nm}(D_n)=\varphi_{nm}(D_n^{t_n})$.
Hence $\varphi_{nm}(t_n)$ normalizes $\varphi_{nm}(D_n)$
and so $L_m=N_{L_m}(\varphi_{nm}(D_n))$.
Since $L=\varprojlim L_m$ it follows that
$D\!:=\varprojlim D_m$ is normal in $L$.
Since $E(T)$ is a compact $L$-space, setting in
Lemma \ref{l:fp}
$X\!:=E(T)$, $G\!:=L$, and $S_n\!:=D_n$,
we find $e\in E(T)$ with $D\le G_e$.
Therefore $D^g\le C$ for some $g\in G$ and,
if $D\neq 1$, making use of
Lemma \ref{l:cent},
we find that $L\cong {\mathbb Z}_p\times {\mathbb Z}_p$
by hypothesis (i), as needed.
Next assume that $D=1$.
It follows from
Lemma \ref{l:invsys}(b)
that $\varprojlim D_m^{L_m}=1$ and
so $L=\varprojlim L_m/D_m^{L_m}$.
Observing that
$L_m/D_m^{L_m}
=(K_m/K_m\cap D_m^{L_m})\amalg \langle t_m\rangle$
Proposition \ref{p:decompfp}
implies that $L\cong {\mathbb Z}_p\amalg {\mathbb Z}_p$,
whence the result when $K$ is procyclic.
For finishing Subcase 2($\alpha)$
we can now assume that $d(K)=2$.
Then
Lemma \ref{l:invsys}(a)
in conjunction with a projective limit argument implies that
$\varphi_{n+1 n}(K_{n+1})\cong K_n$ for every $n$.
By virtue of
Proposition \ref{p:decomphnn},
we have inverse systems of conjugates $K_n'$ and $D_n''$
of the finite $p$-groups $K_n$ and
$D_n$, and $L ={\rm HNN}(K,D,t)$ where $K\!:=\varprojlim K_n'$ and
$D\!:=\varprojlim D_n''$ is procyclic.
We must have $D\neq 1$,
else $L\cong K\amalg \gp t$, and so $2=d(K)=d(L)-1=1$;
a contradiction.
An application of Lemma \ref{l:fp} shows that
$K$ stabilizes a vertex in $T$; it is therefore, up to conjugation,
contained in either $A$ or $B$ and so by hypothesis (ii) is either
free pro-$p$ or free abelian pro-$p$.
In the first case we observe that
Lemma \ref{l:2gen}(b)
implies that $K=D\amalg D^t$ and so
$L=D\amalg \langle t\rangle$ is a free pro-$p$ group.
So assume in the sequel that
$K$ is a free pro-$p$ abelian group.
Note that $L={\rm HNN}(K,D,t)$ contains
$H\!:=K\amalg_DK^t$ which is not abelian.
On the other hand since $E(T)$ is a compact $L$-space, setting in
Lemma \ref{l:fp}
$X\!:=E(T)$, $G\!:=L$, and $S_n\!:=D_n$
we find $e\in E(T)$ with $D\le G_e$.
Hence $D\le C^g$ for suitable $g\in G$.
Since $D\le C^g$, by hypothesis (i) $C_G(D)$ is abelian,
and it contains $H$; a contradiction.
Hence we are done with Subcase 2($\alpha$).
\medskip
\noindent {\it Subcase 2($\beta$).
For each $n\in\mathbb N$ the tree $\Delta_n$ contains an edge.}
\medskip
Lemma \ref{l:2gen}(c) and Proposition \ref{p:treeprod}
imply that $K_n$ can be written as
$K_n=X_n\amalg_{Z_n}W_n$, with cyclic $p$-groups $X_n$,
and there are inverse systems $\{X'_n\}$ and $\{Z''_n\}$
with $Z''_n\le X'_n$ of conjugates of $X_n$ and $Z_n$
respectively.
Define procyclic groups $X=\varprojlim X_n'$ and
$Z=\varprojlim Z_n''$.
We must have $Z\neq X$ else by
Lemma \ref{l:invsys}(a) we could find $n$ with $Z_n=X_n$
and so the decomposition $K_n=X_n\amalg_{Z_n}W_n$
would be fictitious; a contradiction.
Setting in
Lemma \ref{l:fp}
$X:=E(T)$, $G:=L$, and $S_n:=Z_n''$
we find $e\in E(T)$ with $Z\le G_e$.
Hence there is $g\in G$ with $Z\le C^g$.
Now, since $Z\neq X$,
hypothesis (i) implies $Z=1$.
Let $\bar K_n=K_n/{Z_n}^{K_n}$
and let $\bar D_n$ be the canonical image of $D_n$ in $\bar K_n$.
Then, we consider
\begin{equation*}
\bar L_n \!=\! L_n/{Z_n}^{L_n} \!=\!
\text{\rm HNN}(\bar K_n, \bar D_n, \bar t_n) \!=\! \text{\rm HNN}
(X_n/{Z_n}^{X_n}\amalg W_n/{Z_n}^{W_n}, \bar D_n, \bar t_n)
\, .
\end{equation*}
By Lemma \ref{l:2gen}(b),
each pro-$p$ group $K_n$ is at most $2$-generated,
hence considering $\bar L_n$ modulo its Frattini subgroup,
we can conclude that $d(W_n/{Z_n}^{W_n})=1$.
So, taking into account
Lemma \ref{l:2gen}(b) we conclude that
$X_n/{Z_n}^{X_n}$ and $W_n/{Z_n}^{W_n}$ are isomorphic
cyclic $p$-groups.
Thus $\bar L_n\cong X_n/{Z_n}^{X_n} \amalg
\langle\bar t_n\rangle$.
By Lemma \ref{l:invsys}(b) and Proposition \ref{p:decompfp}
we obtain that $L\cong \varprojlim
\bar L_n \cong {\mathbb Z}_p\amalg {\mathbb Z}_p$.
This concludes the proof of the theorem.
\end{proof}
\begin{cor}\label{c:2-free}
Suppose that neither $A$ nor $B$ contains a
$2$-generated non-procyclic abelian subgroup.
Then any $2$-generated subgroup $L$ of $G$
is a free pro-$p$ group.
\end{cor}
\begin{proof}
Suppose that $L$ is a free abelian pro-$p$ group of rank $2$.
Let $T$ be the standard pro-$p$ tree on which $G$ acts.
Then by
\cite[Thm. 3.18]{RZ:00a}
either $L$ stabilizes a vertex or there
is an edge $e$ of $T$ such that $L/L_e\cong{\mathbb Z}_p$.
But $L$ cannot stabilize a vertex; else it is conjugate to
a subgroup of one of the free factors of $G$,
contradicting the supposition.
Therefore $L/L_e\cong {\mathbb Z}_p$ for some edge $e$.
Since $d(L)=2$ we must have $L_e \not= 1$.
Conjugating $L$ by some element of $G$
we may assume that $L_e$ is contained in $C$.
Then, $L=N_G(L_e)=C_G(L_e)$, by
Lemma \ref{l:cent}(a), and,
by the centralizer condition of the theorem,
$L=C\cong {\mathbb Z}_p$, a contradiction.
Thus, by
Theorem \ref{t:freeorabelian},
$L$ must be free pro-$p$.
\end{proof}
|
\section{Introduction}
Planets are believed to predominantly form in the inner regions of circumstellar disks surrounding young stars, the
so-called planet-forming zone (PFZ). Consequently, searches for embryonic planetary systems are expected to be most fruitful if they are focused
on the inner regions of planet-forming disks, whether using direct detection methods or indirect approaches such as imaging signatures of planet-disk interactions.
While typical gas-rich disks around young stars may extend to 100 AU, or more in a few extreme cases, the outer reaches are too tenuous to
form planetary cores within the lifetime of the disk, unless the disk is massive enough to become gravitationally unstable \citep{Boss97,Marois08}.
A few mature planets or substellar objects have been imaged at large radii around A stars, but the majority of the known planetary systems
are believed to have formed within 20 AU \citep{Pollack96}, followed by radial migration inwards due to planet-disk interactions {\it during the gas-rich phase of the disk} \citep{Alibert04},
consistent with the currently known distribution of exo-planets detected with the radial velocity and transit methods.
\subsection{An observational challenge}
A primary difficulty in studying planet formation {\it in progress} is the small angular sizes of the PFZs. The nearest protoplanetary disks are located,
with a few notable exceptions such as the TW~Hya association and a few Herbig Ae stars, at distances in excess of 100 pc. Resolved images of typical PFZs must be obtained
at a spatial resolution better than 0\farcs1. Further compounding the problem is that many of the main tracers of PFZs are found in the infrared
wavelength range (2-200\,$\mu$m) because the relevant gas temperatures are in excess of $100$\,K -- making ground-based observations challenging.
Generally, specialized instrumentation is needed to obtain the
requisite spatial resolution. In recent years, ground-breaking progress has been made in infrared interferometry of the innermost regions of
protoplanetary disks from 1-2\,$\mu$m \citep[e.g.,][]{Millan-Gabet99, Eisner05} and near 10\,$\mu$m \citep{vanBoekel05}. However, interferometric observations of gas
have thus far been limited to single band or relatively low spectral resolving power and to young A and B stars.
This has, for the most part, limited interferometry to dust, gas continuum and hydrogen recombination lines \citep{Tatulli07,Tatulli08,Isella08,Eisner09}.
Our understanding of the structure and dynamics of inner protoplanetary disks and PFZs remains limited. While there appears to be a common end result of disk evolution -- the formation of a planetary system, complete with gas giants and perhaps smaller rocky planets, as well as, presumably, planetesimals, comets and zodiacal dust -- the pathway is unclear.
It is known that disks exist and that they carry most of the angular momentum
of a young stellar system, as evidenced by resolved sub-millimeter imaging of their outer regions \citep{Koerner93, Mannings97, Qi08}, and matching the momentum
distribution of the solar system and other planetary systems. Their spectral energy distributions show us that the disks are gas-rich throughout; if they were not,
the disks would be flat, not flared with essentially pressure-supported scale heights \citep{Kenyon87}.
The presence of accretion shocks and jets tell us that the disks are actively accreting requiring the
gas to be sufficiently viscious to allow angular momentum transport \citep{Koenigl91}. Limited disk lifetimes of about 6 Myr have been inferred, albeit with significant variation, demonstrating the need
for efficient dispersal mechanisms \citep{Haisch01}. Disks do not extend all the way inward to the stellar surface, but exhibit a complex structure
due to the sublimation of dust at high temperatures leading to rapid opacity gradients \citep{Natta01}, and a coupling of the disk to the stellar magnetic field facilitates
both accretion and mass loss \citep{Koenigl91, Shu94}.
Some disks appear to have large excavated inner regions of low dust opacity \citep{Strom89}, compared to the outer disk, but it appears that not
all disks go through such a stage \citep{Muzerolle10}, and it is still an open question whether this is directly related to the formation
of planetesimals or even giant planets, or whether another dispersal mechanism is at play \citep{Alexander09}. Clearly, observing what the gas actually does in the inner disk
impacts our understanding of all these disk properties and enables us to estimate their relative importance for the evolution of
the disk. Ultimately, the motion of the inner disk gas may reveal the presence of accreting protoplanets \citep{Regaly10}.
\subsection{This paper}
\label{this_paper}
We present a spectro-astrometric imaging survey of molecular gas in the PFZs of disks around a sample of solar type stars using
CRIRES on the European Very Large Telescope \citep{Kaufl04}. The primary goal is to directly determine
the basic distribution and kinematics of the gas and to relate this to the process of planet formation and inner disk evolution. For instance,
in a purely passive, non-accreting, disk it may be expected that the gas orbits at essentially Keplerian speeds, dictated only by the mass of the central star, neglecting minor corrections for the mass of the disk itself ($dV/V_{\rm Kepler}\lesssim 10^{-2}$) and for the radial pressure gradient in the disk \citep[$dV/V_{\rm Kepler}\sim 10^{-3}$][]{Cuzzi93}. The presence of double-peaked emission line profiles from protoplanetary disks can indeed be explained by gas in Keplerian orbits \citep{Carr93, Blake04, Pontoppidan08}. However, disks are in general not passive as they accrete and eventually dissipate, so departures from pure Keplerian motion are expected at some level, and is indicated in FU Ori stars for the CO bandhead at 2.3\,$\mu$m \citep{Calvet91, Martin97, Hartmann04}. Do mid-infrared molecular lines -- in fact -- trace material strictly in Keplerian motion, or are there significant non-Keplerian components present, and, if so, are they dominated by infall or outflow motions? What disk radii are traced by rovibrational CO lines; the inner edge of the disk at 0.01-0.1\,AU, the terrestrial planet region at 0.1-1.0\,AU or the giant planet region at 1.0-10\,AU? What is the origin of the absorption components seen in CO rovibrational spectra -- absorption from edge-on disks, foreground clouds, remnant envelopes or disk winds?
Our spectro-astrometric survey was primarily focused on what is probably the best tracer of molecular gas in the inner disk, or at least the
most easily observable -- the fundamental rovibrational band of CO centered at 4.67\,$\mu$m. These CO lines
are traditionally thought to trace warm gas in disks at $\sim$$1\,$AU, based on typical line widths and excitation temperatures
\citep{Najita03,Blake04}, and are invariably bright in nearly all classical T Tauri and Herbig Ae stars. One of the most basic outcomes
of this survey is the direct measurements of the size of the CO line emitting regions.
We demonstrate that astrometric signals in CO were detected for all sources on AU-scales, but with varying amplitude and with an intriguing range of structure:
The CO spectra are divided into three rough phenomenological classes, based on the line shape in combination with the
shape of astrometric spectra: Keplerian disks characterized by double-peaked line-profiles, single-peaked line sources with broad wings, and self-absorbed sources.
This paper is arranged as follows: In \S\ref{observations} the survey and data reduction are described, including a
detailed discussion of the capabilities of the spectro-astrometric mode of CRIRES for super-resolution imaging.
In \S\ref{Kepler_sources}), we discuss the results of fitting simple Keplerian disk models to the data.
The central issue is that many CO line and astrometric spectra {\it cannot} be explained by Keplerian velocity fields.
We introduce a non-Keplerian class of emission lines in \S\ref{Non_Kepler} and explain why a purely Keplerian model fails.
In \S\ref{wind_model}, we develop a 2-dimensional model that adds a disk wind to a regular Keplerian, flared disk, and demonstrate
that this provides a framework for matching all CO line and astrometric spectra from classical T Tauri stars under specific circumstances, namely if the wind is slow
and uncollimated. We suggest that there is a smooth transition from lines dominated by Keplerian motions to wind-dominated lines, possibly
scaling with the mass-loss/accretion rates. In \S\ref{discussion} the implications for our understanding of disk dispersal and the nature of the warm molecular disk surface layer are discussed.
\section{Observations}
\label{observations}
Spectro-astrometric observations were obtained as part of a large CRIRES survey of infrared molecular emission from protoplanetary disks and young stellar objects
within the framework of the European Southern Observatory (ESO) Large Program 179.C-0151 \citep{Pontoppidan11}.
Spectro-astrometry is a highly sensitive method that allows a single telescope to obtain both spatial and kinematic information on gas-phase lines on very small
scales, $<1$\,milliarcsecond, and at very high spectral resolution, $\lambda/\Delta \lambda \sim 100\,000$. The final accuracy of a spectro-astrometric measurement depends linearly
on both the signal-to-noise and the width of the spatial PSF. Basically, the method measures the spatial centroid offset of the spectrum
as a function of wavelength across a line or other spectral feature,
relative to the continuum. This approach can reveal spatial structure on scales much smaller than the
formal diffraction limit of the observation. Spectro-astrometry was first
developed for chromatic imaging and spectroscopy for the detection of stellar binaries in the visible range using specialized instrumentation \citep{Beckers82, Christy83, Aime88}.
The modern form, using an echelle spectrograph, was first presented by \cite{Bailey98}, and is reviewed by \cite{Whelan08}. Infrared ($\lambda\gtrsim 1\,\mu$m) spectro-astrometry of molecular lines with CRIRES was introduced
by \cite{Pontoppidan08}, who presented CO data from three transitional disks, TW~Hya, HD~135344B and SR~21. They showed that sub-milliarcsecond precisions could
routinely be achieved and that the basic geometries of the line emitting regions --
sizes, inclinations and position angles -- could be determined with a high degree of confidence.
Here we extend this sample to a much wider range of disks in terms of stellar type and evolutionary stage.
\subsection{Observing strategy}
Targets were selected for spectro-astrometric observations according to overall brightness and line-to-continuum contrast, as well as to cover
as wide a range as possible in known disk and stellar characteristics. The line-to-continuum contrast is a particularly important parameter to consider, since
the accuracy of the measured astrometric signal depends roughly linearly on this parameter (see \S\ref{Formulation}).
However, we were successful in obtaining high-quality spectro-astrometry for
sources with line-to-continuum ratios spanning $(F-C)/C \sim 0.1-2$, where $F$ is the total flux and $C$ is the continuum flux. The central stars include spectral types
from K7 to late A, and cover luminosities of 0.1 to 100\,$L_{\odot}$. The targets also span a wide range in accretion rates from $\log \dot{M}\sim -9$ to $-6$\,$M_{\odot}\,\rm yr^{-1}$.
Overall, the sample is intended to represent the diversity found among protoplanetary disks to the extent that is possible given the limitations on the size of the sample.
A log of the observations is given in Table \ref{obs_table}, including those already described in \cite{Pontoppidan08}, while Table \ref{source_prop} summarizes
the properties of the central stars.
The lack of a correction for differential refraction between the effective wavelengh of the slit viewing camera, which usually operates at $H$- or $K$-band, and the spectral wavelength precluded the observation of low-elevation targets for much of duration of the campaign\footnote{A correction for differential refraction was
implemented in the observing software on November 28, 2008.}, including
sources in the Taurus and Chamaeleon star forming regions. As a consequence, most targets are located in Ophiuchus, Serpens, Lupus and Corona Australis.
For any given single observation, the wavelength coverage will not be complete due to the presence of saturated telluric absorption lines. This is
particularly true for CO and water, for which the Earth's atmosphere absorbs in the same transitions as those targeted.
For this reason, a number of observations were repeated with a cadence of 3 months to more than a year. The primary purpose was to take advantage of the Earth's velocity around the Sun
to shift the telluric CO lines relative to those of the target disks. By observing targets at different epochs, complete line profiles could be constructed by combining of spectra obtained at each side of a given object's transit date. An example of how observations at two different epochs were combined to complete the spectral coverage of the CO lines is shown in Figure \ref{epoch_comb}. This strategy also allows for a shallow search for variability in the astrometric spectra, although the shifting telluric absorption often make sensitive comparisons difficult. Variability at such time scales may be expected since the Keplerian time scale at the radii traced, $\sim 0.1-1\,$AU is similar to or shorter than the cadence of observations. We can note that we did not detect any obvious variability in the sources observed during several epochs, but that we do not rule variability below the $\sim$20\% level.
Significant artifacts in the astrometric spectra due to flat field and point spread function (PSF) effects may remain. The latter occurs if the PSF is not
exactly rotation symmetric - and PSFs never are. Indeed, \cite{Brannigan06} found
such artifacts to be a common feature of spectro-astrometric observations. The correction for PSF artifacts is therefore an
essential calibration of spectro-astrometry, without which meaningful analysis is not possible. To
effectively correct for PSF artifacts, all astrometric spectra were obtained with the slit oriented at the desired position angle (P.A.), as well as at an antisymmetric $P.A.+180\degr$. In order to
ensure that the instrument was kept as stable as possible, a special CRIRES observing template was developed in which the
grating and prism angle piezos were kept unchanged while the derotator changed the slit position angle. Furthermore,
identical jitter patterns were used for the parallel and anti-parallel slit positions to maximize the reproducibility of
artifacts. The difference average between these two antisymmetric spectra cancel out PSF artifacts while preserving any real signal.
The efficacy of the procedure is demonstrated in Figure \ref{sa_calib}.
\begin{figure}
\includegraphics[width=8cm]{f1.eps}
\caption[]{Example of how observations at several epochs are combined to fill gaps left by saturated telluric lines. The flux spectra (left) and spectro-astrometry (right)
shown are of VV Ser taken at a slit $\rm P.A.=15\degr$. }
\label{epoch_comb}
\end{figure}
\begin{figure}
\includegraphics[width=8cm]{f2.eps}
\caption[]{Example of how parallel and anti-parallel slit positions are used to calibrate the astrometric spectra. Shown is the P.A.$=55\degr$ CO observation of AS 205N.
It is seen that a single spectrum is filled with astrometric artifacts, in this case mostly due to telluric O$_3$ and CO. Also apparent is a residual low frequency artifact due to
uncertainties in the distortion correction, which is also removed by the self-calibration, producing a flat astrometric final spectrum with an RMS of $\sim$0.5\,mas.}
\label{sa_calib}
\end{figure}
\subsection{Data reduction}
\label{reduction}
The data were reduced using our own IDL scripts. The procedure includes flat-fielding, correction for the
non-linearity of the CRIRES detector response following the description in the CRIRES documentation issue 86.2, co-adding individual nod pairs and correction for spatial distortion. The flux
spectra were extracted using optimal extraction \citep{Horne86}, and were corrected for telluric absorption by division with a
spectrum of an early-type standard star. The telluric standards were corrected for small airmass differences using a simple Beer law by minimizing
the telluric noise in a region of the spectrum relatively clear of intrinsic lines. The CRIRES grating position is not reproducible, so small relative shifts of order a few pixels in the dispersion direction
between the science and telluric spectra were applied. Finally, in some cases small differences in effective resolving power between target and standard star observation were corrected by degrading the
resolving power by $0.1-0.2\,\rm km\,s^{-1}$ in either the science or telluric spectrum. Such differences may arise if differences in AO correction cause differences in the degree to which the source
fills the slit. No attempt was made to include an absolute flux calibration and all spectra were normalized to the continuum flux level.
To further improve the signal-to-noise ratio, lines from similar transitions were averaged, wherever possible.
In particular, this is important for the primary survey of the CO rovibrational band around 4.7\,$\mu$m, where typically 8 nearly identical (in terms
of energy and collisional rates) lines can be averaged in a single CRIRES setting. The uncertainty in the spectro-astrometric signals for typical sources is background-limited.
\subsection{Spectro-astrometric formulation}
\label{Formulation}
For the spectro-astrometry, we find that the most numerically stable and reproducible centroid ($X$) estimator as a function of line velocity $v$ is of the form:
\begin{equation}
X(v) = K \frac{\sum_i (x_i(v)-x_0) F_i(v)}{\sum_i F_i(v)}, \mbox{ [pixels] }
\label{sadef}
\end{equation}
where $x_i$ is the spatial location of a pixel and $F_i$ is the flux contained in that pixel.
The centroid $X(v)$ must also be corrected, using a constant factor $K$, for the fact that not all light will be included within the aperture.
That is, $X(v)$ depends on the range over which $i$ is defined in the centroid estimator. In practice, however, $K$ is small, $< 1.5$, and is
estimated using a modeled PSF measured on the continuum of the spectrum. Consequently,
the relative accuracy on the amplitude of the spectro-astrometric signal is likely $\lesssim 10\%$, which is confirmed by repeated
observations of the same sources (see also \S\ref{observations}).
The uncertainty on the centroid is:
\begin{equation}
\sigma^2 (X) = K^2 \frac{\sum_j([j\sum_iF_i-\sum_i(iF_i)]^2 \times \sigma(F_j)^2)}{(\sum_iF_i)^4},
\label{sadef_unc}
\end{equation}
where the dependency of $X$ and $F$ on $v$ is left out for clarity.
At this point, there is one more issue to consider. In our formulation, the flux is a sum of a continuum term and a line term:
\begin{equation}
F_i=F_{C,i}+F_{L,i}.
\label{terms}
\end{equation}
Normally, one would be interested in determining the centroid offsets for the line term $F_L(v)$ only; the presence of the continuum term
causes $X(v)$ to underestimate the true spatial offset. We call this effect {\it continuum dilution} \citep[see also][]{Pontoppidan08}.
Algebraic manipulation of equations \ref{sadef} and \ref{terms} shows that the relation between the measured and true centroid offset is:
\begin{equation}
X_{\rm true}(v)=X_{\rm obs}(v)\times (1+F_C(v)/F_L(v)).
\label{cont_dilution}
\end{equation}
It is seen that the centroid naturally diverges for $F_L\rightarrow 0$. This makes it inconvenient to display the observed spectra in terms of $X_{\rm true}(v)$, and we
consequently display all spectra in terms of $X_{\rm obs}(v)$. The reader should thus be aware that the true spatial extent of the line emission is higher by a factor $1+F_C(v)/F_L(v)$.
In the following, we define a scalar {\it amplitude} of a given astrometric spectrum as the maximal value of the centroid offset: $A={\rm Max}(|X(v)|)$.
This is a convenient model-independent observable that provides a measure of the size of the line emitting region.
\begin{deluxetable}{lllll}
\tablecaption{Log of observations}
\tablehead{
\colhead{Star} & \colhead{PA} & \colhead{Obs. Date} & \colhead{Int. Time} & \colhead{Spectral Range}\\
\colhead{ } & \colhead{ } & \colhead{ } & \colhead{minutes} & \colhead{$\mu$m}
}
\startdata
LkHa 330 & 0\degr & 29/12/2008 & 24& 4.660-4.770 \\
LkHa 330 & 60\degr & 29/12/2008 & 24& 4.660-4.770 \\
LkHa 330 & 120\degr & 29/12/2008 & 24& 4.660-4.770 \\
CW Tau & 0\degr & 1/1/2009 & 24& 4.660-4.770 \\
CW Tau & 60\degr & 1/1/2009 & 24& 4.660-4.770 \\
CW Tau & 120\degr & 1/1/2009 & 24& 4.660-4.770 \\
DR Tau & 0\degr & 14/10/2007 & 24& 4.805-4.901 \\
DR Tau & 60\degr & 14/10/2007 & 24& 4.805-4.901 \\
DR Tau & 120\degr & 14/10/2007 & 24& 4.805-4.901 \\
TW Hya & 63\degr & 26/4/2007& 40& 4.660-4.770 \\
TW Hya & 153\degr & 26/4/2007& 40& 4.660-4.770 \\
HD 135344B & 0\degr & 22/4/2007& 20& 4.645-4.755 \\
HD 135344B & 60\degr & 4/9/2007 & 20& 4.660-4.770 \\
HD 135344B & 120\degr & 5/9/2007 & 20& 4.660-4.770 \\
GQ Lup & 0\degr & 2/5/2008 & 24& 4.660-4.770 \\
GQ Lup & 60\degr & 2/5/2008 & 24& 4.660-4.770 \\
GQ Lup & 120\degr & 2/5/2008 & 24& 4.660-4.770 \\
GQ Lup & 0\degr & 4/8/2008 & 24& 4.660-4.770 \\
GQ Lup & 60\degr & 4/8/2008 & 24& 4.660-4.770 \\
GQ Lup & 120\degr & 4/8/2008 & 24& 4.660-4.770 \\
HD 142527& 60\degr & 7/8/2008 & 12& 4.639-4.749 \\
HD 142527& 150\degr & 7/8/2008 & 12& 4.639-4.749 \\
RU Lup & 0\degr & 26/4/2007& 20& 4.660-4.770 \\
RU Lup & 0\degr & 27/4/2008& 24& 4.660-4.770 \\
RU Lup & 60\degr & 27/4/2008& 24& 4.660-4.770 \\
RU Lup & 120\degr & 27/4/2008& 24& 4.660-4.770 \\
HD 144432& 6\degr & 2/8/2008 & 16& 4.639-4.749 \\
HD 144432& 66\degr & 2/8/2008 & 16& 4.639-4.749 \\
HD 144432& 126\degr & 2/8/2008 & 16& 4.639-4.749 \\
AS 205N & 55\degr & 29/8/2007& 16& 4.660-4.770 \\
AS 205N & 115\degr & 29/8/2007& 16& 4.660-4.770 \\
AS 205N & 55\degr & 29/8/2007& 16& 2.905-2.977 \\
AS 205N & 115\degr & 29/8/2007& 16& 2.905-2.977 \\
AS 205N & 175\degr & 1/9/2007 & 24& 4.660-4.770 \\
AS 205N & 55\degr & 2/5/2008 & 16& 4.660-4.770 \\
AS 205N & 115\degr & 2/5/2008 & 16& 4.660-4.770 \\
AS 205N & 175\degr & 2/5/2008 & 16& 4.660-4.770 \\
DoAr24E S & 30\degr & 3/9/2007 & 20& 4.660-4.770 \\
DoAr24E S & 90\degr & 2/9/2007 & 20& 4.660-4.770 \\
DoAr24E S & 150\degr & 2/9/2007 & 20& 4.660-4.770 \\
SR 21 & 10\degr & 30/8/2007& 32& 4.660-4.770 \\
SR 21 & 70\degr & 30/8/2007& 32& 4.660-4.770 \\
SR 21 & 130\degr & 31/8/2007& 32& 4.660-4.770 \\
RNO 90 & 0\degr & 25/4/2007& 16& 4.660-4.770 \\
RNO 90 & 60\degr & 26/4/2007& 16& 4.660-4.770 \\
RNO 90 & 120\degr & 26/4/2007& 16& 4.660-4.770 \\
VV Ser & 15\degr & 5/9/2007 & 20& 4.660-4.770 \\
VV Ser & 75\degr & 5/9/2007 & 20& 4.660-4.770 \\
VV Ser & 15\degr & 1/5/2008 & 32& 4.660-4.770 \\
VV Ser & 75\degr & 1/5/2008 & 32& 4.660-4.770 \\
VV Ser & 135\degr & 1/5/2008 & 32& 4.660-4.770 \\
S CrA N & 30\degr & 4/9/2007 & 20& 4.660-4.770 \\
S CrA N & 90\degr & 4/9/2007 & 20& 4.660-4.770 \\
S CrA N & 150\degr & 3/9/2007 & 20& 4.660-4.770 \\
R CrA & 0\degr & 1/9/2007 & 12& 4.660-4.770 \\
T CrA & 0\degr & 26/4/2007& 20& 4.660-4.770 \\
T CrA & 90\degr & 26/4/2007& 20& 4.660-4.770 \\
\enddata
\label{obs_table}
\end{deluxetable}
\begin{deluxetable*}{llllllllll}
\tablecaption{Disk and stellar properties}
\tablehead{
\colhead{Source} & Class\tablenotemark{a} & line profile\tablenotemark{b}&\colhead{distance\tablenotemark{c}} & \colhead{$v_{\rm CO (LSR)}$} & \colhead{$L_*$} & \colhead{Sp. T.} & \colhead{$M_*$\tablenotemark{d}} & \colhead{$\dot{M}$\tablenotemark{e}} & \colhead{references\tablenotemark{f}}\\
& & &\colhead{pc} & \colhead{$\rm km\,s^{-1}$} & \colhead{$L_{\odot}$} & & \colhead{$M_{\odot}$} & \colhead{$M_{\odot}\,\rm yr^{-1}$} &
}
\startdata
LkHa 330 & trans. disk & Keplerian & 250 & 9.0 & 16 & G3 & 2.5 & -8.80/-8.80 & 2,12 \\
CW Tau & CTTS & self-abs. &140 & 7.5 &0.8 & K3 & 1.2 & -8.80/-7.99 & 15,19\\
DR Tau & CTTS & single-peak &140 & 10.7 &0.9 & K5 & 1.0 & -7.5/-5.1& 16,17,18,19\\
TW Hya & trans. disk & Keplerian & 51 & 3.0 &0.23 & K7 & 0.7 & -8.80/-8.80 & 14 \\
HD 135344B& trans. disk & Keplerian & 84 & 7.5 &8 & F3 & 1.6 & -8.30/-8.30 &6\\
GQ Lup & CTTS & Keplerian & 150 & 1.0 &0.8 & K7 & 0.8 & -8.00/-8.00 &11\\
HD 142527 & HAeBe & Keplerian &198 & 5.0 &69 & F6 & 3.5 & -7.16/-7.16 & 6\\
RU Lup & CTTS & single-peak & 150 & 3.5 &0.4 & K7 & 0.7 & -7.70/-7.70 & 5\\
HD 144432 & HAeBe & Keplerian & 145 & 6.0 &10 & A9 & 1.7 & 7.07/-7.07 & 6\\
AS 205N & CTTS & single-peak &125 & 4.5 &7.1 & K5 & 1.1 & -6.14/-6.14 & 10\\
DoAr 24E S & CTTS & self-abs. & 125 & 3.5 &1.3 & K7-M0 & 0.7 & -8.46/-8.46 &8\\
SR 21 & trans. disk & Keplerian & 125 & 3.0 & 15 & G2.5 & 2.2 & $<-8.84$&2\\
RNO 90 & CTTS & Keplerian & 125 & -1.5&4.0 & G5 & 1.5 & -- & 3\\
VV Ser & HAeBe & Keplerian & 415 & 7.0&125 & B1-A3 & 3.0 & -6.34/-6.34 & 9,4\\
S CrA N & CTTS & single-peak &130 & 2.4 & 2.3 & K3 & 1.5 & -- &13\\
R CrA & HAeBe & self-abs. & 130 & 5: &100 & B8-F5 & 3.5: & -7.12/-7.12 & 1\\
T CrA & CTTS & self-abs. & 130 & 7: &8 & F0-F5 & 1.6: & $<-8.20$ & 7
\enddata
\tablenotetext{a}{Type of disk -- can be classical T Tauri star (CTTS), Herbig AeBe star or transition disk }
\tablenotetext{b}{Type of CO rovibrational line profile as discussed in \S\ref{this_paper}.}
\tablenotetext{c}{The distances are based on the current best estimates to the parent young clusters of the disks, many of which are determined by parallax measurements of known cluster members. \citep{Dzib10, Torres09, Loinard08}, \citep{Torres09}. The distance to Corona Australis is well determined using the orbit solution for the eclipsing binary TY CrA \citep{Casey98}.
One exception is HD 135344B, which has an uncertain distance of 84-140\,pc \citep{Grady09}.}
\tablenotetext{d}{The mass of the central star is estimated based on the luminosity and spectral type using the evolutionary tracks of \cite{Siess00}.}
\tablenotetext{e}{Range of mass accretion rates found in the literature.}
\tablenotetext{f}{References used for the stellar properties and the mass accretion rates. }
\tablerefs{ [1] \cite{Bibo92}, [2] \cite{Brown07}, [3] \cite{Chen95}, [4] \cite{Dzib10}, [5] \cite{Herczeg08}, [6] \cite{Meijer08}, [7] \cite{Meyer09}, [8] \cite{Natta06},
[9] \cite{Pontoppidan07a}, [10] \cite{Prato03}, [11] \cite{Seperuelo08}, [12] \cite{Salyk09}, [13] \cite{Schegerer09}, [14] \cite{Thi10}, [15] \cite{White01}, [16] \cite{Mora01}, [17] \cite{Muzerolle03},
[18] \cite{Gullbring00}, [19] \cite{Johns-Krull02}}
\label{source_prop}
\end{deluxetable*}
\begin{figure*}
\centering
\includegraphics[width=17cm]{f3.ps}
\caption[]{Keplerian model fits using the simple flat-disk model superimposed on the data. The flux spectra are shown on top, while the lower panels
show the spectro-astrometry at different slit position angles. The red curves show the model fits. Note that the y-axes are given
in continuum-diluted units, and the real physical extent probed by the spectro-astrometry is larger by factors of ($1+F_C/F_L$). }
\label{kepler_models}
\end{figure*}
\begin{figure}
\includegraphics[width=4.2cm]{f4a.eps}
\includegraphics[width=4.2cm]{f4b.eps}
\includegraphics[width=4.2cm]{f4c.eps}
\includegraphics[width=4.2cm]{f4d.eps}
\includegraphics[width=4.2cm]{f4e.eps}
\includegraphics[width=4.2cm]{f4f.eps}
\includegraphics[width=4.2cm]{f4g.eps}
\includegraphics[width=4.2cm]{f4h.eps}
\includegraphics[width=4.2cm]{f4i.eps}
\includegraphics[width=4.2cm]{f4j.eps}
\includegraphics[width=4.2cm]{f4k.eps}
\includegraphics[width=4.2cm]{f4l.eps}
\caption[]{Goodness-of-fit surfaces, using the $\chi^2$ statistic for the simple Keplerian model fits to the Keplerian sources \citep[see also][]{Pontoppidan08}. }
\label{kepler_chis}
\end{figure}
\section{Keplerian sources}
\label{Kepler_sources}
\subsection{Simple geometric models}
Keplerian sources are characterized by double-peaked line profiles in combination with broad astrometric spectra that display an anti-symmetric pattern
at all slit position angles. For disks viewed at inclinations close to face-on, the double-peak may blend into a single peak, but
if the astrometric spectra show an antisymmetric structure at all position angles, they are still considered Keplerian.
The three disks discussed in \cite{Pontoppidan08}, HD 135344B, SR 21 and TW Hya, are in this category. Prototypical disks with clean double-peaked structure
include GQ Lup, RNO 90 and VV Ser. The astrometric and line flux spectra of the Keplerian disks are shown in Figure \ref{kepler_models}.
As shown in \cite{Pontoppidan08} this structure is well explained by
a simple model of a radial, flat distribution of gas in circular orbits around a point mass, and the antisymmetric structure
is due to the relative spatial displacement of red- and blue-shifted gas in an inclined disk. If the slit is oriented along the major axis of the projected disk,
the maximal astrometric displacement amplitude is seen. Conversely, for an exactly axisymmetric and flat disk, a slit
oriented along the disk minor axis produces no astrometric signal. Given at least two slit position angles, the disk position angle
can be determined with a high degree of confidence. The dominant line emitting radii can be determined, using the
maximal astrometric offset for a slit aligned along the disk major axis. Finally, the disk inclination can be determined with confidence if a stellar mass is assumed and vice versa.
We use the same simple geometric model as \cite{Pontoppidan08} to fit the data and determine the basic geometric parameters for the Keplerian disks. The parameters varied
are the inner emitting radius, the stellar mass, the disk inclination, $i$, and position angle, P.A. Figure \ref{kepler_chis} shows the
resulting goodness-of-fit contours for these four parameters. Table \ref{kepler_table} summarizes the best-fit parameters.
As expected, the stellar mass and the disk inclination are degenerate, but in such a way that even an uncertain assumption of the
stellar mass allows an accurate determination of the disk inclination. The disk position angles are absolutely determined, while
the inner radius is somewhat dependent on the choice of the $T(R)$ relation (here assumed to be a power law with exponent $q=-0.5$). In lieu of the inner radius determined from the fit,
the size of the line emitting ring, in an averaged sense, can also be estimated by using the amplitude $A$, defined above, and correcting for continuum dilution (see \S\ref{Formulation}).
\subsection{A size-luminosity relation for ro-vibrational CO}
In Figure \ref{size_lum}, the astrometric amplitudes are plotted versus the stellar luminosities. It is seen that the Keplerian disks show a
clear correlation across nearly four decades in stellar luminosity as $A\propto L_*^{\alpha}$, with a best-fit exponent of $\alpha=0.48$. This is the relation expected for
the radius of a specific equilibrium temperature as a function of stellar luminosity \citep[e.g.,][]{Dullemond01,Monnier02}. A very similar
size-luminosity relation was found for the $K$-band continuum emission from Herbig Ae/Be stars using the Keck interferometer \citep{Monnier05}.
Here, we find that the molecular gas obeys a similar relation, but on larger scales and extending down to sub-solar luminosities.
\begin{figure}
\centering
\includegraphics[width=8cm]{f5a.eps}
\includegraphics[width=8cm]{f5b.eps}
\caption[]{Relation between the spectro-astrometric amplitude (continuum-corrected) and stellar luminosity for the four classes of astrometric spectra. The astrometric
amplitudes have been corrected for continuum dilution and are normalized to a distance of 125\,pc. Top: The Keplerian profiles divided
into the three transition disks from \cite{Pontoppidan08} and the CTTS's from this paper. The Keplerian disks exhibit a tight correlation with $A\propto L_{*}^{0.5}$.
Bottom: The same as the top plot, but for the single-peaked and self-absorbed sources. No correlation is seen. Further, the self-absorbed sources show much greater amplitudes than
the Keplerian disks. }
\label{size_lum}
\end{figure}
The observed points are compared to the radii of different optically thin grey dust equilibrium temperatures, $R_{S}=1.1\sqrt{\epsilon_Q}(L_*/10^3~L_{\odot})^{1/2}(T_s/1500~K)^{-2}$ \citep{Monnier02}, where $\epsilon_Q$ is the ratio of the dust absorption efficiencies at the color temperatures of the incident and reemiiting radiation fields. Note that
the dust temperature also depends on additional radiation sources, such as that of the surrounding disk \citep{Dullemond01}, which will tend to push the radius at a specific
temperature outwards. However, this prescription allows a direct comparison to the analysis of \cite{Monnier05}.
As is seen in Figure \ref{size_lum}, the gas lines are indeed dominated by gas at radii well beyond the interferometrically measured
dust sublimation radii at $T$$\sim$$1000-1500\,$K, and consistently match dust at 350\,K.
In the limit of a disk truncated at a sharp, optically thick, inner disk rim, the CO sizes correspond to dust at somewhat higher temperatures of $\sim$500\,K.
It is important to realize that some molecular gas may still extend inwards, as indicated by the best-fit inner radii and the high velocities of emission
in the line wings \citep{Salyk07}, but the lines are not dominated by that component; the astrometric sizes measure the bulk of the gas emission.
Of particular interest, however, is that some transition disks fall on the size-luminosity relation defined by the classical disks, specifically TW Hya and HD 135344B --
SR 21 falls somewhat above the relation. This is consistent with
the findings of \cite{Salyk07}, \cite{Pontoppidan08} and \cite{Salyk09}; that the inner disk gas of some transition disks often has not been removed
in the same way as the population of small dust grains. For these disks, it now appears that there is not even a difference in the radii forming the CO gas emission,
and that the lines simply follow the expected dependence on luminosity. This can be interpreted as the removal of the small dust opacity component
through the process of planetesimal formation, as this will preserve a significant column of gas in the inner evolved zone of the disk. The caveat is that this is a very small sample
of transition disks, and the inclusion of additional disks may show a greater degree of complexity, including the operation of other mechanisms of
inner disk clearing, such as photoevaporation or clearing by an unseen, but massive, (stellar) companion, both of which would tend to move the inner edge of the
gas-disk outwards. It will be interesting to see how many transition disks, in fact, fall on the relation.
\begin{table}
\centering
\caption{Best fit model parameters for Keplerian sources}
\begin{tabular}{llll}
\hline
\hline
Star & $R_{\rm in}$ [AU] & P.A. & $i$\tablenotemark{a} \\
\hline
LkhA 330 & 4$\pm$1 & 218$\pm$10$\degr$& 12$\pm 2 \degr$ \\
GQ Lup & $<0.1$ & 357$\pm$10$\degr$& 65$\pm 10\degr$ \\
HD 142527 & 0.2$\pm$0.3 &299$\pm$3$\degr$ & 20$\pm 2\degr$ \\
HD 144432 & 2.7$\pm$0.1 &95$\pm$3$\degr$ & 25$\pm 3\degr$ \\
RNO 90 & 0.06$\pm$0.01 & 177$\pm$3$\degr$ & 37$\pm 4\degr$ \\
VV Ser & 0.3$\pm$0.3 &17$\pm$4$\degr$ & 65$\pm 5\degr$ \\
\hline
\end{tabular}
\tablenotetext{a}{Assuming the stellar masses from Table \ref{source_prop}}
\label{kepler_table}
\end{table}
\subsection{Spectro-astrometry of hydrogen recombination lines in Keplerian sources}
Spectro-astrometry suffers from a fundamental symmetry ambiguity. Because the line centroid offsets are
measured relative to the continuum, departures from circular symmetry in the continuum brightness distribution
will be imprinted in the astrometric signal. For instance, if the continuum emission is due to a sharp inner rim in the
dust disk \citep[as in the models of][]{Dullemond01} and the disk is viewed at some inclination, an asymmetry should be
seen in the astrometric line spectrum when the slit is oriented along the disk minor axis. However, it is fundamentally
not possible to disentangle this effect from an azimuthal asymmetry in the line intensity. In other words, CO spectro-astrometry
cannot distinguish between structure in the spatial distribution of line intensity and the spatial distribution of continuum intensity.
This could be remedied if there were an independent line tracer of the stellar location. In this section, we suggest that hydrogen recombination
lines may, for many young stars, constitute such a tracer that may allow for spectro-astrometry of the continuum.
There is a growing consensus that a dominant contributing process to H~I emission from T Tauri and Herbig Ae stars is
magnetospheric accretion \citep{CalvetHartmann92, Muzerolle98, Bary08}, a shift from an original interpretation in the framework of stellar winds
\citep{Hartmann90, Calvet92, Grinin91}, based on observed P Cygni profiles of the Balmer lines in some sources. While winds likely do contribute
to the optical H~I lines, near-IR lines tracing higher energies and densities appear to be dominated by accretion flows, as
indicated by a general lack of blue-shifted absorption \citep{Folha01}. If the lines are indeed dominated by accretion flows,
the astrometry is expected to constrain the H~I emission to within a few stellar radii, corresponding to scales significantly smaller than
the disk co-rotation radii.
We obtained spectro-astrometry along two perpendicular position angles of the Brackett $\alpha$ lines at 4.05\,$\mu$m for two of the Keplerian sources: RNO 90 and HD 142527.
This line traces somewhat lower energies than e.g. the Brackett $\gamma$ line at shorter wavelengths, 2.16\,$\mu$m, but was chosen because the
4\,$\mu$m continuum source is more likely to be comparable to that at 4.7\,$\mu$m. In particular, it is more likely to trace dust emission, which
is not necessarily the case at 2\,$\mu$m, where gas opacity may dominate \citep{Eisner09}. The resulting spectra
are shown in Figure \ref{Bralpha}. No astrometric signals are detected to limits of 0.35\,mas for RNO 90, while a tentative astrometric signal is seen at the 0.2-0.3\,mas level at a P.A$=60\degr$ for HD 142527.
The formal displacement errors in the two sources correspond to similar physical sizes since HD 142527 is at almost twice the distance of RNO 90 (198\,pc versus 125\,pc).
The conclusion is that the dust continuum emission is azimuthically symmetric and spatial asymmetries seen in the CO emission is not due to spatial structure
of the continuum emission.
These results are consistent with an accretion origin of the infrared H~I lines. Alternatively, formation in a very compact stellar
wind cannot be ruled out (a slight blue-shifted asymmetry is seen in the line profiles). Note that a potential stellar wind giving rise to H~I lines should not
be confused with the much more extended disk wind discussed below in the context of the CO lines. More importantly, in the context of the inner disk,
the lack of strong astrometric signatures from the H~I lines
rule out a sharp, inclined inner edge structure of the 4\,$\mu$m continuum emission. This is a result similar to that found with near-IR interferometry
of the 2\,$\mu$m continuum. It is interesting to note that \cite{Whelan04} found highly extended (10-100\,mas) spectro-astrometric signals of the Paschen
$\beta$ lines (tracing the same n=5 level as the Brackett $\alpha$ line used here) from a few T Tauri stars, including DG Tau, which is known to possess a
strong jet.
\begin{figure}
\centering
\includegraphics[width=7cm]{f6a.eps}
\includegraphics[width=7cm]{f6b.eps}
\caption[]{Spectro-astrometry of the H~I Brackett $\alpha$ lines for RNO~90 and HD~142527. Because the H~I lines are highly over-resolved at
the CRIRES resolution, the spectra have been rebinned to a sampling of 12\,$\rm km\,s^{-1}$ to maximize the signal-to-noise. The formal
noise on the HD142527 astrometric spectrum is $\sim$0.13 mas, or $\sim 0.1\,$AU at a distance of 198\,pc, including continuum dilution. The vertical dashed lines indicate
the velocities of the CO rovibrational line centers. }
\label{Bralpha}
\end{figure}
\section{Non-Keplerian (radial) flows}
\label{Non_Kepler}
Only some CO spectra of protoplanetary disks are as simple to interpret as the Keplerian
disks discussed above. Many show a structure not consistent with a strictly defined Keplerian velocity field.
The class of {\it single-peaked} line sources is characterized by a single peaked line spectrum, but with a broad base extending
to $\sim 50\,\rm km\, s^{-1}$ and narrow astrometric spectra that are highly asymmetric about the continuum
at a characteristic position angle. This spectral class was described by \cite{Najita03}. \cite{Bast10} notes
that the narrow central peaks in these sources do not show a splitting at least down to the
CRIRES spectral resolution of $\sim 3\,\rm km\,s^{-1}$.
One obvious explanation for the narrow peak, favored by Occam's razor, involves a Keplerian disk with line emission from large radii with corresponding low Keplerian velocities.
The spectro-astrometry results rule out this scenario. First, if the lines are formed in a Keplerian disk, the central peak
must be emitted from radii of $R_{\rm peak}\sim GM_*(2\sin{i}/v_{\rm CRIRES}^2)$, corresponding to $>30\,$AU for $i>15\degr$. Such extended emission
should be directly spatially resolved by classical imaging with CRIRES (which has a diffraction-limited 4.7\,$\mu$m spatial resolution of $\sim 20\,$AU at 125\,pc),
yet no extended line emission is observed. However, classical imaging still allows for the possibility of disks with $i<15\degr$. This is where spectro-astrometry steps in by constraining
the line emission to much smaller spatial scales ($\lesssim1\,$AU), regardless of inclination, thus ruling out formation in a Keplerian flow at the $\sim$$100\sigma$ level.
\begin{figure*}
\centering
\includegraphics[width=3.5cm]{f7a.eps}
\includegraphics[width=3.5cm]{f7b.eps}
\includegraphics[width=3.5cm]{f7c.eps}
\includegraphics[width=3.5cm]{f7d.eps}
\caption[]{Flux spectra (top) and spectro-astrometry (lower panels) for the single-peaked line sources. The red curves show possible
wind models (see Table \ref{wind_table})}
\label{peaky_sources}
\end{figure*}
AS~205N is the prototypical example of a source with single-peaked, broad winged line profile. The
astrometric spectra of the three clean single-peaked sources are shown in Figure \ref{peaky_sources}.
For these sources, one out of three position angles, separated by $60\degr$ shows an astrometric line signature
centered on the line velocity, but entirely offset in one direction. The other two angles show the antisymmetric
signature that might be expected from a Keplerian velocity field. The amplitude of the astrometric spectra are $\lesssim 1\,$AU, and
the astrometric line is significantly narrower than the flux profile, with the wings missing.
In essence, the combination of narrow line profiles and small spatial extent indicates sub-Keplerian motions.
We explore a model that explains the qualitative properties of the single-peaked class of CO line spectra in \S\ref{wind_model}.
\subsection{H$_2$O and OH in AS 205N}
A number of other single-peaked line disks are known to show strong emission from
water vapor at 3\,$\mu$m as well as throughout the mid-infrared range. \cite{Salyk08} demonstrated the existence of lines due to gaseous
H$_2$O and OH in the $3\,\mu$m hot band for AS~205N. The lines
have low contrast relative to the continuum (5-10\%). However, CRIRES is significantly more sensitive at 3\,$\mu$m than at 4.7\,$\mu$m,
so a detection of an astrometric signal from the water lines may be possible. AS~205N was observed
in August 2007 with 3 different position angles to match the CO observation. No astrometric signals were detected
either for H$_2$O or for OH. The upper limits are in both cases $0.2 \times (1+F_c/F_l)\,$mas, corresponding to $\lesssim 0.5$\,AU
at a distance of 125\,pc, when correcting for continuum dilution, or an emitting area of $\lesssim 0.8\,\rm AU^2$. This is consistent with a measured emitting
area of $0.4\,\rm AU^2$ for the 3\,$\mu$m water lines found by \cite{Salyk08}, and is marginally smaller than the radial extent of the CO emission (0.7\,AU).
Since the excitation energies of the $3\,\mu$m water and OH lines are higher than those of CO, it is likely that the smaller extent can be explained by
an origin in somewhat warmer gas.
\begin{figure}
\includegraphics[width=8cm]{f8.eps}
\vspace{0.4cm} \\
\caption[]{Spectro-astrometry of water and OH lines from the AS 205N disk. }
\label{SA_H2O}
\end{figure}
\subsection{Self-absorbed sources}
Finally, self-absorbed sources have strong absorption components superposed on broad emission lines and are consequently
more complex. While the gas forming the absorption lines in some sources may be completely unrelated to
the disk, it is discussed in section \ref{wind_model} how some self-absorbed sources may be a different representation of
the peaky line sources, namely those viewed at a higher inclination angle.
The self-absorbed sources tend to show the highest amplitude astrometric spectra of the survey. The
physical difference between the single-peaked and self-absorbed sources and the Keplerian disks is
illustrated by their astrometric amplitudes in Figure \ref{size_lum}. These sources do not fall along a neat correlation with the
stellar luminosity as do the Keplerian disks. T CrA and R CrA in particular show astrometric amplitudes as high as 20-30\,AU, indicating
that the absorbing gas is structured on much larger scales than the emitting gas. These sources are likely to be younger and more embedded than the remaining sample, which
may impact the observables.
\begin{figure*}
\centering
\includegraphics[width=3.5cm]{f9a.eps}
\includegraphics[width=3.5cm]{f9b.eps}
\includegraphics[width=3.5cm]{f9c.eps}
\includegraphics[width=3.5cm]{f9d.eps}
\caption[]{Spectro-astrometry for the self-absorbed sources. Note that while a model of DoAr24E S exists that fits the spectro-astrometry, the flux line profile is not well matched.
One explanation for this is that there is an additional compact component to the line. No attempt is made to fit models to T CrA, }
\label{selfabs_sources}
\end{figure*}
\section{A new unified disk+wind model}
\label{wind_model}
\subsection{Parametrization}
\begin{figure*}[ht]
\includegraphics[width=16cm]{f10.eps}
\caption[]{Sketch of a disk wind model that qualitatively reproduces the broad single-peaked CO line spectra and CRIRES spectro-astrometry from highly accreting T Tauri stars
-- in this example, AS 205N, a T Tauri star in Ophiuchus. The left panels show the observed CO rovibrational ($v=1-0$ around 4.7\,$\mu$m) line spectrum and
spectro-astrometry at 3 different slit position angles. The middle panels show two 2D radiative transfer models: one for a purely Keplerian disk, and one
for a Keplerian disk in which a wide-angle molecular wind is launched from the disk surface. The right panel contains a sketch of the basic wind geometry.
}
\label{Wind_sketch}
\end{figure*}
\begin{figure}[ht]
\includegraphics[width=8cm]{f11.eps}
\caption[]{Radial and azimuthal velocities of the gas in the parametrized disk+wind model. For this set of parameters, which reproduce the observables in, e.g., AS 205N, the wind is seen to
be strongly sub-Keplerian beyond 0.5 AU. Note that the finite radius of the inner rim of the disk and a locus distance $>0$ prevents the wind from filling in all polar angles. In this case, the
wind only exists at polar angles of $\theta\gtrsim 50\degr$ (as measured from the disk axis). For angles closer to the disk axis, the velocity field is still Keplerian, although the
gas density here is far too low to generate any line emission. }
\label{Wind}
\end{figure}
A parametrized disk wind model is used to test whether an idealized wind structure can produce a plausible match to the phenomenology of the single-peaked line profiles and associated
asymmetric astrometric spectra. It is based on models for UV resonance and hydrogen line emission from accretion disk winds \citep[e.g.,][]{Knigge95, Kurosawa06},
and represents a computationally convenient structure inspired by numerical results from magneto-hydrodynamical (MHD) simulations of
magnetized disk winds \citep{Koenigl00}.
The basic requirement of the observed spectro-astrometry is that the line forming gas must be orbiting the central star with
strongly sub-Keplerian azimuthal velocities in order to produce the single peak without requiring that the emission is extended
at the spatial resolution of CRIRES ($\sim$0\farcs15). A wide angle wind provides a convenient physical
way of accomplishing this through simple conservation of angular momentum -- as a gas parcel is forced outwards due to the wind pressure,
the azimuthal velocity decreases linearly with radius, in comparison with the underlying Keplerian disk in which the velocity experiences a shallower decrease as $R^{-1/2}$.
This generates gas above the disk that is supported by wind pressure and orbits at low azimuthal velocities.
Following \cite{Kurosawa06}, the wind is constructed as a set of linear streamlines with a locus below the central star at a distance $d$ in units of $R_*$. This
generates a conical wind with no flow along the disk axis. Briefly, the wind is accelerated along the
field lines as:
\begin{equation}
v_{\rm stream} = c_s + (f*v_{\rm esc}-c_s) \times (1-A_{\rm scale}/l+A_{\rm scale})^{\beta},
\end{equation}
where $l$ is the coordinate along the stream line, $c_s$ is the sound speed, $v_{\rm esc}$ is the asymptotic velocity at
the end of the stream line and $A_{\rm scale}$ is the scale of the acceleration region of the wind. $\beta$ is the
wind acceleration parameter. Requiring angular momentum conservation, the azimuthal velocity component is:
\begin{equation}
v_{\phi} = v_{\rm Kepler}(l_0) \times \frac{R(l_0)}{R},
\end{equation}
where $R$ is the radial disk coordinate.
The density of the wind is calculated assuming mass conservation:
\begin{equation}
\rho_{\rm wind} = \frac{\dot{\Sigma}(R)}{v_{\rm stream}|\cos\delta|}\times\left[\frac{d}{S\cos\delta}\right]^2
\end{equation}
Here, $\dot{\Sigma}(w) \propto R^{-p}$ is the local mass-loss rate, $\delta$ is the angle between the stream line and the disk normal and S is the distance to the wind locus.
The exponent of the local mass loss rate is taken to be $p=7/2$ \citep{Krasnopolsky03}. The
total wind mass loss rate can be calculated by integrating over the disk and multiplying by two to include the opposite surface:
\begin{equation}
\dot{M} = 2\int_{R_{\rm inner}}^{R_{\rm outer}} 2\pi R \dot{\Sigma}(R) dR
\end{equation}
\begin{figure}
\centering
\includegraphics[width=8cm]{f12.eps}
\caption[]{Effect of inclination on the line profile in the disk wind model. }
\label{peaky_incl}
\end{figure}
The raytracer RADLite \citep{Pontoppidan09} is used to render model lines and spectro-astrometry for the wind models, based on a generic
model of a flared protoplanetary disk, and assuming level populations in LTE. Specifically, the temperature structure is assumed to be in
equilibrium with the stellar radiation field and dominated by dust heating/cooling. In reality, the heating of the wind is likely to be dominated by
photo-electric heating similar to the heating of the disk atmosphere \citep{Kamp04, Jonkheid04, Gorti04, Dullemond07} or, perhaps, ambipolar diffusion \citep{Safier93}. The cooling
may be dominated by adiabatic expansion and molecular cooling (e.g., partly via the observed CO and H$_2$O lines). However, we restrict ourselves
to qualitative models in this paper (see also \S\ref{Caveats}), since a detailed and appropriate physical treatment of the thermal wind structure required to
match the observations will be likely be a significant study in its own right.
\subsection{Properties of the wind model observables}
Figure \ref{Wind_sketch} illustrates the wind geometry and compares the observables generated using the wind model for
the spectro-astrometry of AS~205N. The total mass-loss rate is $9\times 10^{-9}\,\rm M_{\odot}\,yr^{-1}$, assuming a CO abundance
of $5\times 10^{-5}$ relative to $\rm H$. This mass-loss rate is consistent with a
highly accreting CTTS. Indeed, AS~205N is accreting at a rate of $7\times 10^{-7}\,\rm M_{\odot}\,yr^{-1}$ \citep{Eisner05}. There is
no treatment of the chemistry of the wind in the present model, so if the abundance of CO is lower than that assumed, the mass-loss rate will be correspondingly higher.
However, the numbers appear to be physically reasonable; the ratio of mass-loss to mass-accretion rates is $\dot{M}_{\rm wind}/\dot{M}_{\rm acc}\gtrsim 1.5\%$,
and as this ratio is expected to be as high as 10\% \citep{Koenigl00}, the CO abundance in the wind could in fact be lower than assumed.
There are many free parameters in the wind model, so that shown represents one possibility
that reproduces the main characteristics of the observed astrometric spectra. A full parameter study, beyond the scope of
this paper, may reveal significant degeneracies. That said, some parameters are well constrained, including the position angle of the system.
General properties are also locus distances $d\lesssim 10\,R_*$ and inclinations of $i\lesssim 30\degr$. The velocity field
used for AS~205N is shown in Figure \ref{Wind}.
It is seen that the wind model can explain the key properties of the combined line profile and astrometric spectra:
namely it readily produces single-peaked lines with broad wings. The wings are dominated by the innermost, Keplerian disk,
while the single peak is generated by the sub-Keplerian gas at a few AU. The presence of a strongly
sub-Keplerian component requires a wide-angle, non-collimated wind, in the model parametrized as a small
star-to-locus distance, $d$. Increasing $d$ results in a double-peaked line profile that would
be resolvable with CRIRES. The astrometric spectra become highly asymmetric
at P.A.'s close to the disk minor axis. It is important to note that this wind geometry does not necessarily produce strongly blue-shifted lines --
which might have been expected. The uncollimated winds sees much of the line emission coming from gas moving on trajectories nearly parallel to the disk surface.
For sources viewed at low inclination angles, this generates low radial velocities and prevents large blue-shifts, as
required by the data. The addition of a significant amount of gas at high altitudes where the optical depths toward the central star are low, results in
higher column densities of warm CO, leading to brighter lines with higher line-to-continuum ratios, a property of this class of CO spectra
that has been observed \citep{Bast10}.
Table \ref{wind_table} summarizes the RADLite wind model parameters for the wind-dominated disks. Since we do not
fully explore the parameter space, the entries in the table represent a possible model, but with a caution that significant
degeneracies may exist. Further, the use of LTE level populations and the assumption of coupled dust and gas temperatures
almost certainly introduce inaccuracies in the implied wind structures. Indeed, it was typically necessary to increase the luminosity of the central
source to values well above the known stellar properties to match the amplitude
of the astrometric signal, especially in the case of RU Lup, the reason being that the gas temperatures are significantly higher
than what can be explained by pure coupling to dust.
\subsection{The wind model at higher inclinations as a model for the self-absorbed disks}
As can be seen in Figure \ref{Wind}, there is no wind component along the disk axis due to the finite size of the inner disk rim. Because of this,
a disk wind viewed close to face-on, as is likely the case for AS 205N, allows a free view of the continuum emission from the innermost disk, consistent
with the low extinction toward the central star in many of our targets.
However, if the inclination were higher, say $\sim 45\degr$, the entire wind column will be located between the disk continuum emission and the
observer. This leads to the formation of a strong line absorption component, as illustrated in Figure \ref{peaky_incl}. The Figure shows the line profiles of the prototypical
AS 205N model, but viewed at higher inclination angles. A deep absorption line forms at inclinations higher than $\sim 40\degr$ for this particular model, a result that can
be compared directly to sources exhibiting a self-absorbed line, such as CW Tau. The line profile of this source is qualitatively reproduced by
the AS 205N model viewed at an inclination angle of $55\degr$. The depth of the absorption line is smaller in the model, but can easily be deepened by
increasing the mass-loss rate.
We note that the small spectro-astrometric sample of self-absorbed source includes
sources such as R CrA and T CrA are clearly younger and more embedded than the remaining disk sample, and their
deeper absorption components are consistent with mass-loss rates 2-3 orders of magnitudes higher than that of the AS 205N model.
However, there are more bona-fide isolated disks in our CRIRES survey showing strong, warm CO absorption that were not included as spectro-astrometric targets,
but with properties otherwise resembling CW Tau.
The four self-absorbed disks are not modeled in detail for several reasons. We found it difficult to reproduce the very high amplitudes of
the astrometric offsets for the absorption component for CW Tau and T CrA, likely indicating the presence of extended continuum emission structure.
DoAr 24E S has several absorption components that also cannot be modeled exactly with the wind model without the additional parameters.
However, we do still present a best effort model of DoAr 24E S in Figure \ref{selfabs_sources} to illustrate the required complexity of this star.
\begin{table}
\centering
\caption{Model parameters for wind-dominated sources}
\begin{tabular}{llllll}
\hline
\hline
Star & P.A.\tablenotemark{a} & i\tablenotemark{b} & $L_{\rm model}/L_*$\tablenotemark{c} & d\tablenotemark{d} &$\dot{M}_{\rm wind}$\tablenotemark{e} \\
& & & & $R_*$ & $M_{\odot}\,\rm yr^{-1}$ \\
\hline
DR Tau & 0\degr: & -9\degr & 1.0 & 1 & $4\,10^{-8}$\\
RU Lup & 80\degr & 35\degr & 90 & 8 & $4\,10^{-9}$\\
AS 205N & 235\degr & 20\degr & 1.8 & 5 & $9\,10^{-9}$ \\
DoAr24E S & 235\degr & 20\degr & 1.8 & 5 & $9\,10^{-9}$ \\
S CrA & 15\degr & 10\degr & 1.3& 4 & $1\,10^{-8}$ \\
\hline
\end{tabular}
\tablenotetext{a}{Position angle of the disk major axis.}
\tablenotetext{b}{Inclination of the disk rotation axis. A negative value of the inclination corresponds to the north pole of the disk facing the observer. }
\tablenotetext{c}{Ratio between the luminosity of the central source used in the model and the actual stellar luminosity. A high ratio
indicates that the gas is heated to temperatures significantly higher than those of the dust.}
\tablenotetext{d}{Distance between the locus of the wind and the central star (in units of stellar radius). d=0 corresponds to a spherical wind.}
\tablenotetext{e}{Mass loss rate of the wind assuming $x({\rm CO})=5\,10^{-5}$.}
\label{wind_table}
\end{table}
\section{Discussion}
\label{discussion}
The origin of CO rovibrational emission lines from protoplanetary disks has long been a puzzle. While it has been clear that
the lines are formed close to the star, given the high temperatures and densities required to excite the lines, the great variation
of the line shapes and excitation temperatures -- rotational and vibrational -- have evaded a unified explanation. Specifically, the
lines cannot be explained solely by thermal emission from simple Keplerian disks, except in a few cases. This was already
noted by \cite{Najita03}, but an unambiguous identification of the additional (to a Keplerian disk) radial flow pattern could not be determined. Now, with the addition of
spatial information, as offered by high resolution spectro-astrometry, we can present a more comprehensive picture of the dynamics of molecular gas
on AU-scales in protoplanetary disks.
\subsection{Size of the line emitting regions}
In \S\ref{Kepler_sources} we show that sources with double-peaked (Keplerian) profiles obey a size-luminosity relation with a power-law index of $\alpha=0.5$, as expected for emission
truncated at a radius of constant temperature. Critical ingredients in identifying this relation are the ability to distinguish Keplerian disk-dominated lines from the
wind-dominated and self-absorbed lines, as well as the use of the direct size measure offered by spectro-astrometry. Specifically, the astrometric amplitude is not
dependent on disk inclination, which otherwise makes it difficult to use the line width as an accurate proxy for size.
Another conclusion that can be drawn from the measured sizes of the emitting region from Keplerian sources is that a line origin in the accretion flow can be ruled out. Accretion flows are expected to generated double-peaked CO line profiles \citep{Najita03}. Funnel (along
magnetic field lines) accretion flows are expected to be launched from radii near, or within, the corotation radius \citep{Shu94}. This is, for T Tauri stars, located inside the dust sublimation radius at 2-10 stellar radii, and thus well within the observed CO radii of $\sim$100\,$R_*$. The Br$\alpha$ lines as observed in RNO 90 and HD 142527
have astrometric displacements much smaller than those of CO and evidently traces a different gas component, which may still be an accretion flow. Note that because the H~I lines trace a much wider range of infall velocities than the CO -- H~I line widths are 100-200\,$\rm km\,s^{-1}$ -- they are less affected by Keplerian rotation at the bottom of the flows than
the narrower CO lines \citep{Muzerolle98}. Consequently, they may not display a double-peaked profile even if formed as part of an accretion flow.
Is it expected that CO follows a relation that was developed for dust sublimation? One possibility is that dust shielding plays a significant role
in the survival of CO in the inner regions of disks. In this case, CO would only be able to survive at radii at which the radial column of dust is sufficiently high. This
is consistent with CO existing at radii larger than those of the inner dust rim. However, CO is also know to efficiently self-shield, even at low column densities of
$N(H_2)\sim 10^{18}\,\rm cm^{-2}$ \citep{Visser09}, although uncertainties are large at temperatures higher than a few 100\,K.
\subsection{Non-Keplerian motions and a unifying model for CO rovibrational lines}
We propose that most, if not all, disks have a molecular wind velocity component in addition to pure Keplerian rotation -- one that is effectively traced by
ro-vibrational CO lines, presumably in addition to lines from a wide range of other molecules with strong infrared emission bands. \cite{Najita03} argues against a wind based on
the fact that the lines are not blue-shifted and that very young (stage I) high accretion rate sources sometimes do not show rovibrational CO emission lines.
However, our wind model shows that a strong blueshift of the line emission is not necessary for the uncollimated slow disk winds constrained by angular momentum conservation.
While we do not specifically address the lack of strong CO lines toward some embedded young stellar objects, strong accretion may heat the disk mid-planes
to a point where the temperature inversion in the disk surface required for line emission is no longer possible. Furthermore,
we do find that the contribution of outflowing gas to the line emission is highly variable, possibly in relation to the level of accretion activity onto the central star.
While we have not sought to determine the wind launching mechanism, the data do suggest that the wind is uncollimated and slow ($|v_{\rm wind}|<v_{\rm Kepler}$). It is important to stress
that the wind should not be seen as entirely separate from the Keplerian disk, but rather as a modification. For instance, the
gas does not suddenly stop its Keplerian rotation even as it attains a radial velocity component. For low wind accelerations, it
may be difficult to determine that there is a difference at all from a pure Keplerian velocity field. As the wind is launched, the gas maintains its angular momentum
and slows down so much as it is pushed outwards that it may never escape the system, but re-accrete onto the disk at larger radii.
The mass-loss rates implied by our wind models are large enough that the inner disk gas can be completely redistributed during the lifetime of the disk. This may
limit the time available for planet formation, in line with ideas generally associated with photo-evaporative disk winds \citep{Alexander06}.
\subsection{Wind launching mechanism}
We will not discuss in detail how the molecular wind is launched. However, the requirement that it is uncollimated and slow likely
places significant physical constraints on the launching mechanism. For instance, a magnetic centrifugal wind might be expected to generate winds
that are much too fast as the gas in this case will be locked to the stellar rotation, known to be $v_{*}>v_{\rm Kepler}$ beyond the inner rim of the disk.
A thermal wind is much slower, but may require high ionization fractions and gas temperatures ($T_{\rm gas}\sim 10,000\,$K), which
could be inconsistent with the presence of abundant molecules with rotational temperatures of $\sim 1,000\,$K. The photo-evaporative wind models by
\cite{Alexander06} predict wind launch speeds that may even be slightly subsonic (5-10\,$\rm km\,s^{-1}$), consistent with
our observational finding. It can be noted that the evaporative winds predict a somewhat shallower exponent $p$ for the drop-off of the mass-loss rate $\dot{\Sigma}(w)$ with radius than the $p=7/2$
used here. Experimenting with our model indicates that good fits may also be found with $p=5/2$, although not as readily. Also noteworthy is that the models by \cite{Alexander06} predict significant blueshifts of wind-dominated [Ne~II] line peaks \citep[5-10\,$\rm km\,s^{-1}$][]{Alexander08}, which does not
appear in the CO data, at least compared to the ambient cloud velocity traced by cold absorption components.
\subsection{Implications for radial mixing}
The observational implication of slow gas flows above the canonical ``warm molecular layer'' of the disk may have important implications for
the transport of material in the disk. Specifically, the gas may never be accelerated to velocities allowing it to escape the disk, allowing it to fall back onto the disk at larger radii.
For flow rates of $10^{-9}\,M_{\odot}\rm \,yr^{-1}$, a wind lasting a few Myr will clearly be able to re-distribute a significant fraction, if not all, of the inner disk gas not accreted onto the central star. Similar ideas
where driving the development of early disk wind models, such as the X-wind \citep{Shu94} to explain the redistribution
of solids required to produce the crystalline dust grains in cometary material. It may be important to note that if the outflowing gas falls back onto the disk, there will be
a significant azimuthal velocity differential as sub-Keplerian material impacts the Keplerian disk below. Judging from Figure \ref{Wind}, the differential may be as high as 10-15\,$\rm km\,s^{-1}$
if the fall happens at 10\,AU, but less if the fall happens farther out in the disk. This is sufficient to generate shocks that may be
observable.
\subsection{Caveats, unknowns and new questions}
\label{Caveats}
Clearly, the parametrized wind model is not based on a detailed treatment of the underlying hydrodynamics and radiative transfer. However, it
does represent a phenomenological structure required to match clearly defined observables. It is therefore important to consider
the circumstances under which the physical and chemical structure of the wind may differ from the simple model.
One property of the wind model that does not match the observations is the rotational temperatures of the lines. The model predicts lower
temperatures than observed. However, this can likely be explained by the use of LTE level populations in thermal balance with the dust. The gas that forms the wind will be exposed to a strong IR and UV radiation field from the central star and innermost disk, and is therefore subject to fluorescence pumping as well as
collisional excitation by a gas that is heated and likely partly ionized by photo-electrons. Balancing this are the strong cooling terms offered by
the rovibrational molecular emission, in particular that of CO and water. While previous wind models \citep[e.g.,][]{Safier93} generally
predict very high wind temperatures (10,000 K) and fast winds ($\rm 100\,km\,s^{-1}$), the existence of significant molecular coolants may, in part of the wind flow, maintain
the required low temperatures and velocities. In particular the temperatures may be closer to those given by the
assumption of equilibrium with the stellar radiation field, as assumed in this paper, than the high temperatures implied by a purely
atomic gas. A future detailed heating-cooling balance calculation for the wind model is clearly
an important next step.
More detailed hydro-dynamic modeling is essential to fully assess the implications of slow disk winds. What is the wind launching mechanism?
Does the wind eventually fall back onto the disk and at which radii? How can the wind remain, at least in part, in a molecular form as the material is lofted to altitudes where
dust shielding low and the gas is exposed to a harsh UV radiation field? While a model of the chemistry of the molecular wind is a study in its own right, it can be noted that the
CO column densities required to generate the deep absorption lines seen in the self-absorbed (more inclined) sources are likely high enough to self-shield. Likewise, the
presence of water in the flow, at least in AS 205N -- see Figure \ref{SA_H2O} -- will provide additional shielding against UV photons for a range of molecular species \citep{Bethell09}.
\section{Conclusion}
\label{conclusion}
Using spectro-astrometry to image molecular gas at 0.1-10\,AU in a sample of protoplanetary disks has revealed an intriguing range of structure.
Some disks appear to be dominated by gas in Keplerian rotation about the central star, as expected, while others have a significant {\it slow} radial velocity component consistent with
a wide-angle disk wind. The basic observational evidence for a slow wind is the clear presence of {\it low velocity gas ($\sim 5\rm\,km\,s^{-1}$) within a few AU from the central star.}
It is difficult to generate low velocity gas deep in a potential well, but one simple way of doing so is via angular momentum conservation of an expanding, but initially Keplerian, flow.
While it has long been known that fast atomic winds were common in T Tauri stars, we now find that there is a significant molecular component as well. Further,
it appears that the observed molecular wind must be launched from the disk surface. It is not clear whether the gas in the wind reaches escape velocities, and
may therefore be re-accreted onto the disk at larger radii. The wind reproducing the CO line spectro-astrometry is much slower than that
predicted by X-wind theory, which reaches terminal velocities of several $100\,\rm km\,s^{-1}$ \citep{Shu94}, but is a much better match to photo-evaporative
flows that have poloidal velocities similar to the sound speed \citep[e.g.,][]{Alexander06}. The existence of disk winds with high mass-loss rates have significant implications for the
availability of material for planet formation in the PFZs of protoplanetary disks, and may limit the time available for planet formation. The potential large scale cycling of inner disk material as suggested by the low velocity of the winds will also influence the chemistry of planet-forming material, for instance by exposing a large fraction of the disk mass to the strong UV fields at high disk elevations. Future work will include
the development of a model for how slow {\it molecular} winds are launched, as well as a chemical model of the wind that can explain the survivability of,
at least, CO, H$_2$O and OH at low optical depths and at high elevations above the disk.
\acknowledgments{The authors are grateful to Colette Salyk and Ewine van Dishoeck for comments that improved the manuscript. Joanna Brown and Bill Dent are thanked for
obtaining some of our spectro-astrometric data. A very special thanks goes out to all the Paranal personnel that assisted with
our visitor observations, without whom this study would not have been possible. Support for KMP was provided by NASA through Hubble Fellowship grant \#01201.01
awarded by the Space Telescope Science Institute, which is operated by the Association of
Universities for Research in Astronomy, Inc., for NASA, under contract NAS 5-26555. This paper is based on observations made with ESO Telescopes at the Paranal Observatory
under program ID 179.C-0151.
}
\bibliographystyle{apj}
|
\section{Introduction}
\subsection{Statement of main results}
Let $A$, $B$, $C$ three complex vector spaces,
of dimension $a$, $b$, $c$ respectively.
A tensor $t\in A\otimes B\otimes C$ is said to have {\it rank} $k$
if there is a decomposition
$$t=\sum_{i=1}^k u_i\otimes v_i\otimes w_i$$
with $u_i\in A, v_i\in B, w_i\in C$ and the number of summands
$k$ is minimal. Such a decomposition is said to be {\it unique}
if for any other expression
$$t=\sum_{i=1}^k u'_i\otimes v'_i\otimes w'_i$$
there is a permutation $\sigma$ of $\{1,\ldots , r\}$ such that
$$u_i\otimes v_i\otimes w_i=u'_{\sigma(i)}\otimes v'_{\sigma(i)}\otimes w'_{\sigma(i)}\quad\forall i=1,\ldots, k.$$
When $t$ has a unique decomposition, the vectors $u_i\in A$,
$v_i\in B$, $w_i\in C$ can be {\it identified} uniquely from $t$,
up to scalars.
It is known that the set of tensors of rank $k$
consists of a dense subset of an irreducible
algebraic variety $S_k(Y)$, which is called the {\it $k$-th secant variety}
of the variety $Y$ of tensors of rank one.
This last variety is isomorphic to the (cone over the)
Segre product ${\mathbb{P}}(A)\times{\mathbb{P}}(B)\times{\mathbb{P}}(C)$.
The main result of our paper determines a bound for the rank,
in terms of the dimensions of the
vector spaces, which implies identifiability
\begin{Thm}\label{main}
Let $a\le b\le c$. Let $\alpha$, $\beta$ be maximal such
that $2^\alpha\le a$ and $2^\beta\le b$.
The general tensor $t\in A\otimes B\otimes C$ of rank $k$ has a
unique decomposition, if $k\le 2^{\alpha+\beta-2}$.
\end{Thm}
So if $a$, $b$ are both a power of $2$, then the general tensor of
rank $k$ has a unique decomposition if $k\le\frac{ab}{4}$.
In the general case, the inequality of the theorem can be written as
$k\le 2^{\left(\lfloor \log_2{a}\rfloor+\lfloor \log_2{b}\rfloor-2\right)}$.
Since $\frac{a+1}{2}\le 2^\alpha$ and
$\frac{b+1}{2}\le 2^\beta$, one can say that the unique
decomposition holds if $k\le (a+1)(b+1)/16$.
In our terminology, when the unique decomposition holds
for the general tensor of rank $k$, we will say that
the variety of tensors of rank one is {\it $k$-identifiable}.
Here the meaning of ``general'' is that, among tensors of rank $k$,
the ones which do not have a unique decomposition
consist in a set of zero measure, more specifically
in a proper subvariety of $S_k$.
In particular, the Theorem applies to ``cubic'' tensors.
The general tensor $t\in A\otimes A\otimes A$ of rank
$k$ has a unique decomposition if $k\le \frac{a^2}{16}$
(indeed, the Theorem provides a better bound, when
$a$ is close to a power of $2$).
Our bound is log-asymptotically sharp, in the following sense.
As explained in Proposition \ref{maxk}, one cannot
have a unique decomposition, when the rank exceeds a value $k_{max}=
k(a,b,c)$, which depends on $a,b,c$.
Then $\sup_c \frac{k(a,b,c)}{ab}$ is finite.
On the other hand, even for tensors of small size,
the result is not sharp. In the first cases, with the help of a computer,
we can improve Theorem \ref{main}.
Unique decomposition has been studied by several authors, and there
is a huge amount of literature, on this theme.
Let us remind that Strassen and Lickteig (\cite{Lick}) proved that
the general tensor $t\in A\otimes A\otimes A$ has rank
$\lceil \frac{a^3}{3a-2}\rceil$ for $a\neq 3$
and rank $5$ for $a=3$ (indeed, the case $a=3$ is known to be
{\it defective}, meaning that the corresponding $4$-secant variety
has dimension smaller than the expected one).
In this case, the aforementioned bound implies that, if
$a\geq 3$, then the generic
tensor of rank $k$ can have a unique decomposition only
if $k\le \lceil \frac{a^3}{3a-2}\rceil -1$.
The following theorem shows that this bound is almost always
achieved, for small $a$.
\begin{Thm}\label{main2}
The general tensor $t\in A\otimes A\otimes A$ of rank $k$
has a unique decomposition if $k\le k(a)$ where
$$\begin{array}{cc|rrrrrr|rrr}&a&2&3&4&5&6&7&8&9&10\\
\hline\\
&k(a)&2&3&5&9&13&18&22&27&32
\end{array}$$
\end{Thm}
\vskip.5cm
A more general list, which holds in the non cubic case, is given
in section \ref{fin}.
Comparing the previous table with the table
of the general rank (for $a> 3$, the general rank $-1$
is the best possible achievement),
and with Kruskal's result (see Proposition \ref{Kruskal}),
one can appreciate the improvement.
$$\begin{array}
{cc|rrrrrr|rrr}&a&2&3&4&5&6&7&8&9&10\\
\hline\\
\textrm{gen.rank\ } (a\neq 3)\ \textrm{\cite{Lick}}&\lceil \frac{a^3}{3a-2}\rceil&2&4&7&10&14&19&24&30&36\\
\hline\\
\textrm{Kruskal bound\ \cite{Kru}}&\lfloor\frac{3a-2}{2}\rfloor&2&3&5&6&8&9&11&12&14
\end{array}$$
\vskip.5cm
The more evident lack of uniqueness is when $a=4$ and $k=6$.
The case $a=4$ is particularly interesting due to the
models in phylogenetics \cite{AR,ERSS},
where a basis in ${\mathbb{C}}^4$ can be indexed by the nucleotids
$\{A, C, G, T\}$.
\begin{Thm}\label{main3}
The general tensor $t\in{\mathbb{C}}^4\otimes{\mathbb{C}}^4\otimes{\mathbb{C}}^4$ of rank
$6$ has exactly two decompositions.
\end{Thm}
It is interesting that the exception on uniqueness ($a=4$) holds
very close to the defective case $a=3$. This phenomenon is quite
general and it can be already encountered in the case of symmetric tensors.
\subsection{A few historical remarks}
In this subsection we sketch how our result fills in the literature.
The most celebrated result about uniqueness of decomposition of
tensors is due to Kruskal \cite{Kru}. It is often quoted in
terms of Kruskal's rank. A consequence of Kruskal's criterion
is the following statement, which applies to generic tensors (see
Corollary 3 in \cite{AMR}).
\begin{Prop}\label{Kruskal} ({\bf Kruskal's criterion})
The generic tensor $t\in A\otimes B\otimes C$ of rank $k$
has a unique decomposition if
$$ k \leq \frac 12 \left[\min(a,k)+\min(b,k)+\min(c,k)-2\right]$$
In the cubic case, the generic tensor $t\in A\otimes A\otimes A$
of rank $k$ has a unique decomposition if
$$ k\le \frac{3a-2}{2}$$
\end{Prop}
Kruskal's result is so important in the literature,
that recently there have been published (at least!) three different proofs
\cite{Land,R,St}.
De Lathauwer (\cite{Lat}) proves that the generic tensor
$t\in A\otimes B\otimes C$ of rank $k$ has a unique decomposition if
$ k\le c$ and $k(k-1)\le a(a-1)b(b-1)/2$.
Rhodes, in \cite{R} addresses explicitly, as a problem at the end of the
introduction, the need of sufficient conditions,
stronger than Kruskal's, that guarantee the uniqueness of
the decomposition, for generic tensors. Our Theorem \ref{main} gives a sufficient
condition which improves both Kruskal's and de Lauthawer's bounds.
The tensor decomposition we are looking for are called also
Candecomp or Parafac decompositions in the numerical literature.
Among recent surveys on the topic, see \S 3.2 in \cite{KB} and
Landsberg book \cite{Land0}, which tries to use a language
understandable by both the numerical and the geometrical communities.
From this point of view, one should also consider section 2 of
\cite{AMR}, an interesting bridge between the two worlds.
\subsection{Outline of the proof}
In a line, our technique consists in putting together the
inductive approach of \cite{AOP} with the tool of weak defectivity
developed in \cite{CC1} and \cite{CC2}.
We consider the projective space of tensors ${\mathbb{P}}(A\otimes B\otimes C)$.
In this space, the tensors of rank one give the Segre variety
${\mathbb{P}}(A)\times{\mathbb{P}}(B)\times{\mathbb{P}}(C)$.
Our geometric point of view consists in the use of the celebrated
Terracini's Lemma, which allows to study the identifiability of varieties,
using properties of their tangent spaces. We refer to \cite{CC1}
and \cite{CC2} for a more precise account of the theory behind.
A variety is called {\it tangentially $k$ weakly defective}
($k$-twd, see Definition \ref{twd}) if the span of the tangent
spaces at $k$ general points of $X$, is tangent also in some
other points.
It is a consequence of Terracini's Lemma that, if $X$ is $k$-not
twd, then the general tensor of rank $k$ has a unique decomposition.
So our aim is to prove the $k$-not twd of Segre varieties
$X={\mathbb{P}}(A)\times{\mathbb{P}}(B)\times{\mathbb{P}}(C)$.
The proof is performed by induction, by splitting
$A=A'\oplus A''$ and by specializing some points on the lower
dimensional Segre varieties ${\mathbb{P}}(A')\times{\mathbb{P}}(B)\times{\mathbb{P}}(C)$ and
${\mathbb{P}}(A'')\times{\mathbb{P}}(B)\times{\mathbb{P}}(C)$. It turns out that the
induction works if we prove a stronger statement, concerning
the so called $(k,p,q,r)$-weakly defectivity,
which is defined in section \ref{indsec}.
\subsection{Outline of the paper}
In section \ref{prel} we develop the basic notations on Segre varieties and
weak defectivity. At the end of this section we prove the cases $a\le 7$
of the Theorem \ref{main2}. Section \ref{indsec} contains the definition \ref{kpqr}
of $(k,p,q,r)$-defectivity and the inductive step (Prop. \ref{induz}).
At the end of this section we prove the remaining cases
of the Theorem \ref{main2}.
In the section \ref{pfs} we prove the Theorem \ref{main}.
In section \ref{fin} we prove the Theorem \ref{main3} and we give other
examples of small dimension. Also we expose a list of all the examples
of triple Segre product that we know when the uniqueness
for general tensors of a given rank does not hold.
In section \ref{many}, we show an extension of the
previous results to products of many factors.
\section{Preliminaries on Segre varieties}\label{prel}
Let $A,B,C$ be complex vector spaces, of dimension $a,b,c$ respectively.
Consider the product $X={\mathbb{P}}(A)\times{\mathbb{P}}(B)\times{\mathbb{P}}(C)$.
$X$ is naturally embedded, by means of the Segre map, into
${\mathbb{P}}^N$, where $N=abc-1$.
Sometimes, when there is no need to specify the vector spaces,
we will refer to the variety $X$ also as $\triplet {a-1}{b-1}{c-1}$.
Call $S^k(X)$ the $k$-th secant variety of $X$, defined as the closure
of the union of linear spans of $k$ general points in $X$.
\begin{Def}
$X$ is called $k$-identifiable if a \emph{general} element
in $S_k(X)$ has a unique expression as sum of $k$ elements in $X$.
From the tensorial point of view, this means that a general tensor
of type $a\times b\times c$ and rank $k$, can be written uniquely
(up to scalar multiplication) as a sum of $k$ decomposable tensors.
\end{Def}
\begin{Prop}\label{maxk} There is a maximal rank for which
the $k$-identifiability of tensors is possible, namely
$$k_{max}= \lfloor \frac {N+1}{\dim(X)+1} \rfloor= \lfloor
\frac{abc}{a+b+c-2} \rfloor.$$
\end{Prop}
\begin{proof}
For $k>k_{max}$, the abstract secant variety
$$ AbS^k(X)=\{(x_1,\dots,x_k,u)\in X^k\times{\mathbb{P}}^N:
u\in<x_1,\dots,x_k>\}$$
has dimension bigger than $N$, so that necessarily the general
$u\in S^k(X)$ belongs to infinitely many $k$-secant spaces.
\end{proof}
Our theoretical starting point is a criterion for $k$-identifiability,
which follows from the Terracini's Lemma,
which we will use under the following form (see e.g. \cite{CC1})
\begin{Lem} ({\bf Terracini}) Let $X$ be an irreducible variety and
consider a general point $u\in S_k(X)$. If $u$ belongs to the span
of points $x_1,\dots, x_k\in X$, then the tangent space to $S_k$ at $u$
is the span of the tangent spaces to $X$ at the points $x_1,\dots,x_k$.
\end{Lem}
Our criterion is the following:
\begin{Prop}\label{criterio}
Let $X\subset {\mathbb{P}}^N$ be a non-degenerate, irreducible variety
of dimension $n$. Consider the following statements:
(i) $X$ is $k$-identifiable
(ii) Given $k$ general points $x_1,\ldots x_k\in X$, then the
span $<T_{x_1}X,\ldots, T_{x_k}X>$
contains $T_xX$ only if $x=x_i$ for some $i=1,\ldots k$.
(iii) there exists a set of $k$ particular points
$x_1,\ldots x_k\in X$, such that the span
$<T_{x_1}X,\ldots, T_{x_k}X>$ contains $T_xX$ only if $x=x_i$
for some $i=1,\ldots k$.
Then we have (iii) $\Longrightarrow$ (ii) $\Longrightarrow$ (i).
\end{Prop}
\begin{proof} (iii) $\Longrightarrow$ (ii) follows
at once by semicontinuity.
Let us prove that (ii) $\Longrightarrow$ (i).
Take a general point $u\in S_k(X)$ and assume that $u$ belongs
to the span of points $x_1,\dots, x_k\in X$. By the generality of $u$,
we may assume that $x_1,\dots, x_k$ are general points of $X$.
If $u$ also belongs to the span of points $y_1,\dots, y_k\in X$,
with at least one of them, say $y_1$, not among the $x_i$'s,
then, by Terracini's Lemma, the span of the tangent spaces to $X$ at
the points $x_i$'s, which is the tangent space to $S_k(X)$ at $u$, also
contains the tangent space to $X$ at $y_1$. This contradicts (ii).
\end{proof}
Condition (ii) of the previous Proposition is strongly
related with the notion of $k$-weak defectivity.
In \cite{CC1}, C. Ciliberto and the first author give the following
definition: a variety $X$ is $k$-weakly defective if the general
\emph{hyperplane} which is tangent to $X$ at $k$ general points
$x_1,\dots,x_k$, is also tangent in some other point $y\neq
x_1,\dots, x_k$.
It is clear that a variety which does not satisfy condition (ii)
of the Proposition, is also $k$-weakly defective. However the
converse does not hold.
\begin{Ex} \label{nonbitangent}Consider the Segre product $X={\mathbb{P}}^1\times{\mathbb{P}}^2$.
It is classical (see e.g. Zak's Theorem on tangencies in \cite{zak})
that the tangent space at one point to a smooth variety is not
tangent elsewhere.
On the other hand, a general hyperplane tangent to $X$ at one point,
is also tangent along a line. Indeed, it is well known that
the dual variety of $X$ is not a hypersurface (see \cite{Ein}).
Thus $X$ is $1$-weakly defective.
\end{Ex}
For maintaining the consistency with all the previous
notation in this subject, we dare proposing the following:
\begin{Def}\label{twd} If $X$ satisfies condition (ii) of the previous
Proposition, we will say that $X$ is \emph{ $k$-not tangentially
weakly defective}. Otherwise, we say that $X$ is \emph{$k$-tangentially weakly defective}
($k$-twd, for short).
\end{Def}
We understand that the notation is becoming odd. However,
the increasing number of definitions is a phenomenon which
also occurs in the study of \emph{contact loci}, which
seems however helpful for applications to the Geometry
of secant varieties (see e.g. \cite{CC3}).
Weak defectivity has been intensively studied in \cite{CC1}.
Notice that when $X$ is $k$ weakly defective, then a general hyperplane
tangent to $X$ at general points $x_1,\dots,x_k$ is also
tangent along a positive dimensional variety. We do
not know if a similar phenomenon takes place also
for $k$-twd.
Relations between $k$-weak defectivity and $k$-twd
are probably stronger than expected, at least as far as
one is interested in $k$-identifiability. We do not develop
further this analysis.
Notice than, when we deal with inductive steps in the proofs,
we will need an even more complicated notion of
weak defectivity. Compare with Definition \ref{kpqr} below.
\smallskip
For our purposes, Proposition \ref{criterio} establishes that
$k$-not tangentially weakly defectivity implies
$k$-identifiability,
when $N\geq k(n+1)$.
\begin{Rmk}
Let us notice that, by Proposition \ref{maxk},
if $N+1<(k+1)n$, then $k$-identifiability is excluded.
Thus, the criterion of Proposition \ref{criterio} cannot be applied only for
at most one value of $k$, namely $k=(N+1)/(n+1)$,
which occurs only when $N+1$ is a multiple of $n+1$.
E.g., our criterion could not be applied to
study the $2$-identifiability of $\triplet 111$.
\end{Rmk}
Now we are already able to prove
the first cases of Theorem \ref{main2}.
{\it Proof of the Theorem \ref{main2} in case $a\le 7$.}
The proof is a straightforward application of Proposition \ref{criterio}.
A random choice of $k(a)$ points satisfies condition (iii) of
Proposition \ref{criterio}. Then $X$ is $k$-identifiable.
The Macaulay2 files which we used are available as ancillary files in the arXiv submission of this paper.
\begin{Rmk} More powerful computers and/or better suited algorithms will allow eventually to check the condition (iii) for larger values of $a$,
and we encourage experts in Numerical Algebraic Geometry in going further.
We stopped at $a=7$, because for $a=8$
our algorithm on a common PC consumed too much time and memory. In the next section we
show how the computation for larger values of $a$ can be reduced to other computations for smaller values of $a$.
\end{Rmk}
\section{The inductive statement}\label{indsec}
The inductive criterion makes use of the fact that
if $x=u\otimes v\otimes w$ is a point of
$X={\mathbb{P}}(A)\times{\mathbb{P}}(B)\times{\mathbb{P}}(C)$,
then the tangent space $T_xX$ is the projectification
of the linear space $A\otimes v\otimes w +
u\otimes B\otimes w + u\otimes v\otimes C$.
The idea is to fix two linear subspaces $A',A''$ of $A$,
such that $A=A'\oplus A''$, then split the set of $k$ points in two
subsets and specialize them to the two spaces
${\mathbb{P}}(A')\times{\mathbb{P}}(B)\times{\mathbb{P}}(C)$ and ${\mathbb{P}}(A'')\times{\mathbb{P}}(B)\times{\mathbb{P}}(C)$.
Then, the implication (iii) $\Longrightarrow$ (i) of
Proposition \ref{criterio} suggests that one could play induction.
Unfortunately, the situation is a little bit more complicated,
since one cannot translate condition (ii) of Proposition \ref{criterio}
into the analogous condition on lower-dimensional spaces.
Instead, following the idea of \cite{AOP} (Theorem 3.4)
(suggested also from the Splitting Method of \cite{BCS}),
we need a more elaborated condition.
\begin{Def}\label{kpqr} A triple product
$X={\mathbb{P}}(A)\times{\mathbb{P}}(B)\times{\mathbb{P}}(C)$
is called {\it $(k,p,q,r)$-not weakly defective} if:
for $k$ general points $x_1,\ldots x_k\in X$,
for $p$ general points $u_i\in{\mathbb{P}}(B)\times{\mathbb{P}}(C)$,
for $q$ general points $v_i\in{\mathbb{P}}(A)\times{\mathbb{P}}(C)$,
for $r$ general points $w_i\in{\mathbb{P}}(A)\times{\mathbb{P}}(B)$,
then the span of
$T_{x_i}X$, $A\otimes u_i$, $B\otimes v_i$, $C\otimes w_i$
contains $T_xX$ if and only if $x=x_i$, for some $i=1,\ldots k$.
Otherwise $X$ is called {\it $(k,p,q,r)$-weakly defective}.
Clearly, $(k,0,0,0)$ weak defectivity
coincides with $k$-twd.
\end{Def}
\begin{Rmk}
We will often use the computer algorithm,
available in our arXiv submission,
to prove that some triple Segre product is
$(k,p,q,r)$-not weakly defective.
For instance, the algorithm shows that
$\triplet 222$ is $(1,2,1,1)$-not and $(2,1,1,1)$-not weakly defective.
This is rather interesting, because
${\mathbb{P}}^2\times{\mathbb{P}}^2\times{\mathbb{P}}^2$ is $3$-defective.
\end{Rmk}
\begin{Ex} Consider $A$, $B$, $C$, all of dimension $2$ with basis
$\{u_1,u_2\}$, $\{v_1,v_2\}$, $\{w_1, w_2\}$.
Then $T_{u_1v_1w_1}+u_2v_2C=T_{u_2v_2w_1}+u_1v_1C$.
This shows that ${\mathbb{P}}^1\times{\mathbb{P}}^1\times{\mathbb{P}}^1$ is $(1,0,0,1)$
weakly defective. Nevertheless, $T_{u_1v_1w_1}+u_2v_2C$
has the expected (affine dimension) $6$ and it does not fill the ambient space.
\end{Ex}
\begin{Rmk}\label{ovv} (a) With the previous notation, by semicontinuity
it is clear that when $X$ is $(k,p,q,r)$-not weakly defective,
then it is also $(k,p,q,r)$-not weakly defective,
whenever $(k',p',q',r')\leq (k,p,q,r)$, in the strict ordering.
(b) By semicontinuity, $X$ is $(k,p,q,r)$-not weakly defective
whenever one gets that for {\it particular} sets of points
$\{x_i\}$, $\{u_i\}$, $\{v_i\}$ and $\{w_i\}$ as above,
then the span of
$T_{x_i}X$, $A\otimes u_i$, $B\otimes v_i$, $C\otimes w_i$
contains $T_xX$ if only if $x=x_i$, for some $i=1,\ldots k$.
(c) By Proposition \ref{criterio}, one gets soon that
$(k,0,0,0)$-not weakly defective implies $k$-identifiable.
\end{Rmk}
We will often apply the following reduction step:
\begin{Lem} \label{riduz}
Assume that $\triplet {a-1}{b-1}{c-1}$ is $(k,p,q,r)$-not
weakly defective. Then $\triplet{a'}{b'}{c'}$ is $(k,p,q,r)$-not
weakly defective, for any triple $(a',b',c')>(a-1,b-1,c-1)$
(in the strict ordering).
\end{Lem}
\begin{proof} We need just to prove the statement for
$(a',b',c')=(a,b-1,c-1)$. Write
$X=\triplet a{b-1}{c-1}={\mathbb{P}}(A')\otimes{\mathbb{P}}(B)\otimes{\mathbb{P}}(C)$
so that $\dim(A')=a+1$.
Assume that $X$ is $(k,p,q,r)$-weakly defective.
Thus, for $k$ general points $x_1,\dots, x_k\in X$,
$p$ general points $u_i\in{\mathbb{P}}(B)\times{\mathbb{P}}(C)$,
$q$ general points $v_i\in{\mathbb{P}}(A')\times{\mathbb{P}}(C)$,
$r$ general points $w_i\in{\mathbb{P}}(A')\times{\mathbb{P}}(B)$,
then the span $\Lambda$ of the tangent spaces to $X$ at the $x_i$'s
and the spaces $A'\otimes u_i$, $B\otimes v_i$,
$C\otimes w_i$, is also tangent in another point $y$.
Take a general point $P=(u,v,w)\in
\triplet {a}{b-1}{c-1}$ and consider the projection $\pi$ of $X$ from
$L= u\otimes B\otimes C$. The image of the projection
is $Y={\mathbb{P}}(A)\otimes{\mathbb{P}}(B)\otimes{\mathbb{P}}(C)$
where $A\subset A'$ has codimension $1$.
Furthermore, by the generality of
$P$, $L$ does not meet $\Lambda$, as well as
any line spanned by $y,x_i$.
It follows that the span of the tangent spaces to $Y$
at the general points $\pi(x_1), \dots , \pi(x_k)$
and containing the spaces
$A\otimes\pi(u_i)$, $B\otimes\pi(v_i)$,
$C\otimes\pi(w_i)$ is
also tangent in another point $\pi(y)$. Thus $Y$ is $(k,p,q,r)$-weakly defective.
By induction, we get a contradiction.
\end{proof}
Now we are ready to state and prove our inductive criterion.
Let $X'={\mathbb{P}}(A')\times{\mathbb{P}}(B)\times{\mathbb{P}}(C)$,
$X''={\mathbb{P}}(A'')\times{\mathbb{P}}(B)\times{\mathbb{P}}(C)$.
Note that $A\otimes B\otimes C=\left(A'\otimes B\otimes C\right)\oplus
\left(A''\otimes B\otimes C\right)$.
Denote by $\pi'$ and $\pi''$ the two projections.
\begin{Prop}\label{induz} ({\bf Inductive Step.})
Assume that $X'$ is $(k_1,p+k_2,q_1,r_1)$-not weakly defective
and $X''$ is $(k_2,p+k_1,q_2,r_2)$-not weakly defective.
Then $X$ is $(k_1+k_2,p,q_1+q_2,r_1+r_2)$-not weakly defective.
\end{Prop}
\begin{proof} We specialize $k_1+k_2$ points on $X$ in order
that $k_1$ of them belong to $X_1$ and $k_2$ of them belong to $X_2$ .
Let $x_1,\ldots x_{k_1}\in X'$ and $y_1,\ldots y_{k_2}\in X''$.
Let $A\otimes\tilde v_i\otimes\tilde w_i$ for
$i=1,\ldots p$, be subspaces.
We specialize $q_1+q_2$ points in ${\mathbb{P}}(A)\times{\mathbb{P}}(C)$
in order that the first $q_1$ of them belong to ${\mathbb{P}}(A')\times{\mathbb{P}}(C)$
and the last $q_2$ of them belong to ${\mathbb{P}}(A'')\times{\mathbb{P}}(C)$ .
Call $Q_1$ the span of the first $q_1$ spaces $Bv_i$ and
$Q_2$ the span of the last $q_2$ spaces $Bv_i$.
We specialize $r_1+r_2$ points in ${\mathbb{P}}(A)\times{\mathbb{P}}(B)$ in order that
the first $r_1$ of them belong to ${\mathbb{P}}(A')\times{\mathbb{P}}(B)$ and the last
$r_2$ of them belong to ${\mathbb{P}}(A'')\times{\mathbb{P}}(B)$.
Call $R_1$ the span of the first $r_1$ spaces $Cw_i$
and $R_2$ the span of the last $r_2$ spaces $Cw_i$.
We want to prove that $T=T_{x_1}X+\ldots +T_{x_{k_1}}X+T_{y_1}X+
\ldots +T_{y_{k_2}}X+Q_1+Q_2+R_1+R_2+A\otimes\tilde v_1\otimes
\tilde w_1+\dots +A\otimes\tilde v_p\otimes
\tilde w_p$,
is tangent to $X$ only at $x_1,\ldots x_{k_1},y_1,\ldots y_{k_2}$.
Let $T_xX\subset T$, with $x=u\otimes v\otimes w$.
Then $\pi_1(T_xX)\subset\pi_1(T)$.
Let $u=u'+u''$, $y_j=u''_j\otimes v''_j\otimes w''_j$, $j=1,\dots,k_2$.
At least one among $u'$ and $u''$ is non zero, so let's
assume $u'\neq 0$. Then we get
$\pi_1(T_xX)=A'\otimes v\otimes w+u'\otimes B\otimes w+u'\otimes
v\otimes C$ while $\pi_1(T)=T_{x_1}X'+\ldots +T_{x_{k_1}}X'
+A'\otimes v''_j\otimes w''_j+\ldots + A'\otimes\tilde v_i
\otimes \tilde w_i+ \ldots+Q_1+R_1$ (with $i=1,\dots,p$).
By the assumption that $X'$ is $(k_1,p+k_2,q_1,r_1)$-not weakly
defective it follows that $u'\otimes v\otimes w$ is one among $x_i$.
If also $u'' \neq 0$ the same argument shows that $u''\otimes v\otimes w$
is one among $y_i$, which is a contradiction. Then $u''=0$,
that is $x=u'\otimes v\otimes w$ is one among $x_i$.
It follows that $X$ is $(k_1+k_2,p,q_1+q_2,r_1+r_2)$-not
weakly defective, as we wanted.
\end{proof}
The inductive procedure stops eventually
when we find some condition on weak defectivity, which
does not hold. This does not means, in general, that
our starting example was not $k$-identifiable, but merely
that we specialized the points too much, in order
to expect a meaningful answer.
\vskip 0.8cm
{\it Proof of Theorem \ref{main2} in cases $a=8, 9, 10$.}
In case $a=8$ we start with $22$ points and we want to apply iteratively the
Proposition \ref{induz}.
Splitting one $8$-dimensional vector space of the product
in a direct sum of two $4$-dimensional spaces, one sees that the
$(22,0,0,0)$-not weak defectivity of $\triplet777$
follows if one knows that $\triplet377$ is
$(11,11,0,0)$-not weakly defective. Repeating the procedure with
the second factor, everything reduces to prove
that $\triplet337$ is $(5,7,6,0)$-not weakly defective and
$(6,4,5,0)$-not weakly defective.
The first statement reduces to show that
$\triplet333$ is $(3,3,3,2)$-not weakly defective and
$(2,4,3,3)$-not weakly defective. These statements have finally a reasonable size
and can be checked with a random choice of points with our Macaulay2 algorithm.
The last statement reduces to show that
$\triplet333$ is $(3,2,3,3)$-not weakly defective and
$(3,2,2,3)$-not weakly defective, which follows from the above check and by the Remark
\ref{ovv} (a).
In the case $a=9$ we start with $27$ points and we split the $9$ dimensional space in {\it three} $3$-dimensional summands. The inductive step is better explained by the following table
$$
\begin{array}{ccccccccc}
a & b & c & & k & & p & q & r \\
9&9&9&&27&&0&0&0\\
3&9&9&&9&&18&0&0\\
3&3&9&&3&&6&6&0\\
3&3&3&&1&&2&2&2\\
\end{array}$$
The last statement can be checked again with Macaulay2.
The $a=10$ case starts as follows
$$
\begin{array}{ccccccccc}
a & b & c & & k & & p & q & r \\
10&10&10&&32&&0&0&0\\
5&10&10&&16&&16&0&0\\
5&5&10&&8&&8&8&0\\
5&5&5&&4&&4&4&4\\
\end{array}$$
The second statement reduces to show that
$\triplet144$ is $(1,7,2,2)$-not weakly defective
and
$\triplet244$ is $(3,5,2,2)$-not weakly defective.
Both these statements can be checked with Macaulay2.
This concludes the proof.
\section{Proof of Theorem \ref{main}}\label{pfs}
In order to use the inductive step, we need a starting point.
\begin{Lem}\label{start}
$\triplet 111$ is $(1,0,0,0)$-not and $(0,1,1,1)$-not weakly defective.
\end{Lem}
\begin{proof}
The first fact is true for any smooth variety, see Example \ref{nonbitangent}.
For the second one, we consider $X={\mathbb{P}}(A)\times{\mathbb{P}}(B)\times{\mathbb{P}}(C)$ where $A$, $B$, $C$ have all dimension $2$
and we choose basis $A=\langle a_0, a_1\rangle$, $B=\langle b_0, b_1\rangle$, $C=\langle c_0, c_1\rangle$.
Then, without loss of generality, we may consider the span $T=A\otimes b_0\otimes c_0+a_0\otimes B\otimes c_1+a_1\otimes b_1\otimes C$.
In the monomial basis of $A\otimes B\otimes C$ this span contains all the monomials with the only exception of $a_0\otimes b_1\otimes c_0$
and $a_1\otimes b_0\otimes c_1$. Then, a vector $v=\sum x_{ijk}a_i\otimes b_j\otimes c_k$,
belongs to $X\cap {\mathbb{P}}(T)$ if all the $2\times 2$-minors of the two following flattening matrices vanish
$$\left[\begin{array}{rrrr}
x_{000}&x_{001}&x_{100}&0\\
0&x_{011}&x_{110}&x_{111}
\end{array}\right]\qquad\qquad
\left[\begin{array}{rrrr}
x_{000}&0&x_{100}&x_{110}\\
x_{001}&x_{011}&0&x_{111}
\end{array}\right]$$
A straightforward check on the minors shows that $X\cap {\mathbb{P}}(T)$ consists of the following
six lines in the $5$-dimensional space ${\mathbb{P}}(T)=\{x_{010}=x_{101}=0\}$
(Macaulay2 can be helpful at this step)
$r_0=V(x_{001},x_{000},x_{100},x_{110})$
$r_1=V(x_{000},x_{100},x_{110},x_{111})$
$r_2=V(x_{100},x_{110},x_{111},x_{011})$
$r_3=V(x_{110},x_{111},x_{011},x_{001})$
$r_4=V(x_{111},x_{011},x_{001},x_{000})$
$r_5=V(x_{011},x_{001},x_{000},x_{100})$
which have the property that, for $i\neq j$
$$r_i\cap r_j=\left\{\begin{array}{cl}
\textrm{one point} &\textrm{if\ }i=j+1, j-1\textrm{\ mod\ }6\\
\emptyset&\textrm{otherwise}
\end{array}\right.$$
It follows that ${\mathbb{P}}(T)$ is not tangent anywhere,
because the tangent space at a point meets $X$ in three concurrent lines.
This proves that $X$ is $(0,1,1,1)$-not weakly defective.
\end{proof}
\begin{Rmk}
We will use affine spaces whose dimension is a power of $2$, as
well as sets of points or subspaces whose number is
expressed in terms of powers of $2$, essentially because they allow
the following recursive application of Lemma \ref{induz}:
Assume we want to prove that
$\triplet{2^\alpha-1}{2^\beta-1}{2^\gamma-1}={\mathbb{P}}(A)\otimes
{\mathbb{P}}(B)\otimes{\mathbb{P}}(C)$ is $(2x,2^u, 2^v, 2^w)$-not
weakly defective. Then, by splitting the first linear space $A$ in a
direct sum of two subspaces of dimension $2^{\alpha-1}$ and balancing
the splitting of the number of points and linear spaces, by
Proposition \ref{induz} it is sufficient to prove that
$\triplet{2^{\alpha-1}-1}{2^\beta-1}{2^\gamma-1}$ is
$(x,2^u, 2^{v-1}, 2^{w-1})$-not weakly defective.
We will use this trick so often, in the arguments below.
\end{Rmk}
The final statement will be that, if we order the dimensions
so that $1\leq \alpha\leq \beta\leq \gamma$, then
$X=\triplet{2^\alpha-1}{2^\beta-1}{2^\gamma-1}$ is $(k,0,0,0)$-not weakly defective,
for $k\leq 2^{\alpha+\beta-2}.$
Before showing this fact, we need a series of lemmas.
\begin{Prop} Assume that $X=\triplet{2^\alpha-1}{2^\beta-1}{2^\gamma-1}$ is
not $(k,0,0,0)$-not weakly defective.
Then also $X'=\triplet{2^\alpha-1}{2^\beta-1}{2^{\gamma}}$ is $(k,0,0,0)$-not
weakly defective.
\end{Prop}
\begin{proof}
By Lemma \ref{riduz}.
\end{proof}
So, in order to prove Theorem \ref{main}, we can reduce ourselves to
the case $\beta=\gamma$, $k=2^{\alpha+\beta-2}$.
\begin{Lem}\label{zero} Take $X=\triplet{2^{a_1}-1}{2^{a_2}-1}{2^{a_3}-1}$, with
$a_1,a_2,a_3\geq 1$. Pick non-negative integers $u_1,u_2, u_3$
such that $u_i\leq a_j+a_k-2$, whenever $\{i,j,k\}=\{1,2,3\}$.
Then $X$ is $(0,2^{u_1},2^{u_2},2^{u_3})$-not weakly defective.
\end{Lem}
\begin{proof} We make induction on the sum $a_1+a_2+a_3$.
If $a_1=a_2=a_3=1$, then the numerical conditions imply that
$u_1=u_2=u_3=0$ and the conclusion follows from the fact that
$\triplet 111$ is $(0,1,1,1)$-not weakly defective, which holds by
Lemma \ref{start}.
Assume $a_1>1$ and split the first projective space in a sum of
two spaces of dimension $2^{a_1-1}$. Then there are three
possibilities:
(1) Assume $u_2=u_3=0$. Then, by using Lemma \ref{induz}, the claim
reduces to prove that $\triplet{2^{a_1-1}-1}{2^{a_2}-1}{2^{a_3}-1}$ is
$(0,2^{u_1}, 1, 1)$-not weakly defective and it is
$(0,2^{u_1}, 0, 0)$-not weakly defective. The second condition
is contained in the first. Since $a_1>1$, the six numbers
$a_1-1, a_2, a_3, u_1, 0 , 0$ fulfill the numerical inequalities
of the statement. Hence the claim follows by induction, in this case.
(2) Assume $u_3>u_2=0$. Then the claim
reduces to prove that $\triplet{2^{a_1-1}-1}{2^{a_2}-1}{2^{a_3}-1}$ is
$(0,2^{u_1}, 1, 2^{u_3-1})$-not weakly defective and it is
$(0,2^{u_1}, 0, 2^{u_3-1})$-not weakly defective. The second condition
is contained in the first. One checks that the six numbers
$a_1-1, a_2, a_3, u_1, 0 , u_3-1$ fulfill the numerical inequalities
of the statement. Hence the claim follows by induction.
(3) Assume $u_2,u_3>0$. Then the claim
reduces to prove that $\triplet{2^{a_1-1}-1}{2^{a_2}-1}{2^{a_3}-1}$ is
$(0,2^{u_1}, 2^{u_2-1}, 2^{u_3-1})$-not weakly defective.
One checks that the six numbers
$a_1-1, a_2, a_3, u_1, u_2-1 , u_3-1$ fulfill the numerical inequalities
of the statement. Hence the claim follows by induction.
\end{proof}
\begin{Lem}\label{uno} Take $X=\triplet{2^{a_1}-1}{2^{a_2}-1}{2^{a_3}-1}$, with
$a_1,a_2,a_3\geq 1$. Pick non-negative integers $u_1,u_2, u_3$
such that $u_i\leq a_j+a_k-2$, whenever $\{i,j,k\}=\{1,2,3\}$.
Then $X$ is $(1,2^{u_1}-1,2^{u_2}-1,2^{u_3}-1)$-not weakly defective.
\end{Lem}
\begin{proof} We make induction on the sum $a_1+a_2+a_3$.
If $a_1=a_2=a_3=1$, then the numerical conditions imply that
$u_1=u_2=u_3=0$ and the conclusion follows from the fact that
$\triplet 111$ is $(1,0,0,0)$-not weakly defective (Lemma \ref{start}).
Assume $a_1>1$ and split the first projective space in a sum of
two spaces of dimension $2^{a_1-1}$. Then there are three
possibilities:
(1) Assume $u_2=u_3=0$. Then, by using Lemma \ref{induz}, the claim
reduces to prove that $\triplet{2^{a_1-1}-1}{2^{a_2}-1}{2^{a_3}-1}$ is
$(0,2^{u_1}, 1, 1)$-not weakly defective and it is
$(1,2^{u_1}-1, 0, 0)$-not weakly defective. The first condition follows
by the previous Lemma \ref{zero}. For the second condition,
notice that since $a_1>1$, the the six numbers
$a_1-1, a_2, a_3, u_1, 0 , 0$ fulfill the numerical inequalities
of the statement (and $0=2^0-1$).
Hence the claim follows by induction, in this case.
(2) Assume $u_3>u_2=0$. Then the claim
reduces to prove that $\triplet{2^{a_1-1}-1}{2^{a_2}-1}{2^{a_3}-1}$ is
$(0,2^{u_1}, 1, 2^{u_3-1})$-not weakly defective and it is
$(1,2^{u_1}-1, 0, 2^{u_3-1}-1)$-not weakly defective.
The first condition follows by the previous Lemma. The second condition
follows by induction, since one checks that the six numbers
$a_1-1, a_2, a_3, u_1, 0 , u_3-1$ fulfill the numerical inequalities
of the statement.
(3) $u_2,u_3>0$. Then the claim
reduces to prove that $\triplet{2^{a_1-1}-1}{2^{a_2}-1}{2^{a_3}-1}$ is
$(0,2^{u_1}, 2^{u_2-1}, 2^{u_3-1})$-not weakly defective and it is
$(1,2^{u_1}-1, 2^{u_2-1}-1, 2^{u_3-1}-1)$-not weakly defective.
One checks that the numerical conditions in the statement are still
fulfilled, by the six numbers $a_1-1, a_2, a_3, u_1, u_2-1 , u_3-1$.
Hence the claim follows by induction.
\end{proof}
Now we are ready to prove:
\begin{Thm}\label{premain}
$X=\triplet {2^\alpha-1}{2^\beta-1}{2^\beta-1}$ is $(k,0,0,0)$-not
weakly defective, for $k\leq 2^{\alpha+\beta-2}$.
\end{Thm}
\begin{proof}
Write $\alpha+\beta-2=2p+e$, where $e$ is the remainder.
Now we start our reduction.
($A_1$) One can split the
vector space in the middle as a sum of two spaces of
dimension $2^{\beta-1}$. By using Lemma \ref{induz}, it turns out that
$X$ is $(2^{\alpha+\beta-2},0,0,0)$-not weakly defective
when $\triplet {2^\alpha-1}{2^{\beta-1}-1}{2^\beta-1}$ is
$(2^{\alpha+\beta-3},0,2^{\alpha+\beta-3},0)$-not weakly defective.
($A_2$) Splitting now the third vector space as a sum of
two spaces of dimension $2^{\beta-1}$, and using the Lemma,
this reduces to prove that
$\triplet {2^\alpha-1}{2^{\beta-1}-1}{2^{\beta-1}-1}$ is
$(2^{\alpha+\beta-4},0,2^{\alpha+\beta-4},2^{\alpha+\beta-4})$-not
weakly defective.
($A_3$) Now repeat the procedure, splitting the space in the middle:
Everything reduces to prove that
$\triplet {2^\alpha-1}{2^{\beta-2}-1}{2^{\beta-1}-1}$ is
$(2^{\alpha+\beta-5},0,2^{\alpha+\beta-4}+2^{\alpha+\beta-5},
2^{\alpha+\beta-5})$-not weakly defective.
Now split again the third vector space, and
repeat the steps. At the end of the $(\alpha+\beta-2)$-th step, after
the computation, we find out that we need just to prove that
$\triplet{2^\alpha-1}{2^{\beta-p-e-1}}{2^{\beta-p-1}}$ is $(1,0,
\sum_{i=0}^{p+e-1} 2^i, \sum_{i=0}^{p-1} 2^i)$-not weakly defective.
Notice that all these steps can be performed, because $\beta-p\geq
\beta-p-e\geq 1$. Indeed we have $\alpha\leq \beta$, thus
$2\beta-2\geq 2p+e$, hence $2\beta\geq 2p+e+2>2p+2e$.
Now, $\sum_{i=0}^{p+e-1} 2^i=2^{p+e}-1$ while $\sum_{i=0}^{p-1} 2^i=2^p-1$.
Moreover
\begin{gather} \nonumber
p+e \leq \alpha+(\beta-p)-2 \qquad \mbox{ since } 2p+e=\alpha+\beta-2 \\
\nonumber p\leq \alpha+(\beta-p-e)-2 \qquad \mbox{ since }
2p = \alpha+\beta-e-2. \end{gather}
Thus we may apply Lemma \ref{uno}, and see that
$\triplet{2^\alpha-1}{2^{\beta-p-e-1}}{2^{\beta-p-1}}$ is $(1,0,2^{p+e}-1, 2^p-1)$-not
weakly defective. The result is settled.
\end{proof}
When $\alpha=\beta$, i.e. when the product is balanced, we find that $X$
is $k$-identifiable for $k\leq 2^{2\alpha-2}$.
\smallskip
\noindent{\it Proof of Theorem \ref{main}} Fix $\alpha,\beta$
maximal such that $2^\alpha\leq a$ and $2^\beta\leq b$.
Then, by the previous Theorem, $\triplet{2^\alpha-1}{2^\beta-1}{2^\beta-1}$
is $(k,0,0,0)$-not weakly defective, for $k\leq 2^{\alpha+\beta-2}=
2^\alpha2^\beta/4$. Thus also ${\mathbb{P}}(A)\otimes{\mathbb{P}}(B)\otimes{\mathbb{P}}(C)=
\triplet {a-1}{b-1}{c-1}$ is $(k,0,0,0)$-not
weakly defective, for $k\leq 2^\alpha2^\beta/4$.
The conclusion follows.
\hfill\qed
\vskip 0.5cm
Comparing our result with the maximal $k$ for which the
identifiability of ${\mathbb{P}}(A)\otimes{\mathbb{P}}(B)\otimes{\mathbb{P}}(C)$ makes sense, i.e.
$$k_{max}= \lfloor \frac{abc}{a+b+c-2} \rfloor.$$
(see Proposition \ref{maxk}) and considering that
$ab/3\leq k_{max}\leq ab$,
we see that the bound in the Theorem is, at least log-asymptotically,
sharp, as explained in the Introduction.
In any events, it improves Kruskal's bound for identifiability.
\begin{Rmk} In principle, there are no obstructions in repeating the
argument of Theorem \ref{premain}, when we substitute powers of $2$
with powers of any other integer $p>1$. The final statement is:
\noindent{\it $X=\triplet{p^\alpha-1}{p^\beta-1}{p^\beta-1}$ is
$(k,0,0,0)$-not weakly defective, for $k\leq p^{\alpha+\beta-2}$.}
The proof is achieved very similarly, by splitting, step by step,
a vector space of dimension $p^n$ into $p$ spaces of dimension $p^{n-1}.$
(see e.g. the case $a=9$ in the proof of Theorem \ref{main2}).
We can use this statement, instead of
Theorem \ref{premain}, in the proof of Theorem \ref{main},
obtaining another bound which implies $k$-identifiability.
In most cases, however, the new bound is weaker than the one of the Theorem
\ref{main}. On the other hand, in some specific case, typically
when powers of $3$ are involved, it can be stronger.
To give an example, let us consider $X=\triplet{26}{26}{26}$.
Using Theorem \ref{main}, we obtain $k$-identifiability for
$k\leq 2^{4+4-2}=64$. Using powers of $p=3$, instead, we
get $k$-identifiability for $k\leq 3^{3+3-2}=81$. It is an improvement,
but still a long way from $k_{max}=249$.
\end{Rmk}
\section{Some examples in low dimension}
\label{fin}
In this section, we study the $k$-identifiability of Segre products
$X={\mathbb{P}}(A)\otimes{\mathbb{P}}(B)\otimes{\mathbb{P}}(C)$, when the dimensions $a,b,c$ are small.
We also provide a proof for Theorem \ref{main3}.
\smallskip
\noindent{\it Proof of Theorem \ref{main3}}.
Consider $X=\triplet333$. This product is
$5$-identifiable, by Kruskal's criterion.
On the other hand, accordingly with Proposition \ref{maxk},
one may ask about the $6$-identifiability of $X$.
We are able to prove that $X$ is {\it not} $6$-identifiable,
and the general point in $S^6(X)$ sits in exactly two $6$-secant,
$5$-planes. From the tensorial point of view,
this means that a general $4\times 4 \times 4$ tensor of rank $6$,
can be written as a sum of $6$ decomposable tensors in exactly
$2$ ways (up to scalar multiplication and permutations).
The reason relies in the fact that through $6$ general points
$x_1,\dots,x_6$ of $X=\triplet{3}{3}{3}$, one can draw an elliptic normal
curve $\Gamma$ of degree $12$, which spans a projective space $L={\mathbb{P}}^{11}$,
containing the linear span of $x_1,\dots,x_6$. So, a general point
$u\in S^6(X)$ lies in a linear space $L$ spanned by an elliptic normal
curve $\Gamma\subset X$. By \cite{CC2}, Proposition 5.2, it is known
that $\Gamma$ has $6$-secant order $2$, i.e. there are exactly
two $5$-planes, $6$-secant to $\Gamma$, inside $L$. By \cite{CC2}
Proposition 2.4, if we prove that $\Gamma$ coincides with the contact
locus of a general $6$-tangent hyperplane,
also $X$ must have $6$-secant order equal to $2$.
This last fact can be checked by our Macaulay2 algorithm.
Unfortunately, the existence of an elliptic normal curve, of degree $12$,
passing through $6$ randomly chosen points of $X$, gives only
a probabilistic argument for the existence of such a curve passing through
$6$ general points of $X$. To overcome this problem,
we offer the following theoretical argument.
Consider the projections $z_{i1},\dots,z_{i6}$ of $x_1,\dots,x_6$,
into the $i$-th copy of ${\mathbb{P}}^3$, so that $z_{i1},\dots,z_{i6}$ are general
points of ${\mathbb{P}}^3$. Normal elliptic curves $C$ passing through the
$6$ points of ${\mathbb{P}}^3$ are given by pairs of quadrics through
the points, so they are parametrized by the Grassmannian $G$ of lines
in the space ${\mathbb{P}}^3$ of quadrics through $z_{i1},\dots,z_{i6}$.
In order that three normal elliptic curves $C,C',C''$
in the three copies of ${\mathbb{P}}^3$,
correspond to the same abstract curve, they need to
differ by an element of $PGL(3)$.
So, once we have $C$ ($4$ parameters), we can choose
$\phi,\psi\in PGL(3)$ for the two remaining maps $C\to{\mathbb{P}}^3$
(thus a total of $4+15+15=34$ parameters). On the other hand,
we need to impose that $\phi(C)=C'$ (resp $\psi(C)=C''$)
pass through $z_{21},\dots,z_{26}$ (resp. $z_{31},\dots,z_{36}$).
Since each point imposes $2$ conditions,
we get a total of $24$ algebraic conditions on the $34$ parameters.
Moreover, if we want that after
this correspondence, $C,C',C''$ are projection of the same curve
passing through $x_1,\dots,x_6$, we also need that the projectivity
$\phi: C\to C'$ (resp. $\psi: C\to C''$), composed with
the automorphisms of the curves which sends $z_{11}$ to $z_{21}$
(resp. $z_{11}$ to $z_{31}$), also sends any $z_{1i}$ to $z_{2i}$
(resp. $z_{1i}$ to $z_{3i}$), for $i\geq 2$.
This gives $10$ more conditions,
which are algebraic on the coefficients
of the two quadrics and the entries of the matrices of $\phi,\psi$.
So, we have a total of $34$ conditions, which are algebraic
on the $34$ parameters, i.e. on the projective coordinates
of $G\times PGL(3)\times PGL(3)$. Thus we get at least
a finite number of curves passing through $x_1,\dots,x_6$,
for a general choice of the points.
\hfill\qed
\vskip .5cm
\begin{Rmk} In the previous example, notice that when
the three projections of the points $x_1,\dots, x_k$ differ
by a projectivity, then the number of conditions decreases,
and we find infinitely many normal elliptic curves.
It is easy to see that this implies that a point in the
secant variety $S_6$ of any of these curves, belongs indeed
to infinitely many $6$-secant spaces.
\end{Rmk}
The case of products of projective spaces of dimension $3$ is
particularly interesting, due to its applications to
statistical studies on DNA strings.
If we have many substrings of DNA strings, each formed by three positions,
and we record the occurrence of the four bases in each position, we get a
distribution which can be arranged in a $4\times 4\times 4$ tensor $T$.
The rank $k$ of $T$ suggests the existence of $k$ different types
of substrings, in the probe, such that for each type, the distribution of
bases is independent. So $T$ is the sum of $k$ tensors
$T_1,\dots,T_k$, of rank $1$.
An obvious question concerns the possibility of recovering the
$k$ tensors $T_i$, starting from $T$. When $k\geq 7$, this possibility
is excluded, since $7$ exceeds the maximum given in Proposition \ref{maxk}.
For $k\leq 5$, $k$-identifiability (by Kruskal's criterion)
tells us that, at least theoretically, the reconstruction is possible.
The amazing situation happens for $k=6$. Although one could
expect $6$-identifiability, Theorem \ref{main3} shows that there
are exactly two sets of tensors of rank $1$, whose sum is $T$.
Hence, at least over the complex field, there are exactly two
different sets of distributions, in the $6$ types, that produce
the same distribution $T$.
\medskip
In \cite{AOP} 6.3 one finds the list of known
Segre varieties $X={\mathbb{P}}(A)\times{\mathbb{P}}(B)\times{\mathbb{P}}(C)=
\triplet {a-1}{b-1}{c-1}$ (with $a\le b\le c$) such that
the dimension of $k$-th secant variety is smaller than
the expected value.
Recall that when the dimension of $S_k(X)$ is smaller than the expected value,
i.e. when the variety $X$ is $k$-defective,
then the $k$-identifiability necessarily fails.
A list of known Segre varieties $X$
which are not $k$-identifiable, i.e. such that
the general tensor of rank $k$ in ${\mathbb{C}}^a\otimes{\mathbb{C}}^b\otimes{\mathbb{C}}^c$
has not a unique decomposition, is the following
(for $k<k_{max}$):
{\small
$$\begin{array}{lccl}
&(a,b,c)&k& notes\\
\hline\\
\textrm{defective}&c\ge (a-1)(b-1)+3&(a-1)(b-1)+2\le k
&\textrm{\cite{AOP}}\\
{\textrm{unbalanced}}&&k<\min \left(c,ab\right)&\\
\hline\\
\textrm{defective}&(3,4,4)&5&\textrm{\cite{AOP}}\\
\hline\\
\textrm{defective}&(3,b,b)\quad b\textrm{\ odd}&
\frac{3b-1}{2}&\textrm{\cite{S}}\\[0.1cm]
\hline\\
{\textrm{w. defective}}&3\le a &(a-1)(b-1)+1& \binom d{(a-1)(b-1)+1}
\\ \textrm{unbalanced}& c\ge (a-1)(b-1)+2 &&\textrm{\ decompositions} \\
\textrm{}&&&\textrm{where\ } d={{a+b-2}\choose {a-1}}\\
\textrm{}&&& \textrm{(Theorem \ref{unbal})}\\
\hline\\
{\textrm{w. defective}}&(4,4,4)&6&2\textrm{\ decompositions} \\ \textrm{}&&& \textrm{(Theorem \ref{main3})}\\
\hline\\
{\textrm{w. defective}}&(3,6,6)&8& (**)
\end{array}$$
}
A computer check shows that this list is complete for $c\le 7$.
In the last case marked with (**), the contact variety is a
$4$-fold in ${\mathbb{P}}^{39}$ of degree $108$.
This case needs an ''ad hoc'' analysis which goes beyond the space
of the present note and will be addressed in a forthcoming paper \cite{CMO}.
\vskip.5cm
In the unbalanced case, the identifiability can be proved theoretically.
\begin{Prop}\label{rankab} The general tensor of rank $(a-1)(b-1)$ in
${\mathbb{P}}({\mathbb{C}}^a\otimes{\mathbb{C}}^b\otimes{\mathbb{C}}^c)$
has a unique decomposition as sum of $(a-1)(b-1)$ summands in
$\triplet {a-1}{b-1}{c-1}$ for $c\ge (a-1)(b-1)$.
\end{Prop}
\begin{proof}
Let $\phi\in{\mathbb{C}}^a\otimes{\mathbb{C}}^b\otimes{\mathbb{C}}^c$ be general of rank $(a-1)(b-1)$.
It induces the flattening contraction operator
$$A_{\phi}\colon({\mathbb{C}}^c)^{\vee}\to{\mathbb{C}}^a\otimes{\mathbb{C}}^b$$
which has still rank $(a-1)(b-1)$, by the assumption $c\ge (a-1)(b-1)$.
Indeed, if $\phi=\sum_{i=1}^{(a-1)(b-1)}u_i\otimes v_i\otimes w_i$
with $u_i\in {\mathbb{C}}^a, v_i\in{\mathbb{C}}^b, w_i\in{\mathbb{C}}^c$,
where $w_i$ can be chosen as part of a basis of $C$,
then $\textrm{Im\ }A_{\phi}$ is the span of the representatives
of $v_i\otimes w_i$ for $i=1,\ldots , (a-1)(b-1)$.
It is well known that the projectification of this span,
whose dimension is smaller than the codimension of the Segre variety
$Y={\mathbb{P}}^{a-1}\times{\mathbb{P}}^{b-1}\subset{\mathbb{P}}({\mathbb{C}}^a\otimes{\mathbb{C}}^b)$, meets $Y$
only in these $(a-1)(b-1)$ points
(see for example the Theorem 2.6 in \cite{CC1}).
The claim follows.
\end{proof}
\begin{Prop}
When $c=(a-1)(b-1)$ or $c=(a-1)(b-1)+1$, then the rank of a generic tensor in
${\mathbb{P}}({\mathbb{C}}^a\otimes{\mathbb{C}}^b\otimes{\mathbb{C}}^c)$ is $ab-a-b+2$.
\end{Prop}
\begin{proof}
When $c\ge (a-1)(b-1)+1$, we are in the unbalanced case,
according to the definition 4.2 of \cite{AOP}. In this case the
generic rank is $\min\{c, ab\}$ by (ii)
of the Theorem 4.4 of \cite{AOP}.
Assume $c=(a-1)(b-1)$. Using the same technique, we show that the
secant variety $S_k(\triplet {a-1}{b-1}{c-1})$ has the expected dimension,
for $k\le (a-1)(b-1)$, and fills the ambient space, for $k=(ab-a-b+2)$.
Indeed, with the notations of \cite{AOP},
$T(a-1,b-1,ab-a-b;(a-1)(b-1);0,0,0)$ reduces to $T(a-1,b-1,0;1;0,0,ab-a-b)$,
which is true and subabundant, while $T(a-1,b-1,ab-a-b;ab-a-b+2;0,0,0)$ reduces
(for $b\ge 3$) to $T(a-1,b-1,0;1;0,0,ab-a-b+1)$ and
$T(a-1,b-1,0;2;0,0,ab-a-b-1)$ which are both superabundant and true.
\end{proof}
\begin{Prop}\label{different}
Assume $c\ge (a-1)(b-1)+2$. Then the generic rank in
${\mathbb{P}}({\mathbb{C}}^a\otimes{\mathbb{C}}^b\otimes{\mathbb{C}}^c)$ is at least $(a-1)(b-1)+2$,
and it is equal to $(a-1)(b-1)+2$ in the border case $c=(a-1)(b-1)+2$.
The number of different decomposition of a general
tensor of rank $(a-1)(b-1)+1$ is ${d\choose {(a-1)(b-1)+1}}$
where $d=\deg ({\mathbb{P}}^{a-1}\times{\mathbb{P}}^{b-1})={{a+b-2}\choose {a-1}}$.
This number is always bigger than $1$, with the only exception
$a=b=2$.
\end{Prop}
\begin{proof}
We apply the same argument of the proof of the Proposition \ref{rankab}.
The unique difference is that, now, the dimension of the
projectification of $\textrm{Im\ }A_{\phi}$ equals the codimension
of ${\mathbb{P}}^{a-1}\times{\mathbb{P}}^{b-1}$. Thus we get $d$ points of intersection.
Any choice of $(a-1)(b-1)+1$ among these $d$ points,
yields a decomposition.
\end{proof}
\begin{Rmk}
The case $a=b=3$ of Prop. \ref{different} is connected to the work of tenBerge,
who showed in \cite{tB},
that there are six different decompositions of a general rank $5$
tensor in ${\mathbb{C}}^3\otimes{\mathbb{C}}^3\otimes {\mathbb{C}}^5$, chosen taking $5$ among $6$ possible summands. Our argument,
which we gave for $c\ge 6$, can be extended to the case
$c=5$ and $k=k_{max}=5$, and it gives a geometric explanation of this phenomenon,
indeed the six possible summands correspond to the six intersection points of ${\mathbb{P}}^2\times{\mathbb{P}}^2$
with a general ${\mathbb{P}}^4$.
\end{Rmk}
As a consequence of the two previous results, we get:
\begin{Thm}\label{unbal}
Assume $c\ge (a-1)(b-1)+2$. Then the general tensor of rank $k$ in
${\mathbb{P}}({\mathbb{C}}^a\otimes{\mathbb{C}}^b\otimes{\mathbb{C}}^c)$
has a unique decomposition as sum of $k$ summands in
$\triplet {a-1}{b-1}{c-1}$ if and only if $k\le (a-1)(b-1)$.
\end{Thm}
\section{Products with many factors}\label{many}
At the cost of the growth of the notation,
we can generalize the statement of our main Theorem
\ref{main}, to products of many vector spaces.
In this section, we simply list the corresponding
definitions and results. The proofs are absolutely
straightforward, following the pattern
of the corresponding arguments in the previous sections.
Only the initial step of the induction needs an extra
argument, which is displayed in Lemma \ref{zerostep} below.
\smallskip
For a given set of complex vector spaces $A_1,\dots,A_n$,
with $n\geq 3$ and $\dim A_i\geq 2$, let us give the general:
\begin{Def}\label{p_i} A Segre product
$X={\mathbb{P}}(A_1)\times\dots\times{\mathbb{P}}(A_n)$
is called {\it $(k,p_1,\dots,p_n)$-not weakly defective} if:
for $k$ general points $x_1,\ldots x_k\in X$,
for $p_i$ general points $w_{ij}\in{\mathbb{P}}(A_1)\times\dots\times\hat{\mathbb{P}}(A_i)
\times\dots\times{\mathbb{P}}(A_n)$,
the span of the spaces $T_{x_i}X$, $A_i\otimes w_{ij}$
contains $T_xX$ if and only if $x=x_i$, for some $i=1,\ldots k$.
Otherwise $X$ is called {\it $(k,p_1,\dots,p_n)$-weakly defective}.
\end{Def}
\begin{Rmk}\label{ovvn} (a) With the previous notation, by semicontinuity
it is clear that when $X$ is $(k,p_1,\dots,p_n)$-not weakly defective,
then it is also $(k',p'_1,\dots,p'_n)$-not weakly defective,
whenever $(k',p'_1,\dots,p'_n)\leq (k,p_1,\dots,p_n)$, in the strict ordering.
(b) By semicontinuity, $X$ is $(k,p_1,\dots,p_n)$-not weakly defective
whenever one gets that for {\it particular} sets of points
$\{x_i\}$, $\{w_{ij}\}$, as above,
then the span of
$T_{x_i}X$ and all $A_i\otimes w_{ij}$
contains $T_xX$ if only if $x=x_i$, for some $i=1,\ldots k$.
(c) By Proposition \ref{criterio}, one gets soon that
$(k,0,\dots,0)$-not weakly defective implies $k$-identifiable.
\end{Rmk}
\begin{Lem} \label{riduzn}
Consider $X={\mathbb{P}}(A_1)\times\dots\times{\mathbb{P}}(A_n)$ and assume that,
for a choice of subspaces $A'_i\subset A_i$,
the product ${\mathbb{P}} (A'_1)\times\dots\times{\mathbb{P}}(A'_n)$ is
$(k,p_1,\dots,p_n)$-not weakly defective. Then $X$ is
$(k,p_1,\dots,p_n)$-not weakly defective.
\end{Lem}
The inductive criterion can be rephrased as follows,
always following the lines in \cite{AOP}.
\begin{Prop}\label{induzn} {\bf Inductive Step}
Split the vector space $A_i$ in the sum of two
spaces $A'_i$ and $A''_i$.
Let $X'={\mathbb{P}}(A_1)\times\dots\times {\mathbb{P}}(A'_i)\times\dots\times{\mathbb{P}}(A_n)$,
$X''={\mathbb{P}}(A_1)\times\dots\times {\mathbb{P}}(A''_i)\times\dots\times{\mathbb{P}}(A_n))$,
Assume that the product $X'$ is $(k_1,p'_1,\dots,p_i+k_2,\dots,p'_n)$-not
weakly defective and the product $X''$ is $(k_2,p'_1,\dots,p_i+k_1,\dots,p'_n)$-not
weakly defective.
Then, setting $p_j=p'_j+p''_j$ for $j\neq i$, we get that $X$ is
$(k_1+k_2,p_1,\dots,p_i, \dots, p_n)$-not weakly defective.
\end{Prop}
Now we use again the previous criterion, when the dimension of the
vector spaces are powers of $2$, i.e. when $\dim(A_i)=2^{\alpha_i}$,
for all $i$. We agree to order the spaces, so that
$$ \alpha_1\leq \dots\leq \alpha_n. $$
The following numerical criterion is the exact generalization of Lemmas
\ref{zero} and \ref{uno}.
\begin{Lem}\label{zerostep} Take $X={\mathbb{P}}(A_1)\times\dots\times{\mathbb{P}}(A_n)$, with
$n\geq 3$ and $\dim(A_i)=2^{\alpha_i}\geq 2$.
Pick non-negative integers $u_1,\dots, u_n$
such that, for all $i$:
$$u_i\leq \alpha_1+\dots+\hat \alpha_i+\dots+\alpha_n-(n-1).$$
Then $X$ is $(0,2^{u_1},\dots,2^{u_n})$-not weakly defective and
$(1,2^{u_1}-1,2^{u_2}-1,2^{u_3}-1)$-not weakly defective.
\end{Lem}
\begin{proof} The proof goes by induction. For the inductive step,
one can follow the proof of Lemmas \ref{zero} and \ref{uno},
rephrased for products of many vector spaces.
Thus we only need to check the starting points of the induction, namely
that $Y_n={\mathbb{P}}^1\times\dots\times{\mathbb{P}}^1$ is $(1,0,\dots,0)$-not weakly defective
and $(0,1,\dots,1)$-not weakly defective.
The first fact follows soon, as ${\mathbb{P}}^1\times\dots\times{\mathbb{P}}^1$ is smooth,
so that the general tangent hyperplane is not bitangent.
The second fact follows by induction on $n$. Namely it is true for $n=3$,
as observed in Lemma \ref{uno}. For general $n$,
write $Y_n={\mathbb{P}}(A_1)\times\dots\times {\mathbb{P}}(A_n)$, with
$\dim(A_i)=2$, and split $A_1$ in a direct sum
of two $1$-dimensional spaces $A'$, $A''$.
Using Lemma \ref{riduzn}, one has thus to prove
that $Y_{n-1}={\mathbb{P}}^0\times{\mathbb{P}}^1\times\dots\times{\mathbb{P}}^1$
is $(0,1,0,\dots,0)$-not weakly defective
and $(0,0,1,\dots,1)$-not weakly defective.
The former claim is obvious. The latter
follows by induction.
\end{proof}
We get:
\begin{Prop}\label{pren}
Take $X={\mathbb{P}}(A_1)\times\dots\times{\mathbb{P}}(A_n)$, with
$n\geq 3$ and $\dim(A_i)=2^{\alpha_i}\geq 2$.
Order the $\alpha_i$'s so that $\alpha_1\leq\dots\leq \alpha_n$.
Then $X$ is not $k$-weakly defective, for
$k\leq 2^{\alpha_1+\dots +\alpha_{n-1}-(n-1)}$.
\end{Prop}
It follows that:
\begin{Thm} Take $X={\mathbb{P}}(A_1)\times\dots\times{\mathbb{P}}(A_n)$, with
$n\geq 3$ and $\dim(A_i)=a_i\geq 2$ and, for all $i$,
take $\alpha_i$ maximal, such that $a_i\geq 2^{\alpha_i}$.
Then $X$ is $k$-identifiable, for
$$k\leq 2^{\alpha_1+\dots +\alpha_{n-1}-(n-1)}.$$
\end{Thm}
Comparing our result with the maximal $k$ for which the
identifiability of ${\mathbb{P}}(A_1)\times\dots\times{\mathbb{P}}(A_n)$ makes sense,
which, in the case of a product of many factors, reads as:
$$k_{max}=\lfloor \frac{\prod_{i=1}^{n-1}a_i}
{1+\frac{\sum_{i=1}^{n-1}a_i- (n-1)}{a_n}}\rfloor
$$
we see again that the bound in the Theorem is log-asymptotically sharp.
The inequality of the theorem can be written as
$$k\le 2^{\left(\sum_{i=1}^{n-1}\lfloor\log_2{a_i}-1\rfloor\right)}$$
Since $2^{\alpha_i}\ge\frac{a_i+1}{2}$ we get the general tensor of rank
$k$ is $k$-identifiable if
$$k\le\frac{\prod_{i=1}^{n-1}(a_i+1)}{2^{2n-2}}$$
In \cite{SB} Kruskal bound was extended to the case of $n$ factors.
A sufficient condition for the $k$-identifiability of the general tensor of rank $k$ is
$$2k+n-1\le\sum_{i=1}^n\min(k,a_i)$$
To compare with our condition, in the hypercubic case where $a_i=a$, the bound in \cite{SB}
is $$k\le\frac{n(a-1)+1}{2}$$
while our bound is
$$k\le 2^{(n-1)(\lfloor\log_2{a}-1\rfloor)}$$
For $a\ge 4$ we get also the weaker, but more handy, inequality
$$k\le\left(\frac{a+1}{4}\right)^{n-1}$$
\begin{Ex}
Instead of giving the proofs, which, we repeat, are analogue
to the proofs of the statement of section \ref{pfs}, let us see how
the reduction works in a concrete example.
Take $A_1=\dots =A_5={\mathbb{C}}^{16}$ and consider $X={\mathbb{P}}(A_1)\times\dots\times
{\mathbb{P}}(A_5)$. We want to prove that $X$ is $k$-not weakly defective
for $k=2^{4+4+4+4-4}=4096$.
The reduction step starts as in the following table:
$$
\begin{array}{ccccccccccccc}
A_1 & A_2 & A_3 & A_4 & A_5 & & k & & p_1 & p_2 & p_3 & p_4 & p_5 \\
16 & 16 & 16 & 16 & 16 & & 4096 & & 0 & 0 & 0 & 0 & 0 \\
8 & 16 & 16 & 16 & 16 & & 2048 & & 2048 & 0 & 0 & 0 & 0 \\
8 & 8 & 16 & 16 & 16 & & 1024 & & 1024 & 1024 & 0 & 0 & 0 \\
8 & 8 & 8 & 16 & 16 & & 512 & & 512 & 512 & 512 & 0 & 0 \\
8 & 8 & 8 & 8 & 16 & & 256 & & 256 & 256 & 256 & 256 & 0 \\
8 & 8 & 8 & 8 & 8 & & 128 & & 128 & 128 & 128 & 128 & 128 \\
4 & 8 & 8 & 8 & 8 & & 64 & & 192 & 64 & 64 & 64 & 64 \\
4 & 4 & 8 & 8 & 8 & & 32 & & 96 & 96 & 32 & 32 & 32 \\
4 & 4 & 4 & 8 & 8 & & 16 & & 48 & 48 & 48 & 16 & 16 \\
4 & 4 & 4 & 4 & 8 & & 8 & & 24 & 24 & 24 & 24 & 8 \\
4 & 4 & 4 & 4 & 4 & & 4 & & 12 & 12 & 12 & 12 & 12 \\
2 & 4 & 4 & 4 & 4 & & 2 & & 14 & 6 & 6 & 6 & 6 \\
2 & 2 & 4 & 4 & 4 & & 1 & & 7 & 7 & 3 & 3 & 3
\end{array}
$$
Then use Lemma \ref{zerostep} with $u_1=u_2=3$, $u_3=u_4=u_5=2$.
\end{Ex}
\begin{Rmk} As in the case of triple Segre products,
in principle, there are no obstructions in repeating the
argument, when we substitute powers of $2$
with powers of $3$ (see the proof of Theorem \ref{main2} we gave in case $a=9$),
or any other integer $p>1$.
For some numerical cases, the bound for identifiability that we get
using powers of numbers bigger than two, can be closer to
the maximal value $k_{max}$.
\end{Rmk}
|
\section{Introduction}
The first theoretical calculations of the Cepheids instability strip
done in the 60's assumed that convection was steady with respect to
oscillations. Unfortunately, this ``frozen-in
convection'' approximation led to a cooler red edge than the observed
one as the strong coupling between convection and pulsations occurring in cool
Cepheids was ignored \citep[e.g.][]{Baker65}. Then, following the pioneering
works of
\cite{Unno} and \cite{Gough}, several time-dependent convection (TDC) models
have been developed to investigate the influence of the
convection onto the pulsational stability \citep[e.g.][]{St,K, GW}.
The last up-to-date TDC models actually succeed
in predicting both a red edge close to the observed one and
realistic luminosity curves \citep[e.g.][]{YKB98, Bono, Feuch99,
Koll02}.
However, all of these TDC models suffer from a common weakness due to
the numerous free parameters, usually known as $\alpha$ coefficients, that
describe the turbulent convection
\citep[e.g. the 8 dimensionless parameters in][]{Smolec2010}.
These parameters are either obtained from a fit to the observations or
hardly constrained by theory when taking the asymptotic limit
of stationary convection \citep{BV1, BV2}. A parametric study carried
out by \cite{YKB98} has emphasised the intrinsic degeneracy of TDC
models as similar instability strips have been obtained with different
sets of parameters.
Direct numerical simulations (hereafter DNS) are able to constrain these
TDC models as they fully account for the nonlinearities involved
in the convection-pulsation coupling. These nonlinear simulations
are challenging as they require both
large-amplitude oscillations and convective motions.
In our last 2-D simulations, we have improved the
way acoustic waves are generated by reproducing the self-consistent
excitation operating in variable stars, that is, the $\kappa$-mechanism
\citep[][hereafter GD2011]{paperIII}. The resulting coupling of acoustic modes
with convection is therefore more consistent as the mode amplitude is not
imposed artificially. We have shown in GD2011 that the convective plumes may
either quench the radial oscillations or coexist with the acoustic modes,
depending mainly on the density contrast of the equilibrium model.
The purpose of this letter is to compare the fully nonlinear results
with two main TDC models: (\textit{i}) the first one refers to an
initial formulation of \cite{St} that has been used and improved by
\cite{BS} and \cite{Bono}; (\textit{ii}) the other one has been developed
by \cite{K} and \cite{GW} and is implemented both in the Vienna hydrocode
\citep[e.g.][]{WF,Feuch99} and in the Florida-Budapest one
\citep[e.g][]{YKB98,Koll02}. In these models, a single equation for the
turbulent kinetic energy ${\cal E}_{t}$ is added to the classical mean-field equations
and the main second-order correlations, as for example the convective flux,
are expressed as a function of ${\cal E}_{t}$ only \citep[see e.g.][]{Baker, GW2,
Buchler-ASP, Buchler09}.
We investigate here in more details a particular simulation where the
oscillations strongly modulate the convective flux. The nonlinear results are
first compared at each snapshot with the TDC recipes by computing a
$\chi^2$-statistics of the relevant $\alpha$ coefficients. Secondly, the mean values of
$\alpha$ are used to compare the optimal TDC fluxes with the DNS ones.
The formulation of Stellingwerf is found to be closer to the nonlinear
result than Kuhfu\ss's one: (\textit{i}) the temporal variation of the
$\alpha$ coefficient is weaker; (\textit{ii}) the mean convective
flux is closer to the DNS one, especially in the description of the
overshooting area.
\section{The hydrodynamic model}
We consider a local 2-D box of size $L_x\times L_z$, filled by a
perfect monatomic gas, and centered on both sides of an ionisation
region, of which the associated opacity bump is shaped by a hollow in the
temperature-dependent radiative conductivity profile $K(T)$. Such a
configuration can lead to unstable acoustic modes due to the driving
term $\na \cdot K(T)\na T$ in the energy equation, i.e. this is
the $\kappa$-mechanism \citep{paperI, paperII}. Furthermore, this hollow
in $K(T)$ is deep enough such as the equilibrium temperature gradient is
locally superadiabatic and convective motions develop here according to
Schwarzschild's criterion.
The governing hydrodynamic equations are written in nondimensional
form by choosing $L_z$ as the length scale and $\sqrt{c_pT_{\hbox{\scriptsize top}}}$
as the velocity one, hence the time scale $L_z/\sqrt{c_pT_{\hbox{\scriptsize top}}}$ (with $c_p$
the specific heat and $T_{\hbox{\scriptsize top}}$ the surface temperature).
The resulting nonlinear set of equations is advanced in time
with the high-order finite-difference pencil code\footnote{See
\url{http://www.nordita.org/software/pencil-code/} and
\cite{Pencil-Code}.}, which is fully explicit except for the radiative
diffusion term that is solved implicitly thanks to a parallel alternate
direction implicit (ADI) solver. With the chosen units, the simulation box
spans about $10\%$ of the star radius on both sides of the ionisation region while
the timestep is about 1 minute, such that a simulation typically spans over
4500 days (see GD2011).
\section{Results}
The time evolution of the convective flux obtained in fully nonlinear
2-D simulations is compared with the following TDC expressions developed
by \cite{St} and \cite{K}:
\begin{equation}
\left\lbrace
\begin{aligned}
{\cal F}_{\text{St}}(z,t) & = \alpha_{\text{St}} \dfrac{A}{B} {\cal E}_{t} \sign(\nabla
-\nabla_{\text{ad}})\sqrt{|\nabla -\nabla_{\text{ad}} |}, \\
{\cal F}_{\text{Ku}}(z,t) & = \alpha_{\text{Ku}} A \sqrt{{\cal E}_{t}} \left( \nabla -\nabla_{\text{ad}} \right),
\end{aligned}
\right.
\label{eq:MLT}
\end{equation}
where $\nabla=d\ln T/d\ln p$, $\nabla_{\text{ad}}=1-c_v/c_p$ and
\begin{equation}
{\cal E}_{t}(z,t) =\left\langle\dfrac{{u'_z}^2}{2}\right \rangle,\
A = c_p\left\langle \rho \right\rangle \left\langle T \right\rangle
\text{ and } B= \sqrt{c_p\left\langle T \right\rangle \nabla_{\text{ad}}},
\label{eq:eturb}
\end{equation}
with $p$ and $\rho$ the pressure and density, respectively, and
the brackets denote an horizontal average. Two free dimensionless
parameters $\alpha_{\text{St}}$ and $\alpha_{\text{Ku}}$ are introduced that the simulations allow
to constrain. We first focus on a 2-D DNS that is similar
to the G8 simulation in GD2011, that is, a simulation in which the total
kinetic energy is almost entirely contained in acoustic modes excited by
$\kappa$-mechanism (80\%), the rest being in convective plumes (20\%).
The convective motions do not affect much the growth and the nonlinear
saturation of the unstable radial acoustic mode. We can therefore expect
that the heat transport is strongly modulated by wave motions, what is an
ideal frame for testing the accuracy of time-dependent convection models.
\subsection{Temporal modulation of the convective flux in the DNS}
\begin{figure}
\centering
\includegraphics[width=9cm]{fig/fluxes}
\caption{\textbf{a)} Mean vertical profiles of radiative ${\cal F}_{\text{rad}}$
(solid blue line), turbulent enthalpy ${\cal F}_{\text{conv}}$ (dashed green line), kinetic
${\cal F}_{\text{kin}}$ (dotted black line), modes ${\cal F}_{\text{mod}}$ (dot-dashed magenta line) and total
${\cal F}_{\text{tot}}$ (dotted red line) fluxes, normalised to the bottom
flux $F_{\hbox{\scriptsize bot}}$. \textbf{b)} Temporal power spectrum for the convective flux
only.}
\label{fig:fluxes}
\end{figure}
The imposed bottom flux $F_{\hbox{\scriptsize bot}}$ is mainly transported through the
computational domain by the radiative ${\cal F}_{\text{rad}}$, enthalpy and kinetic ${\cal F}_{\text{kin}}$
fluxes. Following GD2011, the enthalpy flux is divided into the classical
convective flux ${\cal F}_{\text{conv}}$ and the contribution coming from the (unstable)
fundamental mode (hereafter ${\cal F}_{\text{mod}}$) with
\begin{equation}
\left\lbrace
\begin{array}{lll}
{\cal F}_{\text{conv}}(z,t) & = c_p \left\langle \rho u_z T' \right\rangle,\quad
& {\cal F}_{\text{mod}}(z,t) = c_p \left\langle \rho u_z\right\rangle\theta, \\ \\
{\cal F}_{\text{rad}}(z,t) & = -\left\langle K(T) \nabla T \right\rangle,\quad
& {\cal F}_{\text{kin}}(z,t) = \dfrac{1}{2}\left\langle \rho u_z u^2 \right\rangle,
\end{array}\right.
\label{eq:fluxes}
\end{equation}
where the primed quantities denote the fluctuations about the horizontal
average and $\theta$ is the temperature eigenfunction of the fundamental
mode. The resulting time-averaged and normalised fluxes are given in
Fig.~\ref{fig:fluxes}a.
The bulk of the total flux is transported by the
radiative flux, except in the convective zone where ${\cal F}_{\text{conv}}/F_{\hbox{\scriptsize bot}}\simeq
20\%$, while the kinetic flux is negligible (${\cal F}_{\text{kin}}/F_{\hbox{\scriptsize bot}}\leq
1\%$). Concerning ${\cal F}_{\text{mod}}$, one notes that it is hardly as large as ${\cal F}_{\text{conv}}$
in the convective zone. This quantity is a
good signature of the amplitude of the acoustic modes as the higher ${\cal F}_{\text{mod}}$,
the larger the radial oscillations (GD2011), therefore confirming the efficiency
of the $\kappa$-mechanism in this simulation.
The signature of the temporal modulation of the convective flux
is extracted from its Fourier spectrum in time, that is, we first
compute $\widehat{\cal F}_{\text{conv}}(z,\omega)$, with $\omega$
the frequency, and second integrate over the vertical direction $z$
to get the power spectrum $\widehat{\cal F}_{\text{conv}}(\omega)$
(Fig.~\ref{fig:fluxes}b). Several discrete peaks appear about given
frequencies, of which the physical nature is emphasised after superimposing
the linear acoustic eigenfrequencies (the vertical dashed blue lines). The
matching with the fundamental mode frequency $\omega_{00}=3.85$ is
perfect while the weak-amplitude secondary peaks rather correspond to
harmonics of $\omega_{00}$ (i.e. $2\omega_{00},3\omega_{00},\dots$,
shaped by downward-directed vertical gray arrows in
Fig.~\ref{fig:fluxes}b) than to the acoustic overtones $\omega_{01},\
\omega_{02},\dots$. It means that the amplitude of the fundamental
acoustic mode is large enough to generate several harmonics through a
nonlinear cascade and this is again an indication of both the
robustness of the $\kappa$-driving and the relevance of this kind of
convection-pulsation simulation to check the TDC recipes.
\begin{figure*}
\centering
\includegraphics[width=18cm]{fig/contours}
\caption{Evolution of the convective flux in a $(t,\ z)$ plane: (\textbf{a}) in the DNS; (\textbf{b}) with Stellingwerf's formalism ${\cal F}_{\text{St}}$;
(\textbf{c}) with Kuhfu\ss's~formalism ${\cal F}_{\text{Ku}}$ (for
$\alpha_{\text{St}}=\alpha_{\text{Ku}}=1$). The
vertical extent is centered on both sides of the convection zone to emphasise its oscillations.
These snapshots are computed after 1800 periods of oscillations, i.e. well
after the
nonlinear saturation is achieved.}
\label{fig:contours}
\end{figure*}
\subsection{DNS vs TDC models: convective patterns}
We first compare the time evolution of the horizontally averaged
convective flux obtained in the DNS (Eq.~\ref{eq:fluxes}) with its
TDC counterparts ${\cal F}_{\text{St}}$ and ${\cal F}_{\text{Ku}}$ (Eq.~\ref{eq:MLT}). As we are interested here in the
qualitative agreement between the convective patterns in the
plane $(t,\ z)$, we simply assume $\alpha_{\text{St}}=\alpha_{\text{Ku}} =1$.
The three resulting patterns are displayed in Fig.~\ref{fig:contours} for
a time interval spanning about 4 periods of the fundamental acoustic
mode. The black areas denote positive values for the convective flux
and therefore delimit the convective zone. An oscillatory behaviour is
clearly visible in each panel, with a period that looks similar to
the one of the fundamental acoustic mode, that is, almost 4 oscillation
cycles are depicted. This is consistent with the large peak shown around
$\omega_{00}$ in the power spectrum in Fig.~\ref{fig:fluxes}b.
The two TDC patterns also display a good agreement with the
nonlinear simulation regarding the overshooting phenomenon. We
indeed recover the same dark red structures below the convective
zone that correspond to the downdrafts entering in the radiative zone
(this penetration being associated with a negative convective flux).
Nevertheless, we note that the overshooting seems to be more vigorous
in Kuhfu\ss's model than in Stellingwerf's one as these dark
red structures fill in Fig.~\ref{fig:contours}c a larger surface in the bottom
radiative zone than in Fig.~\ref{fig:contours}b.
\subsection{DNS vs TDC models: statistics of coefficients $\alpha$}
\label{stat}
\begin{figure}
\centering
\includegraphics[width=9cm]{fig/alphas}
\caption{Time evolution around the mean value of coefficients
$\alpha_{\text{St}}$ (\textit{upper panel}) and $\alpha_{\text{Ku}}$ (\textit{bottom panel}).
Horizontal gray spans mark the limits of the associated relative
standard deviation.}
\label{fig:alphas}
\end{figure}
A one-to-one comparison between the convective flux in the DNS and
its TDC predictions requires to find the optimal values of $\alpha_{\text{St}}$
and $\alpha_{\text{Ku}}$. This is done by performing several $\chi^2$-tests
at different snapshots in the simulation to track the variations of
coefficients $\alpha$. The resulting fluctuations of $\alpha_{\text{St}}$ and $\alpha_{\text{Ku}}$
around their mean values are shown in Fig.~\ref{fig:alphas}.
The dispersion of the Stellingwerf coefficient $\alpha_{\text{St}}$ (upper panel)
is weaker than the Kuhfu\ss~one (lower panel) as its values are almost within
a $5\%$ range around the mean
$\overline{\alpha_{\text{St}}} = 0.462$.
On the contrary, several outliers with values greater than
$10\%$ are found in the evolution of $\alpha_{\text{Ku}}$ which then appears more
chaotic around the mean value $\overline{\alpha_{\text{Ku}}}=0.076$. As a consequence,
the relative standard deviation (depicted in gray in Fig.~\ref{fig:alphas}) is
weaker in Stellingwerf's case than in Kuhfu\ss's one.
\subsection{DNS vs TDC models: convective fluxes with optimal $\alpha$'s}
\begin{figure}
\centering
\includegraphics[width=9cm]{fig/fluxMLT}
\caption{Mean convective flux in the DNS (solid black line), compared
with the best TDC predictions based on the optimal $\alpha$ values that
came out of the $\chi^2$-test in \S\ref{stat} (Stellingwerf's model in
dashed blue line and Kuhfu\ss's one in dotted green line).}
\label{fig:fluxMLT}
\end{figure}
We recall that TDC models assume that the coefficients $\alpha$ entering
in Eq.~\ref{eq:MLT} are constant. By adjusting the TDC recipes with
the instantaneous convective flux throughout the nonlinear simulation,
the optimal value of these coefficients has been deduced. The final
test, given in Fig.~\ref{fig:fluxMLT}, then consists in the comparison
between the mean nonlinear convective flux taken over the entire
simulation and its best TDC approximations built from these optimal
$\alpha$ values.
This figure emphasises that Stellingwerf's formulation
gives a better agreement than Kuhfu\ss's one. Indeed, the
Kuhfu\ss~model overestimates the overshooting as the (negative) convective
flux remains non-negligible until the bottom of the radiative zone. On
the contrary, the Stellingwerf profile better accounts for the local
penetration of convective plumes and shows the same exponential-like
decay of the negative convective flux when sinking in the
radiative zone \citep[e.g.][]{Dintrans2009}.
However, the two models are rather similar in the bulk of the convective
zone where convection is fully developed. One also notes that they both
predict a negative flux at the top of the convective zone, that is,
an upper overshooting of convection motions near the surface that is
not observed in the nonlinear simulation.
\section{Conclusion}
The main weakness of theories of time-dependent convection (TDC) lies in
the large number of free parameters. This is particularly awkward
when convection is strongly coupled with pulsations as, for example,
near the red border of the Cepheid instability strip where similar
results are obtained with different sets of parameters \citep{YKB98}.
Additional constraints must be found to reduce the intrinsic degeneracy
of models. But the modelling of the interplay between the turbulent
convection and oscillations is really a difficult task due to the strong
nonlinearities involved in this coupling.
One solution to tackle this problem and to bring new constraints
consists in performing fully nonlinear simulations, and this is
the path we have chosen in this work following our first study
in GD2011. Two widely used TDC theories, namely the \cite{St} and \cite{K}
ones, are compared
with results coming from 2-D nonlinear simulations of compressible
convection in which strong acoustic oscillations are self-sustained
by the $\kappa$-mechanism. The heat transport is then modulated by the
fundamental acoustic mode such that this kind of simulation is
relevant to investigate the convection-pulsation coupling
(Figs.~\ref{fig:fluxes}-\ref{fig:contours}).
Focusing on the two TDC formulae for the convective flux, we compute the
evolution of free coefficients ``$\alpha$'' from a $\chi^2$-test applied
to the fully nonlinear results (Fig.~\ref{fig:alphas}). A large temporal
variability is found in both cases that weakens the robustness of
the TDC assumption $\alpha=\hbox{const}$. Moreover, the mean values
$\overline{\alpha}$ are not universal. By
applying the same method to other simulations performed in GD2011
(Table~\ref{tab:alphas}), we indeed do not recover the same
$\overline{\alpha}$ with a relative standard deviation of about
$12\%$ (Stellingwerf) and $18\%$ (Kuhfu\ss).
\begin{table}
\centering
\caption{Values of the optimal Stellingwerf and Kuhfu\ss~$\alpha$
coefficients pulled out from the nonlinear simulations in GD2011.}
\begin{tabular}{ccc}
\toprule
Simulation & $\overline{\alpha_{\text{St}}}$ & $\overline{\alpha_{\text{Ku}}}$ \\
\midrule
{\textbf{G8}} & 0.46 & 0.076\\
G8H9 & 0.47 & 0.099 \\
G8H8 & 0.33 & 0.113 \\
G7 & 0.38 & 0.098 \\
G6 & 0.38 & 0.102 \\
G6F7 & 0.40 & 0.082 \\
G6F5 & 0.38 & 0.067 \\
\bottomrule
\end{tabular}
\label{tab:alphas}
\end{table}
Within these limits, Stellingwerf's formulation is found to give a better
agreement with the nonlinear simulations than Kuhfu\ss's one: ({\it i})
the final mean convective flux ${\cal F}_{\text{St}}$ is closer to its DNS counterpart,
with a much better estimation of the bottom overshooting; ({\it ii})
the temporal dispersion of the $\alpha_{\text{St}}$ coefficient is weaker, then its
enhanced stability (Fig.~\ref{fig:fluxMLT}). This result means that the
time-dependent convective flux better scales with a law ${\cal F}_{\text{conv}} \propto {\cal E}_{t}
\sqrt{\nabla -\nabla_{\text{ad}} }$ than ${\cal F}_{\text{conv}} \propto \sqrt{{\cal E}_{t}}(\nabla -\nabla_{\text{ad}})$,
and the Stellingwerf formulation may probably be preferred in the 1-D
hydrocode used in, e.g., the topical Cepheids models. However, this study
emphasises that both formalisms lead to an artificial overshooting at
the top of the convection zone. We also note that this TDC test
involves the \emph{exact} value of the turbulent kinetic
energy ${\cal E}_{t}$ provided by the nonlinear simulation, and not by the dedicated
1-D TDC equation which is inherently less accurate. As a consequence,
the obtained profiles are certainly the best we can expect from
these TDC recipes.
In this work, the temporal modulation of the convective flux is ensured
by the acoustic modes excited by $\kappa$-mechanism. An interesting
prospect could be the case of a modulation based on the internal
gravity waves excited by convection itself. Indeed, convection can
excite gravity waves in variable stars, either by the means of the
penetration of convective elements into stably stratified regions as
in solar-type stars \citep[e.g.][]{Din05}, or through the so-called
``convective blocking'' mechanism as in white dwarfs or Gamma Doradus
stars \citep[e.g.][]{Pes87}. In both cases, the convective flux is ipso
facto modulated by gravity waves and it may be interesting in that
respect to also check the accuracy of time-dependent convection models.
\begin{acknowledgements}
This work was granted access to the HPC resources of CALMIP under the
allocation 2010-P1021 (\url{http://www.calmip.cict.fr}).
\end{acknowledgements}
\nocite{*}
\bibliographystyle{aa}
|
\subsection*{Introduction}
The black hole entropy calculation in the framework of loop quantum gravity \cite{lqg} is based on the effective description of the
quantum gravitational degrees of freedom at the black hole horizon obtained from a suitable quantization of the classical phase space describing isolated horizons (see \cite{ih} and references therein). In these models the degrees of freedom at the horizon are described by Chern-Simons theories with $SU(2)$ (or $U(1)$) structure groups \cite{ab,jon, su21,su22,su23}.
The simplest models are those where spherical symmetry is imposed already at the classical level. In this case it is natural (although not necessary) to consider
$SU(2)$ (or $U(1)$) Chern-Simons theory with a level that scales with the macroscopic classical area $k\propto a_{\scriptscriptstyle H}$. This makes
the state-counting (necessary for the computation of the entropy) a combinatorial problem which can be entirely formulated in terms of the representation theory of the classical group $SU(2)$ (or $U(1)$): for practical purposes one can take $k=\infty$ from the starting point \cite{counting1, counting2, counting3, hanno}.
However, the perspective considered above can be completely changed if one studies the models in the recently introduced $SU(2)$ Chern-Simons formulation \cite{su21,su22,su23}. The necessity of an $SU(2)$ gauge invariant formulation comes from the requirement that the isolated horizon quantum constraints be consistently imposed in the quantum theory (in \cite{su21} it is shown how the $U(1)$ treatment leads to an artificially larger entropy due to the fact that some of the second class constraints arising from the $SU(2)$-to-$U(1)$ gauge fixing can only be imposed weakly). However, the $SU(2)$ formulation is not unique as there is a one parameter family of classically equivalent $SU(2)$ connections parametrizations of the horizon degrees of freedom. More precisely, in the passage from Palatini-like variables to connection variables that is necessary for the description of the horizon degrees of freedom in terms of Chern-Simons theory (central for the quantization), an ambiguity parameter arises \cite{su22,su23}. This is completely analogous to the situation in the bulk where the Immirzi parameter reflects an ambiguity in the choice of $SU(2)$ variables in the passage from Palatini variables to Ashtekar-Barbero connections (central for the quantization in the loop quantum gravity approach). In the case of the parametrization of the isolated horizon degrees of freedom, this ambiguity can be encoded in the value of the Chern-Simons level $k$, which, in addition to the Immirzi parameter, becomes an independent free parameter of the classical formulation of the isolated horizon-bulk system.
Therefore, it is no longer natural (nor necessary) to take $k\propto a_{\scriptscriptstyle H}$. On the contrary, it seems more natural to exploit the existence of this ambiguity by letting $k$ be arbitrary \footnote{It is a good thing that the effective treatment contains a free parameter arising at the boundary from exactly the analogous reason as the Immirzi parameter in the bulk parametrization of the phase space. This keeps open the possibility that dynamical considerations could lead to cancelation of both ambiguities producing Immirzi parameter independent predictions.}. In this way the $SU(2)$ classical representation theory involved in previous calculations should be replaced by the representation theory of the quantum group $U_q(su(2))$
with $q$ a non-trivial root of unity. Thus quantum group corrections become central for the state-counting problem. In this paper
we study the finite $k$ counting problem by means of simple asymptotic methods.
The powerful methods that have been developed for the resolution of the counting problem in the $k=\infty$ \cite{counting3,hanno} are perhaps generalizable to the finite $k$ case. Here we follow a less rigorous and more physical approach. The formulation is partly inspired from a combination of ideas stemming from different
calculations in the literature \cite{counting2, kaulma, livi}.
\medskip
In the first section, we review some basic facts concerning the quantization of $SU(2)$ Chern-Simons theory
whose physical states are built from the representation theory of the quantum group $U_q(su(2))$ when $q$ is a root
of unity. For that reason, we recall some properties of the representation theory of $U_q(su(2))$,
which allows us to compute the dimension of the Chern-Simons theory Hilbert space ${\cal H}_{CS}$ when the space is a punctured
two-sphere. In the second Section, we give a new integral formulation of the dimension of ${\cal H}_{CS}$ which appear to be much more
convenient to compute black hole entropy. In the last Section, we compute the leading term of the $SU(2)$-black hole entropy and its logarithmic
corrections first for the spherically symmetric black hole and then for the distorted black hole.
We adapt the techniques used in \cite{counting3} and firstly introduced in \cite{counting1} to compute the entropy
of a black hole. We are not going into the mathematical details of these techniques which has been very well exposed
in \cite{counting3} and which are in fact very well-known in the domain of probabilities and used
to understand some properties of random walks. We recover that in the spherically symmetric and distorted black holes, the leading
term of the entropy is proportional to the area and the first corrections are still logarithmic: $S(a) \sim \alpha a + \beta \log a$.
In the spherically symmetric case, $\alpha$ depends on the level $k$ and reaches the value obtained in previous calculations when $k$
goes to infinity; concerning $\beta$ it is independent of $k$ and is given by the value $-3/2$ as expected.
In the distorted case, $\alpha$ grows logarithmically with $k$ and $\beta$ is fixed to the value $-3$.
We finish with a discussion.
\subsection*{1. The Chern-Simons Hilbert Space}
The Chern-Simons theory associated to the group $SU(2)$ is a gauge theory on a three dimensional manifold $M$
governed by the action
\begin{eqnarray}
S_k[A] \; = \; \frac{k}{4\pi} \int_{M} \langle A \wedge dA + \frac{2}{3} A \wedge A \wedge A \rangle
\end{eqnarray}
where $k$ is called the level of the action, $A$ is the local $SU(2)$-connection field and $\langle \cdot,\cdot \rangle$
is a notation for the $\mathfrak{su}(2)$ Killing form.
The Chern-Simons theory became really important when first it was shown to be closely related to three dimensional gravity \cite{AT,Witten1}
and above all when Witten showed \cite{Witten2} its amazing relation to manifold and knots invariants. Indeed, the Chern-Simons path integral
is a manifold invariant whereas the mean values of quantum observables naturally lead to Jones polynomials. For all these reasons, Chern-Simons
theory has been the center of a lot of interests and its quantization is now very well-known when the gauge group is compact, and in
particular when the gauge group is $SU(2)$.
\medskip
The covariant (path integral) and canonical quantizations offer the two main strategies to quantize the Chern-Simons
theory. These approaches are complementary: the covariant quantization leads easily to the fact that the level $k$ must be
an integer when the gauge group is compact \cite{Witten2}; the canonical quantization leads to a precise description of the Hilbert space
when the gauge group is compact but not only (see \cite{CK} for an introduction of the combinatorial quantization for example).
Both quantizations are necessary to understand how the mean value of
Wilson loops observables are related to knots polynomials like the (colored) Jones polynomial or its generalizations.
\medskip
Here we are exclusively interested in the description of the Hilbert space of Chern-Simons theory when the space
is a two-sphere punctured with a number $p$ of particles. At the classical level, each puncture, labelled by $\ell \in [1,p]$,
comes with an unitary irreducible representation $j_\ell$ of the gauge group $SU(2)$. At the quantum level, one shows that
the classical group gauge symmetry is replaced by a quantum group symmetry and the Hilbert space is constructed from
the representation theory of the quantum group $U_q(su(2))$ where $q=\exp(i\pi/(k+2))$ is necessarily a root of unity.
An immediate consequence is that the $SU(2)$ representations labeling the classical punctures become $U_q(su(2))$
representations which concretely implies a cut-off on the punctures' representations which cannot be higher than $k/2$.
Then, the associated Hilbert space is the vector space
\nopagebreak[3]\begin{equation}\label{Hilbert} H_k(j_1,\cdots,j_p) \; = \; \text{Inv}(\otimes_\ell j_\ell)\end{equation}
of invariant tensors in the tensor product $\otimes_\ell j_\ell$
of $U_q(su(2))$ representations endowed with a Hilbert structure defined, as in the classical case, from the Haar measure on the quantum group.
However, we will not be interested in the Hilbert structure of $H_k(j_1,\cdots,j_\ell)$ in the rest of the paper but rather in its
vector space structure and more precisely in its dimension. Indeed, the computation of the $SU(2)$ black hole entropy in Loop
Quantum Gravity needs to be done precisely the calculation of the dimension of the previous vector space.
The calculation of the Hilbert space dimension makes use of the Verlinde coefficients.
In order to introduce these coefficients, we start by recalling some basic facts concerning the representation theory
of $U_q(su(2))$.
\subsubsection*{1.1. Basics of the representation theory of $U_q(su(2))$}
This Section is devoted to recall some basic results on the quantum group $U_q(su(2))$ we will need in the sequel.
We are not going to give a precise definition of this quantum group and a complete description of its properties.
Furthermore we will be interested only on some aspects concerning its representations theory and its recoupling theory.
The (standard) unitary irreducible representations of $U_q(su(2))$ are labelled by integers $j \leq k/2$.
The dimension $d_j$ of the $j$-representation is the same as in the
classical theory and then we have $d_j=2j+1$. Given an element $\xi \in U_q(su(2))$, its representation $\pi_j(\xi)$ is an endomorphism of the vector space
$V_j$.
Many formulae coming from the representation theory of $U_q(su(2))$ are expressed in terms of q-numbers $[x]$ defined for any complex
number $x$ by the relation:
$$
[x] \; = \; \frac{q^x-q^{-x}}{q-q^{-1}} \; = \; \frac{\sin(\frac{\pi}{k+2}x)}{\sin(\frac{\pi}{k+2})}\;.
$$
Invariant $U_q(su(2))$-tensors are defined, by analogy with the classical situation, as tensors which are
invariant under the adjoint action. Note however that the adjoint
action is deformed compared to the classical case and makes use of the antipode instead of the inverse.
Among the invariant tensors, the 3-valent ones
$$\iota(j_1,j_2;j_3): V_{j_1} \otimes V_{j_2} \; \longrightarrow \; V_{j_3}$$
are particularly interesting because all invariant tensors decompose into 3-valent intertwiners.
Three-valent intertwiners are represented as usual by a vertex between three lines colored by the representations
$j_\ell$ as illustrated in the figure (\ref{3valent}).
\begin{figure}[ht]
\psfrag{j1}{$j_1$}
\psfrag{j2}{$j_2$}
\psfrag{j3}{$j_3$}
\centering
\includegraphics[scale=0.3]{3valent.eps}
\caption{Pictorial representation of the 3-valent intertwiner $\iota(j_1,j_2;j_3)$ and its adjoint operator $\iota(j_3;j_1,j_2)$.}
\label{3valent}
\end{figure}
Contrary to what happens in the classical case where one can make a certain choice of normalization
such that the matrix elements of $\iota(j_1,j_2;j_3)$ are invariant under the permutations of $(j_1,j_2)$,
the order between the representations in $\iota(j_1,j_2,j_3)$ does
matter because of the presence of a non-trivial braiding in the quantum case.
In the sequel, we fix the normalization of $\iota(j_1,j_2;j_3)$ such that it satisfies the following fusion rule:
\nopagebreak[3]\begin{equation}
\sum_{j_3} [d_{j_3}] \, i(j_1,j_2;j_3) \cdot i(j_3;j_1,j_2) \; = \; \mathbb I_{j_1\otimes j_2}
\end{equation}
where $\mathbb I_{j_1 \otimes j_2}$ is the identity map in the tensor product of the two representations $j_1$ and $j_2$
and $\iota(j_3;j_1,j_2):V_{j_3} \rightarrow V_{j_1} \otimes V_{j_2}$ is the adjoint of $\iota(j_1,j_2;j_3)$. This relation implies
that the so-called $\theta$-graph is normalized to one
\nopagebreak[3]\begin{equation}
\text{Tr}_{j_1\otimes j_2} (\iota(j_1,j_2;j_3) \cdot \iota(j_3;j_1,j_2)) \; = \;
\text{Tr}_{j_3} (\iota(j_3;j_1,j_2) \cdot \iota(j_1,j_2;j_3)) \; = \; Y(j_1,j_2,j_3) \;,
\end{equation}
which is equivalent to
\nopagebreak[3]\begin{equation}
\iota(j_3;j_1,j_2) \cdot \iota(j_1,j_2;j_3) \; = \; \frac{1}{[d_{j_3}]} \, Y(j_1,j_2,j_3) \,\mathbb I_{j_3}\,.
\end{equation}
We made used of the notation $Y(j_1,j_2,j_3) \in \{0,1\}$ which is one only when $(j_1,j_2,j_3)$ satisfy the triangular inequalities,
otherwise it vanishes. These identities are graphically represented in the figure (\ref{norm}).
\begin{figure}[ht]
\psfrag{j1}{$j_1$}
\psfrag{j2}{$j_2$}
\psfrag{j3}{$j_3$}
\psfrag{=}{$=$}
\psfrag{S}{$\sum_{j_3}[d_{j_3}]$}
\psfrag{C}{$Y(j_1,j_2,j_3)$}
\psfrag{A}{${Y(j_1,j_2,j_3)}{[d_{j_3}]}^{-1}$}
\centering
\includegraphics[scale=0.5]{norm.eps}
\caption{Illustration of the normalization properties: the three relations are in fact equivalent
to the condition that the $\theta$-graph is normalized to one.}
\label{norm}
\end{figure}
\medskip
Of particular interest for the quantization of Chern-Simons theory is the fact that $U_q(su(2))$ is quasi-triangular and
therefore admits an universal R-matrix which is at the origin of the braiding properties associated to the quantum groups. Without
entering too much into the details, let us recall that
$R \in U_q(su(2))\otimes U_q(su(2))$ satisfies in particular the so-called quantum Yang-Baxter equation and other defining
properties that one can find in \cite{CP} for example.
The evaluation of the R-matrix in the tensor product of representations $j_1 \otimes j_2$
is denoted $R_{j_1j_2}=(\pi_{j_1}\otimes \pi_{j_2})(R)$ and defines a braiding operator from $V_{j_1}\otimes V_{j_2}$
to the opposite tensor product $V_{j_2} \otimes V_{j_1}$.
It is useful to represent the R-matrix as in the picture (\ref{braiding}):
if $R$ is represented by an under-crossing (the up-line undercrosses the down-line)
then its inverse $R^{-1}$ is represented by an over-crossing (the up-line overcrosses the down-line).
It is clear from this representation that the product of $R$ by its inverse is the identity because the braiding has been unknoted.
\begin{figure}[h]
\psfrag{j1}{$j_1$}
\psfrag{j2}{$j_2$}
\psfrag{R}{$R$}
\psfrag{R-1}{$R^{-1}$}
\centering
\includegraphics[scale=0.5]{braiding.eps}
\caption{Pictorial representation of the R-matrix and its inverse $R^{-1}$. Both R-matrices are evaluated in $j_1\otimes j_2$.}
\label{braiding}
\end{figure}
\subsubsection*{1.2 From Verlinde coefficients to the Hilbert space dimension}
Now, we have all the ingredients to construct the Verlinde coefficients. These coefficients appeared first \cite{Verlinde}
in the context of conformal field theory
and then it has been realized that they have a very simple algebraic interpretation in the context of quantum groups.
Here we will give only their algebraic definition
and some of their properties which are important in the calculation of the dimension of the Hilbert space $H_k(j_1,\cdots,j_\ell)$.
\medskip
Given two unitary irreducible representations $j_1$ and $j_2$, one defines the Verlinde coefficient $S_{j_1j_2}={S}_{j_2j_1}$ as
the real number determined by the trace on $V_{j_1}\otimes V_{j_2}$ of the operator $R^2$ up to a normalization factor:
\nopagebreak[3]\begin{equation}
{S}_{j_1j_2} \; = \; {\cal Z} \, {\text{Tr} \, (R_{j_2 j_1} R_{j_1j_2})}
\end{equation}
where
\nopagebreak[3]\begin{equation}
{\cal Z}=\sqrt{\frac{2}{k+2}} \,{\sin(\frac{\pi}{k+2})}
\end{equation}
is in fact the partition function of the $SU(2)$ Chern-Simons theory on the 3-sphere $S^3$.
It will be useful in the sequel to use the ``un-normalized" Verlinde coefficient $\widetilde{S}_{j_1j_2}=\text{Tr} \, (R_{j_2 j_1} R_{j_1j_2})$
and the choice of the normalization factor will appear clear soon. Note that $\widetilde{S}_{j_1j_2}$ is the evaluation on the Hopf-link
embedded into the 3-sphere. The Hopf-link is represented in the figure (\ref{Hopflink}).
\begin{figure}[ht]
\psfrag{equiv}{$\equiv$}
\centering
\includegraphics[scale=0.5]{hopflink.eps}
\caption{Representation of the Hopf-link. The evaluation of the associated quantum spin-network colored with the representations
$j_1$ and $j_2$ gives the un-normalized Verlinde coefficient $\tilde{S}_{j_1j_2}$.}
\label{Hopflink}
\end{figure}
The explicit expression of the R-matrix implies that
\nopagebreak[3]\begin{equation}\label{explicitVerlinde}
\widetilde{S}_{j_1j_2} \; = \; [d_{j_1}d_{j_2}] \; = \; \frac{\sin(\frac{\pi d_{j_1} d_{j_2}}{k+2})}{\sin(\frac{\pi}{k+2})}\;.
\end{equation}
These coefficients satisfy many interesting properties which are important to compute the dimension of the physical Hilbert space
$H_k(j_1,\cdots,j_\ell)$ presented above. The properties we will need are given below:
\nopagebreak[3]\begin{eqnarray}
&&\text{The normalization relation:} \;\;\; \sum_{j_3} \widetilde{S}_{j_1j_3} \widetilde{S}_{j_3j_2} \; =
\; \frac{\delta_{j_1j_2}}{{\cal Z}^2}; \label{normalization}\\
&&\text{The fusion relation:} \;\;\; \widetilde{S}_{j_1j_3} \widetilde{S}_{j_2j_3} \; = \; [d_{j_3}] \sum_{\ell} Y(j_1,j_2,\ell) \; \widetilde{S}_{j_3\ell}; \\
&&\text{The recursive relation:} \prod_{i=1}^{n-1}\widetilde{S}_{j_ij_n} = [d_{j_n}]^{n-2} \!\!\sum_{\ell_1,\cdots,\ell_n} \!\!\delta_{\ell_1,0}
\prod_{i=1}^{n-1} Y(j_i,\ell_i,\ell_{i+1}) \widetilde{S}_{\ell_n j_n}.\label{recursive}
\end{eqnarray}
The definition of the normalized Verlinde coefficient becomes clear from the normalization relation.
The recursive relation is a generalization of the fusion relation to any number of unitary irreducible representations
${\bf j}=(j_1,\cdots,j_p)$. The fusion and recursive relation are really easy to prove using
the graphical representations of the Verlinde coefficients. The proof of the fusion relation is given in the picture
(\ref{proof}); the proof of the recursive relation is done along exactly the same lines.
\begin{figure}[ht]
\psfrag{j1}{$j_1$}
\psfrag{j2}{$j_2$}
\psfrag{j3}{$j_3$}
\psfrag{l}{$\ell$}
\psfrag{=}{$=$}
\psfrag{A}{$[d_{j_3}]^{-1}$}
\psfrag{B}{$\sum_\ell Y(j_1,j_2,\ell)$}
\centering
\includegraphics[scale=0.5]{proof.eps}
\caption{Pictorial proof of the fusion relation. We start with the graph on the left. The two arrows are identities: the first one is obtained
applying the decomposition of the identity ``along the vertical dashed line''; the second one is obtained
applying the decomposition of the identity ``along the horizontal dashed line''. Both lead to equivalent expressions and the equality between the
evaluations of the graphs on the right is exactly the fusion relation. We made used of the identities represented in the picture (\ref{norm}).}
\label{proof}
\end{figure}
\medskip
Verlinde coefficients and their properties are particularly interesting to obtain useful formulae for the dimension of
the Hilbert space $H_k(j_1,\cdots,j_p)$.
Indeed, the dimension $N_k({\bf j})=\text{dim}(H_k(j_1,\cdots,j_p))$ is
\nopagebreak[3]\begin{equation}
N_k({\bf j}) \; = \; \sum_{\ell_1,\cdots,\ell_p} \delta_{\ell_1,0} \delta_{\ell_{p+1},0} \prod_{i=1}^p Y(\ell_i,j_i,\ell_{i+1})
\end{equation}
and can be expressed in terms of Verlinde coefficients using
the recursive relation (\ref{recursive}) combined with the normalization relation (\ref{normalization}). Some trivial
calculations lead to the expression:
\nopagebreak[3]\begin{eqnarray}
N_k({\bf j}) \; = \; \frac{2}{k+2} \sin^2(\frac{\pi}{k+2}) \sum_{\ell}[d_\ell]^{2-p} \prod_{i=1}^p \widetilde{S}_{j_i\ell}
\end{eqnarray}
which reduces, after using the explicit formula of Verlinde coefficients (\ref{explicitVerlinde}), to the following well-known formula:
\nopagebreak[3]\begin{equation}\label{dimension}
N_k({\bf j}) \; = \; \frac{2}{k+2} \sum_\ell (\sin(\frac{\pi d_\ell}{k+2}))^{2-p}\prod_{i=1}^p \sin(\frac{\pi d_\ell d_{j_i}}{k+2}) \;.
\end{equation}
\subsection*{2. Equivalent formulae for the Hilbert space dimension}
The expression (\ref{dimension}) for the dimension of the $SU(2)$ Chern-Simons Hilbert space is not very useful to compute the entropy of
a Black Hole. We propose here to give equivalent more interesting formulae.
\subsubsection*{2.1. Chern-Simons Hilbert space and random walk}
The fact that the coefficient $N_k({\bf j})$ are closely related to random walks have been noted and investigated in \cite{livi}
in the classical case, namely when $k$ becomes infinite. Here we show that even in the quantum case (i.e. for a finite $k$)
this link between Chern-Simons and random walk still exists and appears very interesting for the calculation of the entropy.
Some of our formulae have been derived in \cite{kaulma} where however only the classical case (infinite $k$) has been studied at the end.
For obvious reasons of notations, we will consider $\widetilde{N}_k({\bf d})\equiv N_{k-2}({\bf j})$ in the sequel.
We will also make use of the notations $d_i=d_{j_i}=2j_i+1$ and ${\bf d}=(d_1,\cdots,d_p)$.
\medskip
This Section is devoted to propose a random walk interpretation of the dimension $\widetilde{N}_k({\bf d})$.
For that purpose, we proceed in four steps.
\subsubsection*{a) $\widetilde{N}_k({\bf d})$ as a difference of $B_k({\bf d})$-type functions}
To do so, we first use the identities $\sin^2\theta= 1-\cos^2\theta$ and $\cos \theta =\sin(2\theta)/(2\sin^2\theta)$
in the formula
\nopagebreak[3]\begin{equation}
\widetilde{N}_k({\bf d}) \; = \; \frac{2}{k} \sum_\ell (\sin(\frac{\pi d_\ell}{k}))^{2}\prod_{i=1}^p
\frac{\sin(\frac{\pi d_\ell d_{j_i}}{k})}{\sin(\frac{\pi d_\ell}{k})}
\end{equation}
to write it as the difference
\nopagebreak[3]\begin{equation}\label{NfunctionofB}
\widetilde{N}_k({\bf d}) \; = \; B_k({\bf d}) - \frac{1}{4} B_k({\bf d_+})
\end{equation}
where the function $B_k$ which depends on a family of representations ${\bf d}=(d_1,\cdots,d_p)$ or
${\bf d_+}=(d_1,\cdots,d_p,2,2)$ reads
\nopagebreak[3]\begin{equation}\label{definitionofB}
B_k({\bf d}) \; = \; \frac{2}{k} \sum_{d=0}^{k-1} \prod_{\ell} \frac{\sin(\frac{\pi d}{k}d_\ell)}{\sin(\frac{\pi d}{k})}\,.
\end{equation}
The product runs over $\ell \in [0,p]$ or $\ell \in [0,p+2]$ depending whether we are considering $B_k({\bf d})$
or $B_k({\bf d_+})$.
Note that ${\bf d_+}$ is the union of the family of dimensions $\bf d$ with two more equal elements corresponding to the dimension
of the fundamental representation $d_{1/2}=2$.
\subsubsection*{b) Combinatorial expression of $B_k({\bf d})$}
Now we concentrate on the function $B_k({\bf d})$. In particular, we want to exhibit the fact, as in the classical case,
that $B_k({\bf d})$ admits a random walks interpretation. To show this is indeed the case,
we replace each term of the product in (\ref{definitionofB}) by the following expression:
\nopagebreak[3]\begin{equation}
\frac{\sin(\frac{\pi d}{k}d_\ell)}{\sin(\frac{\pi d}{k})} = e^{i\frac{\pi d}{k}(d_\ell -1)} \sum_{n_\ell=0}^{d_\ell -1} e^{-2i\frac{\pi d}{k}n_\ell}\;.
\end{equation}
As a result, the function (\ref{definitionofB}) can be rexpressed as follows:
\nopagebreak[3]\begin{eqnarray}
B_k({\bf d}) = \frac{2}{k} \sum_{d=0}^{k-1} \prod_{\ell =1}^p \sum_{n_\ell =0}^{d_\ell -1} e^{i\frac{\pi d}{k}(d_\ell -1 -2n_\ell)}
= \frac{2}{k} \sum_{d=0}^{k-1} \sum_{\{n_1,\cdots,n_p\}} e^{i\frac{\pi d}{k}(\Delta_p -2N)}
\end{eqnarray}
where we introduced the notations $\Delta_p=\sum_{\ell =1}^p(d_\ell -1)$ and $N=\sum_{\ell =1}^p n_\ell$. Note that
the second sum runs over families of integers $\{n_1,\cdots,n_p\}$ such that each component $n_i\in [0,d_i-1]$.
Permuting the two sums and summing over the variable $d$ lead to:
\nopagebreak[3]\begin{eqnarray}\label{Bk}
B_k({\bf d}) \; = \; \frac{2}{k} \sum_{\{n_1,\cdots,n_p\}} \frac{1 - e^{i\pi (\Delta_p -2N)}}{1 - e^{i\frac{\pi}{k}(\Delta_p - 2N)}} =
\frac{2}{k} \sum_{\{n_1,\cdots,n_p\}} \frac{1 - e^{i\pi \Delta_p}}{1 - e^{i\frac{\pi}{k}(\Delta_p - 2N)}}
\end{eqnarray}
To go further into the calculation, we distinguish the case where $\Delta_\ell$ is odd from the case where $\Delta_\ell$ is even.
\subsubsection*{c) $\Delta_\ell$ odd implies that $\widetilde{N}_k({\bf d})=0$}
The first case, $\Delta_\ell$ odd, is simpler. Indeed, in that case, $1-e^{i\pi\Delta_p}=2$ and then
the function $B_k({\bf d})$ reduces to the form:
\nopagebreak[3]\begin{equation}
B_k({\bf d}) \; = \; \frac{4}{k}\sum_{\{n_1,\cdots,n_p\}} \frac{1}{1 - e^{i\frac{\pi}{k}(\Delta_p - 2N)}} \,.
\end{equation}
From the beginning, we know that $B_k({\bf d})$ is a real-valued function and therefore it equals its real part, i.e.
$B_k({\bf d})={\cal R}_e(B_k({\bf d}))$ where ${\cal R}_e(z)$ denotes
the real part of $z \in \mathbb C$. Moreover, for any value of $\theta$ (different from $0[2\pi]$), the following equality holds:
\nopagebreak[3]\begin{equation}
\frac{1}{1-e^{i\theta}} = \frac{1}{2} + \frac{i}{2} \text{cotan} \frac{\theta}{2}\;.
\end{equation}
As a consequence, the expression of the function $B_k({\bf d})$ simplifies drastically and reduces to:
\nopagebreak[3]\begin{equation}
B_k({\bf d}) =\frac{4}{k} \frac{1}{2} \sum_{\{n_1,\cdots,n_p\}} 1 \; = \; \frac{2}{k} \prod_{\ell =1}^p d_\ell \;.
\end{equation}
Therefore, $B_k({\bf d_+})=4B_k({\bf d})$ and then the dimension of the Hilbert space (\ref{NfunctionofB}) vanishes
in that case. The meaning of this result is simple:
there is no invariant tensor in the tensor product $\otimes_\ell j_\ell$ when $\Delta_p =\sum_\ell (d_\ell -1)$ is odd.
As an example, let us consider the case where all the spins equal $1/2$: $\Delta_p=p$ odd means that there is an odd number
of spins; as expected there is no trivial representation in the tensor product of an odd number of $1/2$ representations.
\subsubsection*{d) $\Delta_\ell$ even: random-walk interpretation of $\widetilde{N}_k({\bf d})=0$}
The second case, $\Delta_p$ even, is far more interesting. In that situation, we would naively say that $B_k({\bf d})$ vanishes because
all the terms in the numerator of the formula (\ref{Bk}) are $1-e^{i\pi \Delta_p}=0$. But a more careful analysis shows that the
denominator can also lead to a singularity. As a result, the non-vanishing contributions to the sum (\ref{Bk}) are those where both
the numerator and denominator vanish. For this to happen, there must exist $s\in \mathbb Z$ such that $\Delta_p-2N=2sk$.
Therefore, we have:
\nopagebreak[3]\begin{equation}\label{random}
B_k({\bf d})=2 \sum_{\{n_1,\cdots,n_p\}}\delta_{\Delta_p -2N[2k]}
\end{equation}
where $\delta_{n[2k]}$ takes the value one if there exists an integer $s$ such that $n=2ks$, otherwise it is null.
Here comes the random walk interpretation of the dimension of the Chern-Simons Hilbert space. We proceed to a changing of
variables: instead of summing over non-negative integers $n_i \in [0,d_i-1]$, we sum over half-integers
$m_i = n_i - \frac{d_i-1}{2} \in [\frac{1-d_i}{2},\frac{1+d_i}{2}]$ (with $m_{i+1}=m_i+1$). Then, the formula (\ref{random}) becomes
\nopagebreak[3]\begin{equation}
B_k({\bf d})=2 \sum_{\{m_1,\cdots,m_p\}} \delta_{m_1+\cdots+m_p[k]}\,.
\end{equation}
A similar formula has been found in \cite{kaulma} and its classical counterpart has been established and studied in \cite{livi}.
As a result, the $B_k$ function appears to be the number of ways to start from the origin $0$ at the $\mathbb Z$ axe and come back at a point
$0[k]$ after $p$ steps $m_i$, each step being bounded as follows $m_i \in [\frac{1-d_i}{2},\frac{1+d_i}{2}]$. These functions have been deeply and
precisely studied in the domain of random walks. And this analogy was used as a central tool in \cite{livi} to obtain asymptotics behavior
of some entropy. In order to be a bit more explicit, we introduce the variable $r=[\Delta_p/(2k)]$ where $[x]$ is the floor function and then:
\nopagebreak[3]\begin{equation}
B_k({\bf d})=2 \sum_{\{m_1,\cdots,m_p\}} \sum_{q=-r}^r \delta_{m_1+\cdots+m_p - qk} \;.
\end{equation}
Let us recall now that we are interested in the number of states $\widetilde{N}_k$ and not in the function $B_k$ itself.
Using previous formulae, we have for $\widetilde{N}_k$ the expression:
\nopagebreak[3]\begin{eqnarray}
\widetilde{N}_k({\bf d}) = 2 \sum_{\{m_1,\cdots,m_p\}} \left( \sum_{q=-r}^r \delta_{m_1+\cdots+m_p - qk} -
\frac{1}{4}\sum_{a,b \in \{-\frac{1}{2},\frac{1}{2}\}} \sum_{q=-s}^s \delta_{m_1+\cdots+m_p+a+b - qk} \right) \nonumber
\end{eqnarray}
where $s=[(\Delta_p+2)/(2k)]$. It is clear that $s$ belongs to the set $\{r,r+1\}$ and to avoid complications
we assume that $r=s$. The case $s=r+1$ would introduce extra terms which are not important at all for what we want to do.
In that case, the previous formula simplifies and after summing over the variables $a$ and $b$ one obtains:
\nopagebreak[3]\begin{equation}\label{randomN}
\widetilde{N}_k({\bf d}) = \sum_{\{m_1,\cdots,m_p\}} \sum_{q=-r}^r ( \delta_{m_1+\cdots+m_p - qk} - \frac{1}{2}\delta_{m_1+\cdots+m_p - qk +1}
-\frac{1}{2}\delta_{m_1+\cdots+m_p - qk -1})\;.
\end{equation}
This expression generalizes the one obtained in the classical case (which corresponds in fact to $r=0$ in our notations).
It is useful to study the asymptotic behavior of the number of states and also to study the effect of a finite $k$.
\subsubsection*{2.2. Integral formula}
Very often, one identifies the number of states $\widetilde{N}_k$ to the dimension of the invariant tensors space
in the tensor product $\otimes_\ell j_\ell$
between representations of the classical group $SU(2)$. This is only true when the ratio $r=0$ which also coincides with the classical limit $k$ goes
to infinity. In that case, $\widetilde{N}_k=\widetilde{N}_\infty$ is expressed as an integral over $SU(2)$ conjugacy classes, or equivalently
over an angle $\theta$.
We proposed to generalize this integral formula to the case where $r \neq 0$.
For that purpose, we start with the relation:
\nopagebreak[3]\begin{eqnarray}
\delta_{m_1+\cdots+m_p + a} \; = \; \frac{1}{2\pi} \int_0^{2\pi}d\theta \; e^{i\theta(m_1+\cdots+m_p+a)}
\end{eqnarray}
defined for any integer $a \in \mathbb Z$. It easily leads to the relation:
\nopagebreak[3]\begin{eqnarray}
\sum_{\{m_1,\cdots,m_p\}}\delta_{m_1+\cdots+m_p + a} \; = \; \frac{1}{2\pi} \int_0^{2\pi}d\theta
\cos(a\theta) \prod_{\ell=1}^p \frac{\sin(d_\ell\frac{\theta}{2})}{\sin \frac{\theta}{2}}\;.
\end{eqnarray}
where the sum runs over $m_\ell \in [\frac{1-d_\ell}{2},\frac{1+d_\ell}{2}]$.
Using this last identity and after some trivial calculations, one shows that the number of states is given by the integral:
\nopagebreak[3]\begin{equation}
\widetilde{N}_k({\bf d}) \; = \; \frac{1}{\pi} \int_0^{2\pi} d\theta \; (\sum_{q=-r}^r \cos(\theta qk)) \sin^2(\frac{\theta}{2}) \prod_{\ell=1}^p \frac{\sin(d_\ell\frac{\theta}{2})}{\sin \frac{\theta}{2}}\;.
\end{equation}
One more simplification occurs due to the trigonometric identity
\nopagebreak[3]\begin{equation}\label{identitytrigo}
\sum_{q=-r}^r \cos(\theta qk) = 1 + 2 \sum_{q=1}^r\cos(qk\theta) = \frac{\sin((r+\frac{1}{2})k\theta)}{\sin \frac{k\theta}{2}} \;.
\end{equation}
As a result, the number of states $\widetilde{N}_k$ takes the form:
\nopagebreak[3]\begin{equation}\label{generalformula}
\widetilde{N}_k({\bf d}) \; = \; \frac{1}{\pi} \int_0^{2\pi} d\theta \; \sin^2(\frac{\theta}{2})
\frac{\sin((r+\frac{1}{2})k\theta)}{\sin \frac{k\theta}{2}} \prod_{\ell=1}^p \frac{\sin(d_\ell\frac{\theta}{2})}{\sin \frac{\theta}{2}}\;.
\end{equation}
This formula generalizes, as announced in the introduction of that section,
the classical one. We see, as expected, that $\widetilde{N}_k({\bf d})$ coincides with the classical formula when $r=0$,
i.e. when $\Delta_p < 2k$. This particular case can be recovered from different arguments: if $\Delta_p <2k$, then each representations in the tensor
product $\otimes_\ell j_\ell$ has a spin $s <k/2$ and therefore one never sees the effect of the cut-off $k$.
When the condition $r=0$ is not satisfied the integral formula defining the number of states differs from the classical one by a different integration measure on the $SU(2)$ conjugacy class. The presence of this new measure might have an effect on the black hole entropy.
\subsubsection*{2.3. An example: all spins equal $1/2$}
To get an intuition of previous formulae, we consider a particular example:
we assume that $d_\ell=2$ for all punctures $\ell \in \{1,\cdots,p\}$, i.e. all the spins equal $1/2$.
Furthermore, we assume that $\Delta_p=p$ is even.
\subsubsection*{a) Classical case: $r=0$ \cite{livi,kaulma}.}
To start with, we also consider first the case $\Delta_p <2k$, i.e. $r=0$ in the previous
notations. This case has been studied deeply in \cite{livi}.
From the random walk expression of the number of states (\ref{randomN}),
one obtains that $\widetilde{N}_k(2,\cdots,2) \equiv {N}_{2}(p)$ is given by:
\nopagebreak[3]\begin{equation}\label{classical1/2}
{N}_{2}(p) \; = \; \sum_{\{m_1,\cdots,m_p\}}(\delta_{m_1+\cdots+m_p} - \delta_{m_1+\cdots+m_p+1}) \; = \;
\binom{p}{p/2} - \binom{p}{p/2-1} \;.
\end{equation}
We have omitted to mention $k$ because $\widetilde{N}_k$ does not in fact depend on $k$ when $r=0$.
The last equality is a result of a trivial combinatorial analysis: given an integer $a \leq p/2$, the number of ways to have
$n_1+\cdots+n_p=a$ where $n_i\in \{-1/2,1/2\}$ is given by the number
of ways to choose $(p/2+a)$ elements from a set of $p$ elements which is precisely given by the binomial coefficient $\binom{p}{p/2+a}$.
From the expression of the binomial coefficients in terms of factorials, one ends up with the following formula:
\nopagebreak[3]\begin{equation}
{N}_{2}(p) \; = \; \frac{2}{p+2}\frac{p!}{((p/2)!)^2}
\end{equation}
We obtain an exact combinatoric formula for the number of states in that particular case. The asymptotic of ${N}_2(p)$
is therefore straightforward to obtain from the stirling formula which states that:
\nopagebreak[3]\begin{equation}\label{stirling}
p! \; \sim \; \sqrt{2\pi p} (p/e)^p \;\;\;\; \text{for large values of $p$.}
\end{equation}
Using this very well-known result, one shows the following asymptotic behavior:
\nopagebreak[3]\begin{equation}\label{F2}
{N}_2(p) \; \sim \; \sqrt{\frac{8}{\pi}}p^{-3/2} 2^p \;.
\end{equation}
This formula coincides with the one found in \cite{kaulma} from different arguments. In particular, we recover the same leading order and
the same sub-leading corrections to the ``entropy":
\nopagebreak[3]\begin{equation}
S_{1/2}(p) \; \equiv \; \log{N}_2(p)\; = \; 2^p \; - \; \frac{3}{2} \log p \; + \; {\cal O}(1).
\end{equation}
\subsubsection*{b) Quantum corrections: $r>0$.}
Let us now relax the condition that $\Delta_p<2k$, i.e. $r=[p/(2k)]$ is now a non-zero integer. In that case, it is
a bit more involved to obtain a combinatoric expression for the number of states but, using similar arguments as previously,
one can show that:
\nopagebreak[3]\begin{equation}
{N}_2(p) \; = \; \sum_{q=-r}^r \binom{p}{p/2-qk} -\frac{1}{2}\binom{p}{p/2-qk-1} - \frac{1}{2} \binom{p}{p/2-qk+1}\;.
\end{equation}
Note that we still omit to mention explicitly the dependence in $k$ even if now $N_2(p)$ does depend on $k$.
To go further, we separate the $q=0$ contribution from the others in the sum and, using trivial symmetries properties of binomial coefficients, we get:
\nopagebreak[3]\begin{eqnarray}
{N}_2(p) & = & \binom{p}{p/2} - \binom{p}{p/2-1} \nonumber \\
& & +\sum_{q=1}^r 2\binom{p}{p/2-qk} -\binom{p}{p/2-qk-1} - \binom{p}{p/2-qk+1}\,.
\end{eqnarray}
Thus, we obtain a correction to the previous classical case (\ref{classical1/2}) and each term (for a given value of $q \in [1,r]$)
in the remaining sum is a finite linear
combination of binomial coefficients which reduces to the following form after some calculations:
\nopagebreak[3]\begin{equation}\label{remainingterms}
2\frac{1-p-2q^2k^2}{(p/2+1)^2 -q^2k^2} \binom{p}{p/2-qk}\;.
\end{equation}
As a result, the asymptotic behavior of each term in the previous sum is governed by the asymptotic behavior of the binomial coefficient
$\binom{p}{p/2-qk}$. We are interested in the asymptotic for a large value of $p$ but also a large value of $k$ such that these two numbers have
the same scaling namely $p/k$ remains constant. Indeed, in the black hole context, both $p$ and $k$ are proportional to the area of the horizon
which tends to infinity (in unit of Planck area). Thus, the leading term in the asymptotic expansion of (\ref{remainingterms}) is given by
the leading term of the expansion of binomial coefficient of the form $\binom{p}{\alpha p}$ for large values of $p$ where $\alpha \in [0,1/2[$.
From Stirling formula (\ref{stirling}), one shows that
\nopagebreak[3]\begin{equation}
\binom{p}{\alpha p} \sim \frac{1}{\sqrt{2\pi \alpha(1-\alpha)p}}g(\alpha)^{-p} \;\;\; \text{with} \;\;\; g(\alpha)=
\alpha^\alpha(1-\alpha)^{1-\alpha}\;.
\end{equation}
It is straightforward to show that the function $g$ satisfies the bound $g(\alpha) > 1/2$ and therefore the previous binomial
coefficient grows like $g(\alpha)^{-p} < 2^p$. As a consequence, the ``classical term" dominates the asymptotics in the sense that:
\nopagebreak[3]\begin{equation}
{\binom{p}{\alpha p}}/{\binom{p}{p/2}} \; \sim \; (2g(\alpha))^{-p} \; \rightarrow \; 0
\end{equation}
for $\alpha<1/2$. This result shows that the quantum corrections (due to the finiteness of $k$) do not affect the asymptotic
expansion of ${N}_2(p)$ neither at the leading neither at the subleading order.
\subsection*{3. Entropy of the $SU(2)$ Black Hole}
Here we adapt the techniques used in \cite{counting3} and firstly introduced in \cite{counting1} to compute the entropy
of a black hole. We are not going into the mathematical details of these techniques which has been very well exposed
in \cite{counting3} and which are in fact very well-known in the domain of probabilities and used
to understand some properties of random walks.
We propose to reproduce these results in a more ``intuitive" or physical way: we will omit many mathematical details
which appear in a first time non necessary. In particular, we show that it is not necessary to go to complex analysis and number theory
to get the asymptotic behavior of the entropy.
Before going to the details of the entropy calculation, let us briefly recall how black holes are described
in loop quantum gravity and how we compute the entropy.
In the context of LQG, a local definition of a black hole is introduced through the concept of {\em isolated horizons} (IH), regarded as a sector of the phase-space of GR containing a horizon in equilibrium with the external matter and gravitational degrees of freedom. This local definition is used for the black-hole entropy calculation since the quantization of such a system allows to define a Hilbert space which is the tensor product of a boundary and a bulk terms.
The entropy of the IH is then computed by the formula $S={\rm tr}(\rho_{\scriptscriptstyle IH}\log\rho_{\scriptscriptstyle IH})$, where the density matrix
$\rho_{\scriptscriptstyle IH}$ is obtained by tracing over the bulk d.o.f., while restricting to horizon states that are compatible
with the macroscopic area parameter $a$. Assuming that there exists at least
one solution of the bulk constraints for every admissible state on the boundary, the entropy is given by
$S=\log(N(a))$ where $N(a)$ is the number of horizon states. Since the theory on the horizon is associated to Chen-Simons theory with punctures, the entropy calculation problem boils down to the counting, in the large horizon area limit, of the dimension of the Hilbert space (\ref{Hilbert}), which, for a given configuration of punctures spins ${\bf d}=(d_1,\cdots,d_p)$, is expressed by the formula (\ref{generalformula}).
\subsubsection*{3.1. The Laplace transform method: basic idea}
The Laplace transform method allows in certain cases to obtain the asymptotic behavior of a function
$F(p)$ for large values of $p$.
For simplicity, we assume that the function $F$ is defined for $p$ integers
but the method applies in the case where $F$ is a function of a real number $x$.
The idea consists first in considering the Laplace transform $\widetilde{F}(s)$ defined a priori for $s \geq 0$ by:
\nopagebreak[3]\begin{equation}\label{Laplace}
\widetilde{F}(s) \; = \; \sum_{p=0}^\infty e^{-ps} F(p) \;.
\end{equation}
The Laplace transform appears as a series and therefore might be not defined at all or it might be defined for some values of the
real positive variable $s$ only. To understand if the series is convergent or divergent, one looks
at the asymptotic behavior of $F(p)$ at large $p$. To be more concrete, let us propose some examples.
\begin{enumerate}
\item If $F(p)\sim p^\alpha$ for some real number $\alpha$, then the series (\ref{Laplace}) is convergent for any values of $s$.
\item If $F(p)\sim e^{\alpha p^2}$ for some positive real number $\alpha$, then the series (\ref{Laplace}) is divergent for any values
of $s$ and then the Laplace transform is never defined. On the contrary, if $\alpha$ is negative, then the series in convergent
and the Laplace transform is well defined for all values of $s$.
\item If $F(p) \sim e^{\alpha p}$ for some positive real number $\alpha$, then the series is convergent for $s>\alpha$. In the case
where $s<\alpha$, the series diverges and therefore is ill-defined. The case $s=\alpha$ is critical: the convergence properties of $\widetilde{F}$
in that regime depends on the subleading behavior of $F(p)$.
\end{enumerate}
We are particularly interested in the last case because in the context of black hole entropy the number $N(a)$ of microstates corresponding
to a given macroscopic area $a$ is exponential in $a$, namely $N(a) \sim e^{\alpha a}$ and all the problem is to find the coefficient $\alpha$.
Let us now come back to the general discussion.
If we find $\alpha>0$ such that $\widetilde{F}(s)$ is defined for $s>\alpha$ and undefined for $s<\alpha$, then we conclude that
$F(p)\sim e^{\alpha p}$ for large $p$. This is more a physical argument than a rigorous proof because we assume the asymptotic
behavior of $F(p)$. However, it is also possible to prove rigorously the asymptotic behavior as it was done in \cite{counting1}
and \cite{counting3} where the reader can find the mathematical details.
In the language of probabilities, $s_c=\alpha$ is called the critical exponent of $F(p)$.
Before going further, let us recall that, in the particular case where $F(p)\sim e^{\alpha p}$, it is possible to invert
the Laplace transform and to recover to function $F(p)$ from $\widetilde{F}(s)$ according to:
\nopagebreak[3]\begin{equation}
F(p) = \frac{e^{s_0 p}}{2\pi}\int_0^{2\pi} dx \; \widetilde{F}(s_0+ ix) e^{ixp}
\end{equation}
where $s_0>\alpha$ is a real number.
\medskip
In fact, it is possible to extend this technique to obtain also the subleading terms in the asymptotics expansion of $F(p)$.
To understand this point, we assume that $F$ behaves as follows $F(p)\sim e^{\alpha p}Q(p)$ where $Q(p)$ is an algebraic function
whose dominant term at large $p$ is $Q(p)\sim p^{\beta}$ where $\beta$ is a real number. We generalize the Laplace transform
to the following function of the two real variables $s$ and $t$:
\nopagebreak[3]\begin{equation}\label{generalLaplace}
\widetilde{F}(s,t) = \sum_{p=1}^\infty e^{-ps} \, p^{-t} \, F(p)\;.
\end{equation}
When $t=0$, this function coincides with the standard Laplace transform up to the constant $F(0)$. The point is to evaluate it at the critical value
$s_c=\alpha$. Indeed, $e^{-\alpha p}F(p) \sim p^{\beta}$ and therefore the convergence properties of the series $\widetilde{F}(\alpha,t)$ depends on
the asymptotic behavior of $p^{\beta-t}$: if $\beta-t>-1$, the series is divergent; if $\beta - t<-1$, the series is convergent. Therefore, we proceed
as we do to obtain $\alpha$, we define the critical value $t_c$ as the minimal value of $t$ such that $\widetilde{F}(\alpha,t)$ is well-defined.
Thus, the coefficient $\beta$ is fixed by $\beta=t_c-1$.
Of course, we can repeat this technique
to obtain recursively all the corrections to the leading term in the asymptotic expansion of $F(p)$. But, for this method to work,
we must know the form of the asymptotic behavior of the function $F(p)$. Once we know that the function $F(p)\sim e^{\alpha p} p^\beta$
at large $p$, our method allows to obtain the critical exponents $\alpha$ and $\beta$.
\subsubsection*{3.2. A simple application of the Laplace transform method}
Let us show that we can use this very simple technique to obtain the asymptotic behavior of the number of states $\widetilde{N}_k({\bf d})$
when all the spins are equal. To that aim, we introduce the notation $F_d(p)=\widetilde{N}_k({\bf d})$ with $j_1=\cdots=j_p=j$
and $d=2j+1$:
\nopagebreak[3]\begin{equation}\label{pipo}
F_d(p) \; = \; \int_0^\pi d\theta \; \mu_k(\theta) \left(\frac{\sin (d\theta)}{\sin \theta}\right)^p
\end{equation}
where $\mu_k(\theta)$ is a continuous non singular function obtain directly from (\ref{generalformula}):
\nopagebreak[3]\begin{equation}
\mu_k(\theta) \; = \; \frac{2}{\pi}\, \sin^2\theta \, \frac{\sin((2r+1)k\theta)}{\sin(k\theta)}\;.
\end{equation}
We omit to mention the dependence in $k$ of $\mu_k(\theta)$ and of $F_d(p)$ for clarity reasons.
Furthermore, we know, from random walks arguments, that the asymptotics of $F_d(p)$ is dominated by its classical part only;
in other words, we consider $r=0$ in the sequel. We will discuss the quantum corrections later.
Now, we assume that $F_d(p) \sim e^{\alpha p}$ for large values of $p$.
If the asymptotic behavior assumption is true, then the Laplace transform
$\widetilde{F}_d(s) = \sum_{p=0}^\infty e^{-sp}F(d,p)$
is well-defined for $s > \alpha$ and not-defined for $s<\alpha$. To obtain the critical exponent $\alpha$, we need to simplify the expression
of $\widetilde{F}_d(s)$. To do so, we exchange the sum defining $\widetilde{F}_d$ and the integral over $\theta$ in the definition of $F_d$
(\ref{pipo}). We obtain the following expression:
\nopagebreak[3]\begin{equation}\label{integralF}
\widetilde{F}_d(s) \; = \; \frac{2}{\pi}\int_0^\pi d\theta \, \sin^2\theta \, \sum_{p=0}^\infty
\left(e^{-s} \frac{\sin (d\theta)}{\sin \theta}\right)^p
\end{equation}
Then, assuming that $s$ is large enough, we perform the sum over $p$ and then:
\nopagebreak[3]\begin{eqnarray}
\widetilde{F}_d(s) = \frac{2}{\pi} \int_0^\pi d\theta \, \sin^2\theta \, Z(s,\theta) \;\;\; \text{with} \;\;\;
Z(s,\theta) = \left( 1 - e^{-s}\frac{\sin (d\theta)}{\sin \theta}\right)^{-1} \nonumber \\
\end{eqnarray}
The next step consists in analyzing the structure of the singularities of $\widetilde{F}_d(s)$. It is clear that its singularities
come from the poles of the function $Z(s,\theta)$ viewed as a function of $\theta$.
We immediately see that $Z(s,\theta)$ admits a pole (viewed as a function of $\theta$)
if and only if $e^{-s} \geq d^{-1}$, i.e. $s \leq \log d$. More precisely, we have the following:
\begin{enumerate}
\item When $s=\log d$, then $Z(s,\theta)$
admits an unique pole which is $\theta_0=0$.
\item When $s < \log d$, then $Z(s,\theta)$ admits also at least one pole $\theta_0 \neq 0[\pi]$. At the vicinity of $\theta_0$, the
(inverse of the) function $Z$
behaves as follows:
$$Z^{-1}(s,\theta_0+\varepsilon) \simeq -\varepsilon(d\text{cotan}(d\theta_0) - \text{cotan}(\theta_0)).$$
As a consequence, the integral (\ref{integralF}) is divergent.
\end{enumerate}
As expected, $\widetilde{F}(d,s)$ is defined only for $s>\log d$. Therefore, $s_c=\log d$ is the
critical exponent and we have the asymptotic behavior:
\nopagebreak[3]\begin{equation}
F_d(p) \; \sim \; d^p \;.
\end{equation}
Before computing the sub-leading corrections, let us make some important remarks.
In the first remark, we come back to the exchange of the sum over $p$ and the integral over $\theta$ in the computation
of the Laplace transform (\ref{integralF}). This step can be justified in our case but this is not always the case.
More precisely, the exchange makes sense if the sum over $p$ is defined, namely if $\vert e^{-s}\sin(d\theta)/\sin\theta \vert <1$
for all $\theta$. This is exactly the condition we obtained to compute the critical exponent.
In the second and last remark, we come back to the ``quantum corrections" of $F_d(p)$. We know that the number of states is given by:
\nopagebreak[3]\begin{eqnarray}
F_d(p) & = & \frac{2}{\pi}\int_0^\pi d\theta \, \sin^2\theta \left(\frac{\sin (d\theta)}{\sin \theta}\right)^p \nonumber\\
& & +\frac{4}{\pi}\sum_{q=1}^r \int_0^\pi d\theta \, \sin^2\theta \, \cos(qk\theta)
\left(\frac{\sin (d\theta)}{\sin \theta}\right)^p \;.
\end{eqnarray}
The first term is the classical contribution and the rest are the quantum corrections. We want to compute the asymptotic behavior
of these corrections at the large $p$ limit, and we take,
at the same time, $k$ large as well with a fixed ratio $\rho=p/2k$.
As a consequence the number $r=[\rho]$ remains fixed in this limit. Therefore, the contribution of $F_d(p)$ corresponding to $q \neq 0$ reads:
\nopagebreak[3]\begin{equation}
\frac{4}{\pi}\sum_{q=1}^r \int_0^\pi d\theta \, \sin^2\theta \, \cos(\frac{q\theta}{2\rho} p)
\left(\frac{\sin (d\theta)}{\sin \theta}\right)^p \;.
\end{equation}
The calculation of the Laplace transform is more involved
in that situation due to the presence of a fast oscillating function in the integrand. A consequence is that the naive exchange of the infinite sum
over $p$ and the integral over $\theta$ is not justified.
\medskip
Now, we go further and compute the sub-leading terms. For that purpose, we concentrate on the classical contribution only (formally we take $t=0$)
and we define the general Laplace transform of $F_d(p)$ (\ref{generalLaplace}):
\nopagebreak[3]\begin{equation}
\widetilde{F}_d(s,t) \; = \; \sum_{p=0}^\infty e^{-ps} \, p^{-t} \, F_d(p)\;.
\end{equation}
Permuting the integral with the sum leads to the expression:
\nopagebreak[3]\begin{eqnarray}\label{generalizedexample}
\widetilde{F}_d(s,t) = \int_0^\pi \mu_k(\theta) \sum_{p=0}^\infty p^{-t} \left( e^{-s} \frac{\sin(d\theta)}{\sin \theta}\right)^p
= \int_0^\pi \mu_k(\theta) \, \text{Li}_t\left(\frac{\sin (d\theta)}{d\sin \theta}\right)\,.
\end{eqnarray}
Indeed, we recognize the polylogarithm function $\text{Li}_t(z)$ defined for any couple of complex numbers $(t,z)$ such that
$\vert z \vert <1$ by the series:
\nopagebreak[3]\begin{equation}
\text{Li}_t(z) \; = \; \sum_{p=1}^\infty p^{-t} \, z^p \;.
\end{equation}
Following the general idea we described above, we first evaluate $\widetilde{F}_d(s,t)$ at the critical value $s_c=\log d$.
Then, we look for the critical value $t_c$ such that $\widetilde{F}_d(s_c,t)$ is well-defined for $t>t_c$ but not defined if $t<t_c$.
We know that $\widetilde{F}_d(s_c,t)$ might be not defined because of the singularity of the integrand
$\text{Li}_t\left(\frac{\sin (d\theta)}{d\sin \theta}\right)$ at $\theta=0$. To compute $t_c$, we analyze the behavior of the integrand
around $\theta=0$:
\begin{eqnarray*}
\text{Li}_t\left(\frac{\sin (d\theta)}{d\sin \theta}\right) & \sim & \text{Li}_t(\theta^2\frac{1-d^2}{6}) \\
& \sim & \text{Li}_t(e^{\frac{1-d^2}{6}\theta^2}) \\
& \sim & \Gamma(1-t) \left(\frac{d^2-1}{6}\right)^{t-1} \theta^{2(t-1)}
\end{eqnarray*}
where $\Gamma(t)$ is the Gamma function. We used some asymptotic properties of the polylogarithm function. As a consequence, at the vicinity
of $\theta=0$, the integrand of (\ref{generalizedexample}) behaves as:
\nopagebreak[3]\begin{equation}
\mu_k(\theta) \, \Gamma(1-t) \left(\frac{d^2-1}{6}\right)^{t-1} \theta^{2(t-1)} \; \sim \; \frac{2}{\pi} \, \Gamma(1-t) \left(\frac{d^2-1}{6}\right)^{t-1}\theta^{2t}
\end{equation}
because $\mu_k(\theta)\simeq \frac{2}{\pi} \theta^2$ for $\theta$ small.
Therefore, the integral over $\theta$ is defined when $2t>-1$ and not defined when $2t<-1$; then the
critical value of $t$ is $t_c=-1/2$. As a consequence, we obtain the value
of the second critical exponent
$$\beta = -1 + t_c=-3/2$$
which is independent of the dimension $d$. Finally, we can establish that:
\nopagebreak[3]\begin{equation}
F_d(p) \, \sim \, d^p p^{-3/2} \;\;\;\; \text{then} \;\;\;\; \log F_d(p) = p \log d -\frac{3}{2} \log p + {\cal O}(1)\;.
\end{equation}
In particular, we recover from another method the asymptotic behavior of $F_2(p)$ given in (\ref{F2}). This asymptotics has been obtained in
\cite{livi} from random walks arguments. The case $d=2$ has also been consider in \cite{kaulma}.
\subsubsection*{3.3. Asymptotic of the entropy}
We apply, the method illustrated above to compute the asymptotic behavior of a spherically symmetric and a distorted $SU(2)$ black hole
entropy in loop quantum gravity.
\subsubsection*{a) The spherically symmetric Black Hole}
The calculation of the entropy of the spherically symmetric $SU(2)$ black hole has been done precisely in \cite{counting3}
and has been investigated earlier in \cite{kaulma} when the level in infinite. In \cite{su22, su23}, we have shown that
the level $k$ and the area $a$ can be considered as independent variables. For that reason,
we are going to reformulate the results obtained in \cite{counting3} when $k$ is finite focussing on physical arguments and avoiding the
number theory and the complex analysis aspects. These aspects are not necessary to get the main ideas and the main results
if one assumes that the number of states grows exponentially with the area.
\medskip
The entropy $S(a)=\log N(a)$ of a spherically symmetric black hole of macroscopic (adimensionalized) area $a$ is defined from
the number of states
\nopagebreak[3]\begin{equation}\label{SSBH}
N(a) = \sum_{p=0}^\infty \sum^{k+1}_{d_1,\cdots,d_p} \delta(a-\frac{\sum_{\ell=1}^p\sqrt{(d_\ell-1)(d_\ell +1)}}{2}) \widetilde{N}_{k}({\bf j})\,.
\end{equation}
The finiteness of the level $k$ appears in two different places: in the sums which run from $2$ to $k+1$ and in the expression
of $\widetilde{N}_{k}({\bf j})$. It is important to note again that we will consider $k$ and $a$ as independent variables
and we will study the entropy for large $a$ but finite $k$.
This consideration allows us to define the Laplace transform of $N(a)$
\nopagebreak[3]\begin{equation}
\widetilde{N}(s) = \int_0^\infty da\, e^{-as}N(a) \,.
\end{equation}
One difference with the previous Section is that now the variable $a$ is continuous.
After some calculations
and assuming that $s$ is large enough, we end up with the following expression:
\nopagebreak[3]\begin{eqnarray}
\widetilde{N}(s) & = & \sum_{p=0}^\infty \sum^{k+1}_{d_1,\cdots,d_p} \int_{0}^\pi d\theta \, \mu_k(\theta)
\left( \prod_{\ell=1}^p \frac{\sin(d_\ell \theta)}{\sin \theta} e^{-\frac{s}{2} \sqrt{(d_\ell -1)(d_\ell +1)}}\right)\\
&=& \sum_{p=0}^\infty \int_0^\pi d\theta \,\mu_k(\theta) \left(\sum^{k+1}_{d=2}\frac{\sin d\theta}{\sin \theta} e^{-\frac{s}{2} \sqrt{(d-1)(d+1)}}\right)^p\\
&=& \int_0^\pi d\theta \, \mu_k(\theta) \left( 1 - \sum_{d=1}^k\frac{\sin (d+1) \theta}{\sin \theta} e^{-\frac{s}{2} \sqrt{d(d+2)}}\right)^{-1}
\end{eqnarray}
Again, we exchanged the sums over the variables $d_\ell$ and $p$ with the integral over $\theta$.
We proceed as in the previous section and conclude that the critical value of $s$ is the highest zero of the function
\nopagebreak[3]\begin{equation}
1 - \sum_{d=1}^k \frac{\sin (d+1)\theta}{\sin \theta} e^{-\frac{s}{2} \sqrt{d(d+2)}}
\end{equation}
which is reached for $\theta=0$. Therefore, the critical exponent $\alpha$ is the unique solution of the equation
\nopagebreak[3]\begin{equation}\label{barberosum}
1 - \sum_{d=1}^k (d+1) \, e^{-\frac{\alpha}{2} \sqrt{d(d+2)}} \; = \; 0 \,.
\end{equation}
With that definition, $\alpha$ depends on the level $k$.
For increasing values of the level $k$, the solutions $\alpha$ of the previous equation reach fast an asymptotic value as plotted in Figure \ref{Fig:s-kSpheric}.
\begin{figure}[ht]
\centering
\includegraphics[scale=1]{s-kSpheric.eps}
\caption{Increasing critical values of the exponent $\alpha $ for different $k\in \mathbb N$. Already for $k\geq 4$ an asymptotic value is reached.}
\label{Fig:s-kSpheric}
\end{figure}
The asymptotic value coincides as expected with the value $\alpha_{\infty}$ found in \cite{counting3} when $k\rightarrow \infty$.
Furthermore, we can estimate how the difference $\Delta \alpha= \alpha_\infty - \alpha$ decreases when $k$ increases.
$\Delta \alpha$ decreases exponentially with $k$ in the sense that it exits two real positive constants $A$ and $B$
such that
$$
\vert \alpha_k - \alpha_\infty \vert \; < \; A \, e^{-Bk} \,.
$$
\medskip
Next, we compute the subleading corrections. The difficulty of this problem is that we could not find a way
to put the generalized Laplace transform in a ``suitable" form as it was the case for the "toy-example"
we considered above. We will nonetheless circumvent this difficulty as follows.
First, we will evaluate $\widetilde{N}(s)$ at the vicinity of the critical value $\alpha$, i.e. $s=\alpha + \varepsilon$ for
$\varepsilon$ small, and then we will compute the generalized Laplace transform $\tilde{N}(\alpha,-t)$ at the critical value $\alpha$
as follows:
\nopagebreak[3]\begin{equation}
\widetilde{N}(\alpha,-t) \; = \; \frac{\partial^{t} \widetilde{N}(\alpha + \varepsilon)}{\partial \varepsilon^t}\vert_{\varepsilon=0}\;.
\nonumber
\end{equation}
assuming that $t$ is an integer. Finally, we will see that it makes sense to extend the obtained formula to half-integers (and also
real numbers in fact) which will allows us to extract the sub-leading corrections to the entropy.
\medskip
Let us start, as announced, by the following calculation:
\nopagebreak[3]\begin{eqnarray}\label{bla}
\widetilde{N}(\alpha + \varepsilon) \; \simeq \;
\int_0^\pi d\theta \, \mu_k(\theta) \left( 1 - \sum_{d=1}^{k} \frac{\sin (d+1)\theta}{\sin \theta} e^{-\frac{\alpha}{2} \sqrt{d(d+2)}}
(1-\frac{\varepsilon}{2} \sqrt{d(d+2)})\right)^{-1}
\end{eqnarray}
We know that the singularity of $\widetilde{N}(s+\varepsilon)$ when $\varepsilon$ goes to zero is due to the singularity of the integrand
$$f_\varepsilon(\theta) \; = \; \mu_k(\theta) \left( 1 - \sum_{d=1}^{k} \frac{\sin (d+1)\theta}{\sin \theta} e^{-\frac{\alpha}{2} \sqrt{d(d+2)}}
(1-\frac{\varepsilon}{2} \sqrt{d(d+2)})\right)^{-1}$$
in (\ref{bla}) when $\theta$ goes to zero. Therefore, we concentrate on the behavior of $f_\varepsilon(\theta)$
at the vicinity of $\theta=0$:
\begin{eqnarray*}
f_\varepsilon(\theta) \; \simeq \; (2r+1) \frac{2\theta^2}{\pi} \left( \sum_{d=1}^{k} (\varepsilon (d+1)\sqrt{d(d+2)} + \frac{\theta^2}{6}d(d+1)(d+2)) e^{-\frac{\alpha}{2}\sqrt{d(d+2)}}\right)^{-1}
\end{eqnarray*}
Now, we start from this expression to study the singularities of the generalized Laplace transform $\widetilde{N}(\alpha,t)$
evaluated at the critical value. Indeed, when $t$ is a positive integer, we can compute:
\nopagebreak[3]\begin{equation}\label{generalizedLaplace}
\widetilde{N}(\alpha,-t) \; = \; \frac{\partial^{t} \widetilde{N}(\alpha + \varepsilon)}{\partial \varepsilon^t}\vert_{\varepsilon=0}\;
\end{equation}
The last quantity is expressed as an integral over the variable $\theta$ whose eventual singularity is due to the behavior of the integrand around
$\theta=0$ given by:
\nopagebreak[3]\begin{eqnarray}
\frac{\partial^{t} f_\varepsilon(\theta) \vert_{\varepsilon=0}}{\partial \varepsilon^t} & \simeq &
(-1)^t t!\frac{2\theta^2}{\pi} \frac{ \left( \sum_{d=1}^k (d+1)\sqrt{d(d+2)}e^{-\frac{\alpha}{2} \sqrt{d(d+2)}}\right)^t}{\left( \frac{\theta^2}{6} \sum_{d=1}^k d(d+1)(d+2) e^{-\frac{\alpha}{2} \sqrt{d(d+2)}}\right)^{t+1}} \nonumber\\
& \sim & \theta^2 \, \frac{1}{\theta^{2(t+1)}} = \theta^{-2t}\;.\nonumber
\end{eqnarray}
We assume that the behavior of the integrand of $\widetilde{N}(\alpha,-t)$ remains the same even if $t$ is any real number.
As a consequence, $\widetilde{N}(\alpha,t)$ is singular when $2t > 1$ and the critical value of $t$ is $t_c=1/2$.
Then, the critical exponent $\beta=-t_c-1=-3/2$ and we recover the asymptotic expansion \cite{counting3}:
\nopagebreak[3]\begin{equation}
N(a) \; \sim \; e^{\alpha a} \, a^{-3/2} \;\;\;\; \text{for large $a$},
\end{equation}
with $\alpha$ given as in Figure \ref{Fig:s-kSpheric}.
Again, the finiteness of $k$ does not modify the sub-leading corrections when $k$ is large.
\medskip
Let us finish this section by two remarks. The first one concerns the effect of the finiteness of the level $k$ in the entropy.
As we have just seen, $k$ does modify the leading but does not modify the subleading corrections of the entropy. In that sense,
the logarithmic corrections seems to be universal and independent of the Immirzi parameter even in the $SU(2)$ spherically symmetric
black hole.
The second one concerns the techniques
we used: our calculations have to be viewed more as ``physical" arguments than rigorous proofs
of the asymptotic behavior of the entropy.
The nice point is that we can recover the ``right'' results very easily without entering too much into technical aspects.
\subsubsection*{b) The distorted Black Hole}
In the distorted case \cite{su23}, the black hole is described in terms of two commuting Chern-Simons theories
associated to the same level $k$. As in the symmetric case, the area $a$ of the black hole and the level $k$ are
considered as independent variables. At the fundamental level, the description of the distorted black hole in terms
of microstates is rather different from the description of the symmetric black hole. Indeed, each puncture colored
with a $SU(2)$ representation $j$ coming from the bulk decomposes into two $SU(2)$ representations $j^+$ and
$j^-$ when it crosses the black hole. A macroscopic state is therefore characterized by the number $p$ of punctures and a
family $(j_\ell,j^+_\ell,j_-^\ell)_\ell$ ($\ell \in [1,p]$) of $3p$ representations of $SU(2)$ such that
$(j_\ell,j^+_\ell,j^-_\ell)$ satisfy the triangular inequality for each $\ell$ and an additional constraint.
To describe this additional constraint, we associate canonically the $SU(2)$ generators $J_{+}^i$, $J_{-}^i$, and $J^i$
to each puncture: the Casimir of these operators $J_{\pm}^2$ and $J^2$ are fixed by the representations
$j_\pm$ and $j$ in the standard way.
The constraint reads
\nopagebreak[3]\begin{equation}\label{EPRL}
~C^i(p)= J_-^i-J_+^i-\alpha(J_+^i+J_-^i)=0.
\end{equation}
with
\nopagebreak[3]\begin{equation}\label{alfa}
\alpha\equiv\frac{ J_{+}^2-J_{-}^2}{J^2},
\end{equation}
Now $C^i$ and $D^i=J^i_++J^i_-+J^i=0$ (implicitly imposed above) cannot be simultaneously strongly imposed as they do not form a Lie algebra. One has to impose them weakly and there are two possibilities.
In order to see this let us exploit the fact that there is a strict analogy with the way the simplicity constraints are imposed in the EPRL-FK model \cite{FK, EPRL}. Observe first that equation (\ref{EPRL}) has the very same form of the linear simplicity constraints of the EPRL-FK models where the role of the Immirzi parameter is here played by $\alpha$.
The first possibility of weak imposition consists of
taking \nopagebreak[3]\begin{equation}\label{op1} j_{\pm}=(1\pm\alpha) j/2\end{equation} implying \nopagebreak[3]\begin{equation}
j=j_++j_-.
\end{equation}
It can be checked that this choice is
consistent with alpha as given in (\ref{alfa}). With this then one can check that for an admissible state $|\psi\rangle$ one has
$$C^2 |\psi\rangle=\hbar^2(1-\alpha^2)j |\psi\rangle,$$
which vanishes in the (semiclassical) limit $\hbar\to 0$, $j\to \infty$ with $\hbar j$ kept constant. Moreover, one has that
\nopagebreak[3]\begin{equation}\label{weak} \langle\phi |C^i|\psi\rangle=0\end{equation}
for arbitrary pairs of admissible states.
In other words, in this first possibility the constraint $C^i$ are satisfied strongly in the semiclassical limit, and weakly in the sense of matrix elements in general.
The second possibility is not to impose the condition $(\ref{op1})$ and leave $j_{\pm}$ completely free and only constrained by the triangular inequalities with $j$.
In that case it has been shown \cite{SF} that $(\ref{weak})$ is still satisfied. This second possibility is still compatible with the classical limit but is weaker than the previous
one.
\vskip 0.5cm
\noindent{\bf{\em Entropy calculation A}}
In this case we impose condition (\ref{op1}) and hence $j=j_++j_-$.
Such a state characterizes a
black hole of macroscopic area
$$
a \; = \; \frac{1}{2} \, \sum_{\ell=1}^p\sqrt{(d_\ell^++d_\ell^--1)(d_\ell^++d_\ell^- +1)}
$$
in unit of $\ell_p^2$.
As a consequence, the number of microstates $N(a)$ associated to a distorted black hole of area $a$
is given by the formula:
\nopagebreak[3]\begin{eqnarray}
N(a) =\sum_{p=0}^\infty \sum_{d^{\pm}_1,\cdots,d^{\pm}_p} \delta(a-\frac{\sum_{\ell=1}^p\sqrt{(d_\ell^++d_\ell^--1)(d_\ell^++d_\ell^-+1)}}{2}) \, \widetilde{N}_{k}({\bf j^+})\widetilde{N}_{k}({\bf j^-}),
\end{eqnarray}
where $d^\pm_\ell=2j_\ell^\pm+1$.
Following the steps of the previous Section, we introduce the Laplace transform $\tilde{N}(s)$ of the number of states $N(a)$.
It is given by:
\begin{eqnarray*}
\widetilde{N}(s)&=& \sum_{p=0}^\infty \sum^{k+1}_{{\bf d^+},{\bf d^-}} \, \int_{0}^\pi d\theta^+ \, \mu_k(\theta^+)
\int_{0}^\pi d\theta^- \, \mu_k(\theta^-) \nonumber\\
&\cdot&\left( \prod_{\ell=1}^p \frac{\sin(\frac{d^+_\ell \theta^+}{2})}{\sin \frac{\theta^+}{2}}
\frac{\sin(\frac{d^-_\ell \theta^-}{2})}{\sin \frac{\theta^-}{2}} e^{-\frac{s}{2} \sqrt{(d_\ell^++d_\ell^- -1)(d_\ell^++d_\ell^- +1)}}\right),
\end{eqnarray*}
where the sums run over the families
${\bf d}^\pm\!=\!(d_1^\pm,\cdots,d_p^\pm)$
of representations dimensions.
Following the same strategy as in the spherically symmetric case, the previous expression can be simplified as follows:
\begin{eqnarray*}
\widetilde{N}(s) &=& \sum_{p=0}^\infty \int_0^\pi d\theta^+ \,\mu_k(\theta^+) \int_{0}^\pi d\theta^- \, \mu_k(\theta^-)\nonumber\\
&\cdot&\left( \sum^{k+1}_{d^+,d^-=1} \frac{\sin (\frac{d^+\theta^+}{2})}{\sin \frac{\theta^+}{2}}
\frac{\sin (\frac{d^- \theta^-}{2})}{\sin \frac{\theta^-}{2}} \, e^{-\frac{s}{2} \sqrt{(d_\ell^++d_\ell^- -1)(d_\ell^++d_\ell^- +1)}}\right)^p\\
&=& \int_0^\pi d\theta^+ \int_{0}^\pi d\theta^- \; \frac{\mu_k(\theta^+)\mu_k(\theta^-)}{D^A_k(\theta^+,\theta^-;s)},
\end{eqnarray*}
with
\begin{eqnarray}
D^A_k(\theta^+,\theta^-;s) \; = \; 1 -\sum_{d^\pm=0}^k \, \frac{\sin (\frac{(d^++1)\theta^+}{2})}{\sin \frac{\theta^+}{2}}
\frac{\sin (\frac{(d^-+1) \theta^-}{2})}{\sin \frac{\theta^-}{2}} \,
e^{-\frac{s}{2} \sqrt{(d_\ell^++d_\ell^- +1)(d_\ell^++d_\ell^- +3)}} \,.\label{DkA}
\end{eqnarray}
In the definition of $D^A_k$, the sums run over $d^\pm \in [0,k]$, which are related to the spin variables $j^\pm$
by the relation $d^\pm=2j^\pm$ due to the changing of variables.
As in the spherically symmetric case, we conclude that the critical value of $s$ is the highest value for which $D^A_k$, viewed as
a function of the angles $\theta^\pm$, admits a zero. It is reached when $D^A_k$ admits one zero at $\theta^\pm=0$.
Therefore, the critical exponent $\alpha$ is the unique solution of the equation
\nopagebreak[3]\begin{equation}\label{k-sA}
1 - \sum^{k}_{d^\pm=0} (d^++1)(d^-+1) e^{-\frac{\alpha}{2} \sqrt{(d_\ell^++d_\ell^- +1)(d_\ell^++d_\ell^- +3)}} \; = \;0 \, .
\end{equation}
The exponent $\alpha$ depends on $k$ as in the spherically symmetric case and the numerical solution has been plotted
in Figure \ref{Fig:s-kDistorted1A}.
\begin{figure}[ht]
\centering
\includegraphics[scale=1]{s-kDistorted1A.eps}
\caption{In the figure, we have plotted the values of the exponent $\alpha $ as function of $k\in \mathbb N$ for the first integers. The plot shows as, similarly to the spherically symmetric black hole counting, in the case of a weak imposition of the constraint $C^i=0$ through the relation $j=j_++j_-$, an asymptotic value for $\alpha$ is quickly reached as $k\geq 4$.}
\label{Fig:s-kDistorted1A}
\end{figure}
{\bf
\noindent{\em Entropy calculation B} }
Let us now concentrate on the case of the weaker imposition of the constraint $C^i=0$, where all pairs of admissible states are taken into account. In this case, the black hole of macroscopic area is
$$
a \; = \; \frac{1}{2} \, \sum_{\ell=1}^p\sqrt{(d_\ell-1)(d_\ell +1)}
$$
in unit of $\ell_p^2$.
As a consequence, the number of microstates $N(a)$ associated to a distorted black hole of area $a$
is given by the formula:
\nopagebreak[3]\begin{eqnarray}
N(a) & = &\sum_{p=0}^\infty \sum_{d_1,\cdots,d_p} \delta(a-\frac{\sum_{\ell=1}^p\sqrt{(d_\ell-1)(d_\ell +1)}}{2}) \nonumber \\
&\cdot& \sum^{k+1}_{d^\pm_1,\cdots,d^\pm_p=1} \left(\prod_{\ell=1}^pY(j_\ell,j_\ell^+,j_\ell^-)\right) \, \widetilde{N}_{k}({\bf j^+})\widetilde{N}_{k}({\bf j^-}),
\end{eqnarray}
where $d^\pm_\ell=2j_\ell^\pm+1$ and we recall that, in order to implement the admissibility condition, $Y(j_\ell,j_\ell^+,j^-_\ell)=1$ if $(j_\ell,j^+_\ell,j^-_\ell)$ satisfy
the triangular inequality, it vanishes otherwise.
Following the steps of the previous case A, we introduce the Laplace transform $\tilde{N}(s)$ of the number of states $N(a)$.
It is given by:
\begin{eqnarray*}
\widetilde{N}(s) & = & \sum_{p=0}^\infty \sum_{\bf d} \sum^{k+1}_{{\bf d^+},{\bf d^-}} \, \int_{0}^\pi d\theta^+ \, \mu_k(\theta^+)
\int_{0}^\pi d\theta^- \, \mu_k(\theta^-) \nonumber\\
&\cdot&\left( \prod_{\ell=1}^p Y(j_\ell,j_\ell^+,j^-_\ell) \, \frac{\sin(\frac{d^+_\ell \theta^+}{2})}{\sin \frac{\theta^+}{2}}
\frac{\sin(\frac{d^-_\ell \theta^-}{2})}{\sin \frac{\theta^-}{2}} e^{-\frac{s}{2} \sqrt{(d_\ell -1)(d_\ell +1)}}\right)\nonumber\,.
\end{eqnarray*}
Following the same strategy as in the case A, the previous expression can be simplified as follows:
\begin{eqnarray*}
\widetilde{N}(s) &=& \sum_{p=0}^\infty \int_0^\pi d\theta^+ \,\mu_k(\theta^+) \int_{0}^\pi d\theta^- \, \mu_k(\theta^-)\nonumber\\
&\cdot& \left( \sum_d \sum^{k+1}_{d^+,d^-=1} Y(j_\ell,j_\ell^+,j^-_\ell) \frac{\sin (\frac{d^+\theta^+}{2})}{\sin \frac{\theta^+}{2}}
\frac{\sin (\frac{d^- \theta^-}{2})}{\sin \frac{\theta^-}{2}} \, e^{-\frac{s}{2} \sqrt{(d-1)(d+1)}}\right)^p\\
&=& \int_0^\pi d\theta^+ \int_{0}^\pi d\theta^- \; \frac{\mu_k(\theta^+)\mu_k(\theta^-)}{D^B_k(\theta^+,\theta^-;s)},
\end{eqnarray*}
with now
\begin{eqnarray}
D^B_k(\theta^+,\theta^-;s) \; = \; 1 -\sum_{d} \sum_{d^\pm=0}^k Y(j_\ell,j_\ell^+,j^-_\ell) \, \frac{\sin (\frac{(d^++1)\theta^+}{2})}{\sin \frac{\theta^+}{2}}
\frac{\sin (\frac{(d^-+1) \theta^-}{2})}{\sin \frac{\theta^-}{2}} \, e^{-\frac{s}{2} \sqrt{d(d+2)}} \,.\label{Dk}
\end{eqnarray}
Again, the spins variables $j^\pm$
are related to the sums variables by $d^\pm=2j^\pm$ due to the changing of variables. The variable $d$ is also related to the spin variable
$j$ by $d=2j$.
Similarly to the previous cases, we conclude that the critical value of $s$ is the highest value for which $D^B_k$, viewed as
a function of the angles $\theta^\pm$, admits a zero. It is reached when $D^B_k$ admits one zero at $\theta^\pm=0$.
Therefore, the critical exponent $\alpha$ is the unique solution of the equation
\nopagebreak[3]\begin{equation}\label{k-s}
1 - \sum_d \sum^{k}_{d^\pm=0} Y(j_\ell,j_\ell^+,j^-_\ell) \, (d^++1)(d^-+1) e^{-\frac{\alpha}{2} \sqrt{d(d+2)}} \; = \;0 \, .
\end{equation}
The exponent $\alpha$ again depends on $k$ and the numerical solutions has been plotted
in Figure \ref{Fig:s-kDistorted1B}.
\begin{figure}[ht]
\centering
\includegraphics[scale=1]{s-kDistorted1B.eps}
\caption{In the plot, the circles indicate the values of the exponent $\alpha$ as function of $k\in \mathbb N$ for the first integers; the squares represent values of the function $c\log{k}$ for the same values of $k$, where $c=2\sqrt{3}+O(1/k)$.}
\label{Fig:s-kDistorted1B}
\end{figure}
To have a more physical intuition of the behavior of $\alpha$ as a function of $k$, let us assume that all the spins $j$
are fixed to $1/2$: this means that the edges of the spin-network in the bulk which intersect the black hole surface
are colored by $1/2$-spins. This assumption will give us the ``shape'' of the function $\alpha(k)$ for large values of $k$ since
the main contributions to the entropy come from small values of the bulk spin $j_\ell$. Let us call $\alpha_{1/2}$ the value
of $\alpha$ where only $j=1/2$ spins contribute and $\alpha_{1/2}$ satisfies:
\nopagebreak[3]\begin{equation}
\sum_{d^+=0}^k \sum_{d^-=d^+-1}^{d^++1} (d^++1)(d^-+1) e^{-\alpha_{1/2}\frac{\sqrt{3}}{2}} \;= \; 1 \,.
\end{equation}
A straightforward calculation leads to the following expression relating $\alpha_{1/2}$ and $k$:
\nopagebreak[3]\begin{equation}
e^{\alpha_{1/2}\frac{\sqrt{3}}{2}} \; = \; 3 \sum_{n=1}^{k+1} n^2 \; = \; \frac{(k+1)(k+2)(2k+3)}{2} \,.
\end{equation}
As a consequence, in the limit where $k$ is large
$$\alpha_{1/2} \sim 2\sqrt{3} \, \log k$$
which means that $\alpha$ grows logarithmically with $k$.
Further evidence of this behavior of $\alpha$ is given by the numerical solution of eq. (\ref{k-s}) plotted above (Figure \ref{Fig:s-kDistorted1B})
where all spins are taken into account and not only $1/2$ spins. Note that the numerical value $c$ in the asymptotic formula
$\alpha\sim c\log k$ is such that $c=2\sqrt{3}+O(1/k)$ as expected.
\medskip
Let us now concentrate on the sub-leading corrections. We proceed exactly as in the spherically symmetric case:
we first evaluate $\tilde{N}(s)$ at the vicinity of the critical point $\alpha$, i.e. $s=\alpha + \varepsilon$ for
a small $\varepsilon$; we can then study the singularities of the generalized Laplace transform $\widetilde{N}(\alpha,t)$ evaluated at the critical value through the relation (\ref{generalizedLaplace}), by expanding the integrand $f_\varepsilon(\theta^+,\theta^-)$ around $\theta^+=0=\theta^-$; finally, we look at the maximal value
of $t_c$ for which $\tilde{N}(\alpha,t)$ is well-defined. The critical exponent $\beta$ is then given by $\beta=-t_c-1$. More precisely, in the distorted case we have
\begin{eqnarray*}
&&\frac{\partial^{t} f_\varepsilon(\theta^+,\theta^-) \vert_{\varepsilon=0}}{\partial \varepsilon^t}
\sim \frac{\theta^{+2}\theta^{-2}}{(\theta^{+2(t+1)}+\theta^{-2(t+1)})} =\frac{\rho^4}{\rho^{2(t+1)}}\frac{\sin^2({\varphi})\cos^2(\varphi)}{(\sin^{2(t+1)}(\varphi)+\cos^{2(t+1)}(\varphi))},
\end{eqnarray*}
where in the last equality we have changed to polar coordinates $\theta^+=\rho \sin(\varphi)$, $\theta^-=\rho \cos(\varphi)$.
From the previous equation we get that $$\tilde{N}(\alpha, -t)\simeq \int d\rho \,\rho^{-2t+3}.$$ Consequently, $\tilde{N}(\alpha, t)$ is singular when $t > 2$, in this case the critical value of $t$ is $t_c=2$.
Therefore, the critical exponent now is $\beta=-t_c-1=-3$ and the asymptotic expansion reads:
\nopagebreak[3]\begin{equation}
N(a) \; \sim \; e^{\alpha a} \, a^{-3} \;,
\end{equation}
where $\alpha$ is given in Figure \ref{Fig:s-kDistorted1A} or \ref{Fig:s-kDistorted1B} according to the prescription that defines the allowed states.
From the previous expression for the Laplace transform, we see that, in the distorted case, the constant factor in front of the logarithmic corrections becomes $3$.
\subsection*{Conclusion}
This paper has been devoted to the calculation of leading and sub-leading terms of the $SU(2)$
black hole entropy in Loop Quantum Gravity when the black hole is spherically symmetric \cite{su21,su22}
and when it is distorted \cite{su23}. To reach this aim, we derived first, by means of the recoupling theory of the quantum group $U_q(su(2))$, a new integral formula, resulting to be very useful, for the dimension of the
Hilbert space of $SU(2)$ Chern-Simons theory on a punctured two-sphere, which enters the definition of the Hilbert space of the $SU(2)$ spherically symmetric and distorted black hole as derived in \cite{su22,su23}.
Successively, we revised the technique of the Laplace transform method, exposed in detail in \cite{counting3} and firstly introduced in \cite{counting1}, to study the asymptotic behavior (in the large area limit) of the entropy associated to these two statistic mechanical ensembles.
The entropy of a $SU(2)$ spherically symmetric black hole has been already studied in \cite{counting3} when the
level $k$ is infinite. Here, following a paradigm-shift introduced in \cite{su23}, we considered the case where the level
$k$ of the Chern-Simons theory and the macroscopic area of the black hole $a$ are independent variables and we studied the effect of a
finite $k$. We showed that, if one takes into account the
finiteness of the level, the entropy of type I isolated horizons is not modified at least up to the subleading corrections, therefore recovering the results of \cite{counting3}. Moreover, the critical exponent of the leading order $\alpha$, which is now a function of the level $k$, reaches fast an asymptotic value for large $k$, as shown in the plot in Figure 6.
Concerning the entropy of a distorted $SU(2)$ black hole, this is something which has never been studied before. In this case, for each puncture coming from the bulk, there are two punctures associated to it on the horizon and the Hilbert space becomes now the direct product of two $SU(2)$ Chern-Simons Hilbert spaces with same level $k$ \cite{su23}. The $SU(2)$ symmetry is implemented by the insertion of an intertwiner between the three punctures (one from the bulk and two from the horizon), therefore, the area constraint still plays an important role. Using the techniques developed for the spherically symmetric case, we have performed the counting of the enlarged Hilbert space number of states and shown that the entropy is again proportional to the horizon area to the leading order. In the distorted case, one can distinguish two different models according to the way second class constraints are imposed weakly. In the strongest imposition of the constraints the results do not differ in a qualitative sense from those obtained in the spherically symmetric case. However, if the second class constraints are only imposed weakly, in the Gupta-Bleurer sense, then the critical exponent $\alpha$ does not go to an asymptotic value for increasing values of the level $k$ but grows logarithmically with it, as shown in Figure \ref{Fig:s-kDistorted1B}. In that sense, our model is
consistent with ``any'' value of the Immirzi parameter.
\subsubsection*{Acknowledgments}
This project was partially supported by the ANR. D.P. was supported by the {\em Marie Curie} EU-NCG network.
|
\section{Introduction}
The properties of dilute polymer solutions under shear flow have
been studied
intensively~\cite{smith:99,schr:05,teix:05,gera:06,doyl:00,cela:05,cher:05,puli:05,schr:05_1,delg:06,wink:06_1,wink:04,ripo:06,aust:99,wang:90}.
Recent advances in experimental single-molecule techniques even
provide insight into the dynamics of individual polymers under
shear
flow~\cite{smith:99,schr:05,teix:05,schr:05_1,gera:06,doyl:00}.
Similarly, the dynamics of individual polymers in a melt has been
addressed
extensively~\cite{mcle:02,kroe:04,rubi:03,kim:09,jose:07,jose:08}.
However, we are far from a similar understanding of the dynamics
of semidilute polymer solutions, although insight into the
behavior of such systems is of fundamental importance in a wide
spectrum of systems ranging from biological cells, where transport
appears in dense environments \cite{elli:03}, to turbulent drag
reduction in fluid flow. Moreover, in semidilute solutions of long
polymers, viscoelastic effects play an important role. Due to the
long structural relaxation time, the internal degrees of freedom
of a polymer cannot relax sufficiently fast under non-equilibrium
conditions and an elastic restoring force tries to push the system
towards its original state. Here, a deeper understanding can be
achieved by mesoscale hydrodynamic simulations
\cite{kapr:08,gomp:09}.
The dynamical behavior of dilute and semidilute polymer solutions
is strongly affected or even dominated by hydrodynamic
interactions~\cite{doi:86,kapr:08,gomp:09}. From a theoretical
point of view, scaling relations predicted by the Zimm model at
infinite dilution, e.g., for the dependence of dynamical
quantities as viscosity and relaxation time on the polymer length,
are, in general, accepted and
confirmed~\cite{doi:86,ahlr:99,muss:05}. However, as the
concentration of the polymer is increased beyond the segmental
overlap concentration $c^*$, where the volume occupied by polymer
coils is equal to the total volume, the dynamics becomes more
complex due to intermolecular excluded volume
interactions~\cite{dege:79,doi:86,rasp:95,pate:92}. For this
regime, the scaling theory established by de Gennes describes the
polymer dynamics using the concept of 'blobs'~\cite{dege:79}. Here
a blob consists of $g$ monomers and has the radius $\xi$. A
polymer chain comprised of $N_m$ monomers can be regarded as
composed of $N_m/g$ blobs which are hydrodynamically independent.
Inside of a single blob, the dynamics follows the predictions of
the Zimm model in dilute solution. On length scales larger than
$\xi$, hydrodynamic and excluded volume interactions are screened
due to chain overlap. Thus, the polymer dynamics on this scale can
be described by the Rouse model. When the concentration is further
increased, the polymer dynamics is dominated by entanglement
effects, which arise from physical uncrossability of chain
segments for sufficiently long polymers. Based on this theory, the
relaxation time and also the zero-shear viscosity in the
semidilute regime can be scaled by using the concentration ratio
$c/c^*$, where $c$ is the segment concentration. Various
experiments confirm the predicted dependencies for the relaxation
time and viscosity~\cite{pate:92,taka:85,rasp:95,adam:83,heo:05}.
However, a systematic simulation study is still lacking, even
though single-chain hydrodynamic simulations are well
established~\cite{pier:91,dunw:91,pier:92,ahlr:99,lyul:99,pete:99,jend:02,hsie:04,muss:05,ryde:06,send:07,zhan:09}.
The difficulty is that in the semidilute regime a large polymer
overlap is necessary, whereas at the same time the segmental
density has to be rather low to retain hydrodynamic interactions,
which requires the simulation of long polymers \cite{ahlr:01}.
The large length- and time-scale gap between the solvent and
macromolecular degrees of freedom requires a mesoscale simulation
approach in order to assess their structural, dynamical, and
rheological properties. Here, we apply a hybrid simulation
approach, combining molecular dynamics simulations (MD) for the
polymers with the multiparticle collision dynamics (MPC) method
describing the
solvent~\cite{male:99,male:00,male:00_1,kapr:08,ripo:04,ripo:05,muss:05,gomp:09,padd:06}.
As has been shown, the MPC method is very well suited to study the
non-equilibrium properties of polymers
\cite{wink:04,ryde:06,cann:08,chel:10}, colloids
\cite{ripo:06,padd:04,wyso:09}, and other soft-matter object such
as vesicles \cite{nogu:04} and cells \cite{nogu:05,mcwh:09} in
flow fields.
Experiments~\cite{smith:99,ledu:99,schr:05,teix:05,schr:05_1,gera:06,doyl:00},
theoretical studies~\cite{wang:90,wink:06_1}, and
simulations~\cite{liu:89,pier:95,pete:99,aust:99,pier:00,hsie:04,stol:05,ryde:06}
of individual polymers under shear-flow conditions exhibit large
deformations and a strong alignment of the polymers. Moreover, a
large overlap is present in a semidilute solution of long
polymers. A typical simulation requires $10^5$ - $10^6$ monomers
and $10^7$ - $10^8$ fluid particles. Hence, despite the adopted
mesoscale approach, large systems can only be studied on a
massively parallel computer architecture. Here, we present results
of large-scale simulations of semidilute polymer solutions under
shear. The simulations were performed with our program MP$^2$C
(massively parallel multiparticle collision
dynamics)~\cite{sutm:10}, which exhibits excellent scaling
behavior on the massively parallel architecture of the IBM Blue
Gene/P computer \cite{sutm:11}. For the MPC fluid, we find a
linear increase of the speedup with increasing number of cores in
a strong scaling benchmark up to $2^{12}, \ 2^{14}, \ 2^{16}$
cores for $\ 10^7, \ 8 \times 10^7, 6 \times 10^8$ fluid
particles, respectively.
The paper is organized as follows. In Sec. II, the model and
simulation approache are described. The equilibrium properties of
the system are presented in Sec. III. Sec. IV is devoted to the
structural and conformational properties of the system under
stationary shear flow. In Sec. V, results are presented for the
rheological properties and finally, Sec. VI summarizes or
findings.
\section{Model and Parameters}
The solution consists of $N_p$ linear flexible polymer chains embedded in an
explicit solvent. A linear polymer is composed of $N_m$ beads of
mass $M$ each, which are connected by harmonic springs. The bond
potential is
\begin{equation}
U_b=\frac{\kappa}{2}\sum^{N_m-1}_{i=1}
\left( \vert \vec{r}_{i+1}-\vec{r}_i\vert - l\right)^2,
\end{equation}
where $l$ is the bond length and $\kappa$ the spring constant.
Inter- and intramolecular excluded-volume interactions are taken
into account by the repulsive, shifted, and truncated
Lennard-Jones potential
\begin{equation}
U_{LJ}= 4\epsilon\left[\left(\frac{\sigma}{r}\right)^{12}-\left(\frac{\sigma}{r}\right)^{6}+\frac{1}{4} \right]\Theta\left(2^{1/6}\sigma-r\right),
\end{equation}
where $\Theta(x)$ is the Heaviside function [$\Theta(x) =0$ for $x
<0$ and $\Theta(x) = 1$ for $x\geqslant0$]. The dynamics of the
chain monomers is determined by Newton's equations of motion, which are
integrated by the velocity Verlet algorithm with the time step
$h_p$ \cite{alle:87}.
The solvent is simulated by the multiparticle collision (MPC)
dynamics method \cite{male:99,male:00,kapr:08,gomp:09}. It is
composed of $N_s$ point-like particles of mass $m$, which interact
with each other by a stochastic process. The algorithm consists
of alternating streaming and collision steps. In the streaming
step, the particles move ballistically and their positions are
updated according to
\begin{equation}
\vec{r}_i(t+h)=\vec{r}_i(t)+h\vec{v}_i(t) ,
\end{equation}
where $i =1, \ldots, N_s$ and $h$ is the time interval between
collisions. In the collision steps, the particles are sorted into
cubic cells of side length $a$ and their relative velocities, with
respect to the center-of-mass velocity of the cell, are rotated
around a randomly oriented axis by a fixed angle $\alpha$, i.e.,
\begin{equation}
\vec{v}_i(t+h)=\vec{v}_i(t)+(\mathbf{R}(\alpha)-\mathbf{E})(\vec{v}_i(t)-\vec{v}_{cm}(t)),
\end{equation}
where $\vec{v}_i(t)$ denotes the velocity of particle $i$ at time
$t$, $\mathbf{R}(\alpha)$ is the rotation matrix, $\mathbf{E}$ is
the unit matrix, and
\begin{equation}
\vec{v}_{cm} = \frac{1}{N_c}\sum^{N_c}_{j=1}\vec{v}_j,
\end{equation}
is the center-of-mass velocity of the particles contained in the
cell of particle $i$. $N_c$ is the total number of solvent
particles in that cell.
The solvent-polymer coupling is achieved by taking the monomers
into account in the collision step, i.e., for collision cells containing monomers, the center-of-mass velocity reads
\begin{equation}
\vec{v}_{cm}(t)=\frac{\sum^{N_c}_{i=1}m\vec{v}_i(t)+\sum^{N_c^m}_jM\vec{v}_j(t)}{mN_c+MN_c^m},
\end{equation}
where $N^m_c$ is the number of monomers within the considered
collision cell. To insure Galilean invariance, a random shift is
performed at every collision step~\cite{ihle:01}. The collision
step is a stochastic process, where mass, momentum and energy are
conserved, which leads to the build up of correlations between the
particles and gives rise to hydrodynamic interactions.
To impose a shear flow, for the short chains, we apply
Lees-Edwards boundary conditions \cite{alle:87}. A local
Maxwellian thermostat is used to maintain the temperature of the
the fluid at the desired value~\cite{huan:10}.
A parallel MPC algorithm is exploited for systems of long chains,
which is based on a three-dimensional domain-decomposition
approach, where particles are sorted onto processors according to
their spatial coordinates \cite{sutm:10}. Here, shear flow is
imposed by the opposite movement of two confining walls. The walls
are parallel to the $xy$-plane and periodic boundary conditions
are applied in the $x$- and $y$- directions. The equations of
motion of the solvent particles are modified by the wall
interaction~\cite{wink:09}. We impose no-slip boundary conditions
by the bounce-back rule, i.e., the velocity of a fluid particle is
reverted when it hits a wall and phantom particles in a wall are
taken into account. The same rule is applied for monomers when
colliding with a wall~\cite{gomp:09}.
The simulation parameters are listed in Table \ref{table1}. All
simulation are performed with the rotation angle
$\alpha=130^{\circ}$. Length and time are scaled according to
$\tilde{r}_{\beta}=r_{\beta}/a$, $\beta \in \{x,y,z\}$, and
$\tilde{t}=t\sqrt{k_BT/(ma^2)}$, which corresponds to the choice
$k_BT=1$, $m=1$, and $a=1$, where $T$ is the temperature and $k_B$
the Boltzmann constant. The collision time is $\tilde h=0.1$. The
parameters yield the shear viscosity $\tilde \eta = \eta/
\sqrt{mk_BT/a^4} = 8.7$. A large rotation angle $\alpha \gtrsim
90^{\circ}$ and a small time step $h$ are advantages to obtain
high fluid viscosities, low Reynolds numbers, and larger Schmidt
numbers. The selected values yield the fluid Schmidt number $Sc
\approx 14$, i.e., a fluid is simulated rather than a gas
\cite{ripo:05,gomp:09}. Between MPC collisions, the monomer
dynamics is determined by molecular dynamics simulations for
$h/h_p$ steps, with $\tilde{h}_p=0.002$. Moreover, we choose
$l=a$, $\sigma=a$, $k_BT/ \epsilon =1$, and $\tilde{\kappa}=\kappa
a^2 /(k_BT)=5\times10^3$. The large spring constant
$\tilde{\kappa}$ ensures that the mean of the bond length changes
by less than $0.5\%$ and the variance of the bond length
distribution by $3\%$ only, even at the largest shear rate.
\begin{table*}[t]
\caption{\label{table1} List of simulation parameters. $L_x$,
$L_y$, $L_z$ denote the dimensions of the simulation box, $\dot
\gamma$ the shear rate, and $\left\langle N_c \right\rangle$ is
the mean number of fluid particles in a collision cell. For $N_p$,
$c$, $c^*$, and $\dot \gamma$ the smallest and largest values used
are given. Actual concentrations are provided in figure captions.}
\begin{ruledtabular}
\begin{tabular}{ccccccccc}
$N_m$ & $N_p$ & $L_x/a$$\times$ $L_y/a$$\times$ $L_z/a$ &$\langle R_{g0}^2\rangle/l^2$& $c/l^{-3}$ & $c^*/l^{-3}$&$c/c^*$ & $\dot{\gamma}/\sqrt{k_BT/(ma^2)}$ &$\langle N_c\rangle$ \\
\hline
20& 10 -- 200& 20$\times$20$\times$20 &7.05& 0.025 -- 0.5 &0.26&0.098 -- 1.96 & $7.5\times10^{-4}$ &5\\
40& 10 -- 100 & 20$\times$20$\times$20 &17.29& 0.05 -- 0.5 &0.13&0.38 -- 3.76 & $7.5\times 10^{-4}$ &5 \\
50& 10 -- 512 & 50$\times$50$\times$50 &24.51& 0.004 -- 0.205&0.098&0.041 -- 2.08 & $10^{-4} - 3 \times 10^{-1}$ &10 \\
250&50 -- 3000 & 450$\times$75$\times$75 &163.49& 0.0049 -- 0.296&0.029&0.17 -- 10.38 & $10^{-6} - 3\times 10^{-2}$&10\\
\end{tabular}
\end{ruledtabular}
\end{table*}
\section{Equilibrium properties}
\label{sec_equil}
Before we will address polymer solutions under shear, the scaling
behavior of equilibrium properties is discussed, in order to
determine the crossover at which a solution starts to follow the
expected scaling laws of a semidilute solution.
\subsection{Conformational properties}
The mean square radius of gyration $\langle R^2_{g0} \rangle$ in
dilute solution obeys the scaling relation
\begin{equation} \label{gyrat}
\langle R^2_{g0} \rangle \propto N_m^{2\nu} ,
\end{equation}
with an exponent $\nu\approx$ 0.59 for a good solvent
\cite{dege:79,doi:86}. The obtained values for $\langle
R^2_{g0}\rangle$ are listed in Table~\ref{table1} for various
chain lengths. A fit of Eq.~(\ref{gyrat}) to our simulation data
obtained at the lowest concentrations (cf. Table 1) yields the
exponent $\nu$=0.61, in good agreement with theory and
experimental data~\cite{dege:79,doi:86}.
As the concentration increase, the polymer coils start to overlap
when the monomer concentration $c=N_m N_p/V$ exceeds the value
$c^* = N_m/V_p$, with the volume of a polymer $V_p = 4 \pi \langle
R_{g0}^2\rangle^{3/2}/3$ and the totally available volume $V$
\cite{doi:86}. Scaling considerations predict the dependence
\begin{equation} \label{eq:Rg_c}
\langle R_g^2 \rangle =\langle R_{g0}^2 \rangle \left(\frac{c}{c^*}\right)^{(2\nu-1)/(1-3\nu)}
\end{equation}
for the radius of gyration at concentrations $c \gg c^*$
\cite{dege:79,doi:86}. This relation has been confirmed
experimentally~\cite{daou:75} and by computer simulations
\cite{stol:05,peli:08}.
\begin{figure}[t]
\begin{center}
\includegraphics*[width=.4\textwidth,angle=0]{Rg_c_2.eps}
\caption{Relative mean square radii of gyration as a function of
the scaled concentration $c/c^*$ for the chain lengths $N_m=20$ (${\color{green}
\blacktriangle }$), $40$ ($\blacklozenge$), $50$ (${\color{blue} \bullet }$),
and $250$ (${\color{red} \blacksquare }$).
The solid line indicates the dependence of Eq.~(\ref{eq:Rg_c}) for $\nu=0.61$. }
\label{Rg_Rg0}
\end{center}
\end{figure}
Figure \ref{Rg_Rg0} shows radii of gyration for various polymer
lengths and concentrations. Our simulation results follow the
scaling predictions. For $c\ll c^*$, $\langle R_g^2\rangle$ is
independent of polymer concentration. At $c/c^* \approx 1$, the
coil size starts to decrease and for $c \gg c^*$, $\langle
R_g^2\rangle \sim c^{-0.265}$ with $\nu = 0.61$.
As suggested by de Gennes, the coil overlap implies a screening of
excluded volume and hydrodynamic interactions on length scales
larger than the blob size $\xi$ \cite{dege:79,doi:86}. Below this
length, the swollen conformations and hydrodynamic interactions
are maintained. The correlation length $\xi$ is independent of
chain length $N_m$ and is only a function of monomer concentration
at strong overlap, which yields the scaling relation
\begin{equation} \label{eq:xi_c}
\xi=\langle R_{g0}^2 \rangle^{1/2} \left(\frac{c}{c^*}\right)^{-\nu/(3\nu-1)}
.
\end{equation}
The crossover is reflected in the polymer structure factor
\begin{equation}
S(\vec{q}) = \frac{1}{N_m}\sum^{N_m}_{i,j=1}\langle
\exp[-i \vec{q} (\vec {r}_i-\vec{r}_j)]\rangle ,
\end{equation}
which exhibits the power-law dependence $S(\vec{q}) \sim
q^{1/\delta}$ for $1 \ll q \langle R_g^2 \rangle^{1/2} \ll \langle
R_g^2 \rangle^{1/2}/ l$ with $\delta=1/2$ for a melt and
$\delta=\nu$ in good solvent \cite{stro:07}. Hence, in a
semidilute solution two regimes are expected, separated by the
correlation length $\xi$: A good solvent behavior on length scales
smaller than $\xi$ and a Gaussian chain behavior on length scales
larger than $\xi$, i.e.,
\begin{equation}
S(q) \sim \left\{ \begin{array}{ccc}
q^{-2} & \mbox{for}
& 2\pi/\langle R_g^2 \rangle^{1/2} < q < 2\pi/ \xi, \\
q^{-1/\nu} & \mbox{for} & 2\pi/ \xi < q < 2\pi / l,
\end{array}\right. .
\end{equation}
\begin{figure}[t]
\includegraphics*[width=.4\textwidth,angle=0]{structure_factor_L50_L250.eps
\caption{\label{fig1} Structure factors of polymers of length
$N_m=250$ for the concentration ratios $c/c^*= 0.17$ ({\color{green}-----}),
$c/c^*= 2.77$
({\color{blue}- - -}), $c/c^*= 5.19$ ({\color{red}- $\cdot$ -}),
and $c/c^*= 10.38$ (-----), and
of $N_m=50$ in the dilute regime (- $\cdot$$\cdot$ -).
The dashed and solid straight lines represent power-law functions with the
exponents $1/\nu = 1/0.61$ and $2$, respectively. Inset: Dependence of the
blob size on the concentration. }
\end{figure}
In Fig.~\ref{fig1} polymer structure factors are shown for $N_m=
250$ in dilute and semidilute solutions as well as for $N_m =50$
in dilute solution. In order to obtain the two regimes separated
by $\xi$, the polymer chains have to be sufficiently long to
provide not only a ratio $c/c^*\gg 1$ but also a low segment
concentration $c$. As shown in the figure, for dilute solutions,
$S(q)$ decays with an exponent $1/\nu$, where $\nu =0.61$. The
polymers in the semidilute regime show a crossover from the
scaling behavior $S \sim q^{-2}$ at small $q$ to the behavior $S
\sim q^{-1/\nu}$ at large $q$ values. The crossover between the
two regimes corresponds to $q \approx 2\pi/\xi$. The values for
$\xi$ are presented in the inset of Fig.~\ref{fig1} and are found
to be in excellent agreement with the scaling
prediction~(\ref{eq:xi_c}) with $\nu = 0.61$. Thus, the scaling
relation captures the concentration dependence of the blob size
for the considered range very well.
\subsection{Dynamics}
\label{equ_dynamics}
The polymer dynamics is dominated by hydrodynamic interactions in
dilute solution. Theoretical results on the concentration
dependence of the relaxation times for small overlap
concentrations are presented in Refs.~\cite{muth:78,muth:84}.
Compared to the infinite-dilution limit, a term linear in the
concentration is obtained, which is consistent with experimental
data~\cite{pate:92}. However, the experimental data can also well
be fitted by an empirical exponential function~\cite{pate:92}.
In semidilute solution, the intermolecular interactions between
different chains become increasingly important. The dynamics of
the polymers can be classified according to their intermolecular
interactions as unentangled or entangled~\cite{pate:92,rasp:95}.
In the {\em unentangled regime}, the monomers move according to
Brownian motion in all three spacial directions and their dynamics
can be described by the Rouse behavior of polymers which consist
of blobs. Thus, the longest relaxation relaxation time reads as
\begin{equation}
\tau=\tau_{b} \left(\frac{N_m}{g} \right)^2,
\label{t_blob_rouse}
\end{equation}
where $\tau_{b}$ is the blob relaxation time and $g$ the number of
monomers in a blob. Inside of a blob, the dynamics follows the
Zimm behavior
\begin{equation}
\tau_{b} \sim \left( \frac{\xi}{l} \right)^3 ,
\label{t_blob_zimm}
\end{equation}
and the longest relaxation time exhibits the concentration
dependence
\begin{equation}
\tau = \tau_0 \left(\frac{c}{c^*} \right)^{(2-3\nu)/(3\nu-1)} .
\label{t_unentangled}
\end{equation}
In the \textit{entangled regime}, polymers are assumed to exhibit reptation
motion inside a tube caused by the presence of neighboring chains.
The monomer dynamics is then described by reptation
theory~\cite{dege:79}, where the longest relaxation time obeys the
relation
\begin{equation}
\tau \sim \left(\frac{c}{c^*} \right)^{(3-3\nu)/(3\nu-1)}.
\label{t_entangled}
\end{equation}
By calculating end-to-end vector correlation functions, which
exhibit an exponential decay, we determined the longest polymer
relaxation times $\tau$ for various concentrations. The relaxation
time $\tau_0$ at infinite dilution is obtained by extrapolation
to zero concentration. The obtained values for $\tau_0$ are shown
as a function of polymer length in the inset of
Fig.~\ref{Relaxation_c}. Their length dependence is well described
by the power-law $\tau_0 \sim N_m^{3 \nu}$, with $\nu=0.6$, in
accord with predictions of the Zimm model \cite{doi:86}.
Figure~\ref{Relaxation_c} depicts the dependence of the relaxation
time on concentration. In the vicinity of $c/c^* = 1$, $\tau$
follows the scaling prediction (\ref{t_unentangled}) for an
unentangled semidilute polymer solution. With increasing
concentration, $\tau$ increases faster, which we attribute to
strong intermolecular interactions. Although, there are no
entanglements in our system for $c/c^*<10$, the predicted
dependence for entangled polymer melts is indicated by the solid
line for illustration. This dependence is not reached and requires
longer polymers or higher concentrations. A very similar
dependence has been obtained experimentally in Ref. \cite{pate:92}
over comparable concentration and relaxation time ranges.
\begin{figure}[t]
\begin{center}
\includegraphics*[width=.4\textwidth,angle=0]{Relax_time_250_new5.eps}
\caption{Concentration dependence of longest polymer relaxation times $\tau$ for
the polymer lengths $N_m=50$ ($\bullet$) and $250$ (${\color{blue} \blacksquare }$).
The dashed line indicates the prediction
Eq.~(\ref{t_unentangled}) and the continues line Eq.~(\ref{t_entangled})
with $\nu=0.6$. Inset: Polymer length dependence of the relaxation time at
infinite dilution. The solid line shows the dependence
$\tau_0 \sim N_m^{3 \nu}$ with $\nu = 0.6$.
\label{Relaxation_c}}
\end{center}
\end{figure}
According to the Zimm model~\cite{zimm:56,doi:86,harn:96,muss:05},
hydrodynamic interactions strongly affect the diffusive
dynamics of polymers in solution and lead to the time dependence
\begin{equation}
g_2(t)=\langle ([{\bm r}_i(t)-{\bm r}_{cm}(t)]-[{\bm r}_i(0)-{\bm r}_{cm}(0)])^2 \rangle \sim t^{2/3}
\label{g2_dilute}
\end{equation}
of their monomer mean squared displacement in the center-of-mass
reference frame for time scales larger than the Brownian time
\cite{dhon:96} and smaller than the longest relaxation time at
which $g_2(t)$ saturates. In the semidilute regime, hydrodynamic
interactions are screened for time scales larger than $\tau_b$,
the time needed to diffuse a blob diameter \cite{ahlr:01}.
Consequently, after a time $\tau_b$, the dynamics is Rouse-like
and $g_2(t) \sim t^{1/2}$~\cite{ahlr:01}. Figure~\ref{msd_9}
displays $g_2(t)$ for various concentrations for polymers of
length $N_m=250$.
For a dilute solution with $c/c^*=0.17$, $g_2(t)\sim t^{2/3}$ in
the time interval $10^{-3} < t/\tau_0 < 10^{-1}$ as expected. For
$t > \tau$, $g_2$ slowly approaches a plateau value. At larger
concentrations, $g_2(t)$ displays a $t^{2/3}$ dependence which
turns into a $t^{1/2}$ behavior at a time $\tau_b$, which decrease
with increasing concentration. The concentration dependence of the
mean squared displacement reflects the screening of hydrodynamic
interactions in the semidilute regime. However, the dependence of
$\tau_b$ on concentration, which is linked to the screening length
$\xi_{H}$ according to $\tau_b \sim \xi^3_{H}$ and $\xi_{H} \sim
c^{-\gamma}$, where $\gamma$ is predicted to be $1$
\cite{free:74,edwa:84}, $0.6$ \cite{fred:90}, or $0.5$
\cite{dege:79,ahlr:01}, respectively, cannot be obtained
unambiguously from our simulations, because the different time
regimes are too short. Simulations of longer polymers are
necessary to arrive at clear and pronounced diffusion regimes.
\begin{figure}[t]
\begin{center}
\includegraphics*[width=.4\textwidth,angle=0]{msd9.eps}
\caption{Mean square displacements of monomers in the center-of-mass
reference frame for the concentrations $c/c^*=0.17$ ({\color{red}- - -}),
$c/c^*=2.77$ (- $\cdot$ -), $c/c^*=5.19$ ({\color{blue}-- -- --}),
$c/c^*=10.38$ (-----) of polymers of length $N_m=250$.
The short lines indicate the dependencies $g_2 \sim t^{2/3}$ and
$g_2 \sim t^{1/2}$, respectively.
\label{msd_9}}
\end{center}
\end{figure}
\section{Semidilute polymer solutions in shear flow}
We now discuss the properties of polymer solutions in shear flow.
At infinite dilution, the flow strength is characterized by the
Weissenberg number $\mathrm{Wi}=\dot{\gamma}\tau_{0}$, where
$\dot{\gamma}$ is the bare shear rate. For $\mathrm{Wi}\ll 1$, the
weak shear flow regime, the chains are able to undergo
conformational changes before the local strain has changed by a
detectable amount, while for $\mathrm{Wi}\gg 1$, the chains are
driven by the flow and they are not able to relax back to the
equilibrium conformation. This is illustrated in
figure~\ref{fig:snapshots}, which displays snapshots for various
flow rates. At small Weissenberg numbers, the polymers are only
weakly perturbed and are close to their equilibrium conformations,
whereas large $\rm Wi$ imply large deformations and a strong
alignment with flow.
As pointed out in section~\ref{equ_dynamics}, the polymer
relaxation times depend on concentration. Thus, in the following,
some properties will be characterized by the Weissenberg number
$\mathrm{Wi}_c= \mathrm{Wi} \, \tau(c)/\tau_0 = \dot \gamma
\tau(c)$. The question is, to what extent the influence of
concentration on the polymer dynamics can be accounted for by a
concentration-dependent Weissenberg number. As we will see, this
concept applies well for all structural and dynamical properties.
\begin{figure}
\includegraphics*[clip,angle=270,width=0.4\textwidth]{low.eps}\\[1ex]
\includegraphics*[clip,angle=270,width=0.4\textwidth]{high.eps}
\caption{Snapshots of systems with $N_p=800$ polymers of length $N_m=250$
for the Weissenberg
numbers $\mathrm{Wi_c}=18$ (top) and $\mathrm{Wi_c}=184$ (bottom).
For illustration, some of the chains are highlighted in red.
\label{fig:snapshots}}
\end{figure}
\begin{figure}[t]
\begin{center}
\begin{tabular}{cc}
\begin{minipage}{1.65in}
\includegraphics*[width=\textwidth]
{Gimp1_2D_400_0001_xy.eps}
\end{minipage}
&
\begin{minipage}{1.65in}
\includegraphics*[width=\textwidth]
{Gimp1_2D_400_0001_xz.eps}
\end{minipage}
\\
\begin{minipage}{1.65in}
\includegraphics*[width=\textwidth]
{Gimp1_2D_400_03_xy.eps}
\end{minipage}
&
\begin{minipage}{1.65in}
\includegraphics*[width=\textwidth]
{Gimp1_2D_400_03_xz.eps}
\end{minipage}
\\
\end{tabular}
\caption{Monomer density distributions in the flow-gradient plane
(a), (c) and flow-vorticity plane (b), (d) for the Weissenberg
numbers $\mathrm{Wi}_c = 1$ (a), (b) and $\mathrm{Wi}_c = 307$
(c), (d). The concentration is $c/c^*=1.6$ and the chain length is
$N_m=50$. \label{shape_shear}}
\end{center}
\end{figure}
\begin{figure}[h]
\begin{center}
\begin{tabular}{cc}
\begin{minipage}{1.65in}
\includegraphics*[width=\textwidth]
{Gimp1_2D_40_001_xy.eps}
\end{minipage}
&
\begin{minipage}{1.65in}
\includegraphics*[width=\textwidth]
{Gimp1_2D_512_001_xy.eps}
\end{minipage}
\\
\end{tabular}
\caption{Monomer density distributions in the flow-gradient plane
for the shear rate $\tilde {\dot{\gamma}} = 10^{-3}$ and the
concentrations $c/c^*=0.16$ (a) and $c/c^*=2.08$ (b), which corresponds to the
Weissenberg numbers $\mathrm{Wi}_c = 6.2$ and $\mathrm{Wi}_c =11$, respectively.
The chain length is $N_m=50$.
\label{shape_c}}
\end{center}
\end{figure}
\begin{figure}[t]
\begin{center}
\includegraphics*[width=.4\textwidth,angle=0]{tensor_250_new_10.eps}
\caption{Deformation ratios $\delta G_{xx}$ as function of
Weissenberg number. Open symbols correspond to
systems with $N_m=50$ for $c/c^*=0.16$ ($\circ$),
$c/c^*=1.6$ (${\color{red}\triangle}$), and $c/c^*=2.08$
(${\color{blue}\square}$). Filled symbols denote results for
$N_m=250$ with the concentrations $c/c^*=0.17$ ($\bullet$), $c/c^*=2.77$
(${\color{green}\blacktriangle}$), $c/c^*=5.19$
(${\color{blue}\blacklozenge}$), and $c/c^*=10.38$
(${\color{red}\blacksquare}$). In the inset, the
same data are shown as function of $\mathrm{Wi}$.
\label{delta_G_xx}}
\end{center}
\end{figure}
\subsection{Conformations}
The average shape of an individual chain in solution under shear
is illustrated in Fig.~\ref{shape_shear} by the density
distribution of monomer positions with respect to the polymer
center of mass. At low Weissenberg numbers
(Fig.~\ref{shape_shear}(a), (b)), the polymers are only weakly
deformed and aligned with respect to the flow direction $x$,
whereas they are considerably stretched and aligned in the flow
direction and are compressed in the gradient and vorticity
direction for high shear rates (Fig.~\ref{shape_shear}(c), (d)).
Figure~\ref{shape_c} shows that the extent of deformation and
alignment depends upon polymer concentration. At the same
Weissenberg number, a larger deformation and a more pronounced
alignment is found at higher concentrations.
\subsubsection{Radius of gyration}
Polymer deformation and orientation are characterized
quantitatively by the gyration tensor
\begin{equation}
G_{\beta \beta'}=\frac{1}{N_m}\sum^{N_m}_{i=1}\langle
\Delta r_{i,\beta} \Delta r_{i,\beta'}\rangle,
\label{equ:Gtensor}
\end{equation}
where $\Delta r_{i,\beta}$ is the position of monomer $i$ in the
center-of-mass reference frame of the polymer.
In Fig.~\ref{delta_G_xx}, the relative deformation along the flow
direction
\begin{equation}
\delta G_{xx}=\frac{G_{xx}-G_{xx}^0}{G_{xx}^0},
\label{delta_G}
\end{equation}
where $G_{xx}^0=\langle R_g^2 \rangle/3$ is the gyration tensor at
equilibrium for the particular concentration, is shown for various
concentrations and polymer lengths. A significant polymer
stretching appears for $\mathrm{Wi}_c
>1$. At large shear rates, the stretching
saturates at a maximum, which is smaller than the value
corresponding to a fully stretched chain ($G_{xx} \approx l^2
N_m^2/12$) and reflects the finite size of a polymer. This is
consistent with experiments on DNA \cite{schr:05,schr:05_1}, where
the maximum extension is on the order of half of the contour
length, and theoretical calculations \cite{wink:10}. It is caused
by the large conformational changes of polymers in shear flow,
which yields an average extension smaller than the contour length.
Nevertheless, molecules assume totally stretched conformations at
large Weissenberg numbers during their tumbling dynamics.
Interestingly, a universal dependence is obtained for $\delta
G_{xx}$ as function of a concentration-dependent Weissenberg
number $\mathrm{Wi}_c$ at a given polymer length, whereas in terms of
$\mathrm{Wi}$, polymers at larger concentrations exhibit a stronger
stretching at the same $\mathrm{Wi}$, as shown in the inset of
Fig.~\ref{delta_G_xx} \cite{stol:05}. The latter is evident, since
the longest relaxation time of a polymer at higher concentrations
is larger and hence the polymer is more strongly deformed at the
same shear rate.
Theoretical calculations for single polymers in dilute solution
predict the dependence $\delta G_{xx} = C_{x}\mathrm{Wi}^2$ for $\mathrm{Wi} <1$,
where $C_x$ is a universal constant. The renormalization group
calculations of Ref. \cite{wang:90} yield $C_{x}=0.27$, whereas a
calculation based on a Gaussian phantom chain model yields $C_{x}
\approx 0.3$ \cite{brun:93,carl:94,pier:95,wink:06_1,wink:10}. As
shown in Fig.~\ref{delta_G_xx}, the simulations confirm the
quadratic dependence on the shear rate; $\delta G_{xx}$ is
independent of chain length for $\mathrm{Wi} <1$ and $C_x \approx 0.1$.
For $\mathrm{Wi} >10$, finite size effects appear and different asymptotic
values are assumed for the two chain lengths. We like to stress
that our simulations are in agreement with the molecular dynamics
simulation results of Ref.~\cite{pier:95} and the SANS data of
Refs.~\cite{lind:88,lind:89}.
\begin{figure}[t]
\begin{center}
\includegraphics*[width=.4\textwidth,angle=0]{tensor_250_new_17.eps}
\caption{
Relative deformations in the gradient and vorticity direction (inset).
Open symbols correspond to systems with $N_m=50$ for $c/c^*=2.08$
(${\color{blue}\square}$). Filled symbols denote results for
$N_m=250$ with the concentrations $c/c^*=0.17$ ($\bullet$),
$c/c^*=1.38$ (+), $c/c^*=2.77$ (${\color{green}\blacktriangle}$),
$c/c^*=5.19$ (${\color{blue}\blacklozenge}$), and $c/c^*=10.38$
(${\color{red}\blacksquare}$). \label{delta_G_yy}}
\end{center}
\end{figure}
In the gradient and the vorticity directions, the polymers are
compressed, with a smaller compression in the vorticity direction
as shown in Fig.~\ref{delta_G_yy}. To highlight the universal
properties of the systems, we present the ratios $G_{\beta
\beta}/G_{\beta \beta}^{00}$ $(\beta \in \{y,z\})$, where
$G_{\beta \beta}^{00} = \left\langle R^2_{g0} \right\rangle/3$ is
calculated from the radius of gyration in dilute solution at
equilibrium. At low shear rates---there is no detectable shear
deformation by shear---polymers shrink by concentration effects
for $c/c^* > 1$ (cf. Fig.~\ref{Rg_Rg0}). This is illustrated in
Fig.~\ref{delta_G_yy} for $\mathrm{Wi}_c<10$ and various concentrations.
The dashed lines indicate the values of $\left\langle R^2_g
\right\rangle/\left\langle R^2_{g0} \right\rangle$ from
Fig.~\ref{Rg_Rg0}. Evidently, the ratios $G_{\beta
\beta}/G^{00}_{\beta \beta}$ are consistent with the values
$\left\langle R^2_g \right\rangle/\left\langle R^2_{g0}
\right\rangle$ for each concentration. The ratio for the shorter
chain $N_m=50$ and concentration $c/c^*=2.08$ is close to that of
the longer chain with a similar concentration ratio $c/c^*=2.77$.
This is consistent with the fact that $\left\langle R^2_g
\right\rangle/\left\langle R^2_{g0} \right\rangle$ is independent
of chain length (cf. Fig.~\ref{Rg_Rg0}). With increasing shear
rate, the various curves progressively approach a universal
function, which decays as $\sim\mathrm{Wi}^{-0.45}$ over the
considered $\mathrm{Wi}$-range. Hence, we obtain a different scaling
behavior of the radius of gyration along the flow direction and
the transverse directions. The reason is that a high monomer
density is maintained along the flow direction due to polymer
stretching, whereas the density in the transverse directions
decreases by flow-induced polymer shrinkage.
The exponents of the power-law decay of $G_{yy}$ and $G_{zz}$
compare well with the experimental data on single DNA
molecules~\cite{schr:05_1}. Similarly, simulations (with/without
hydrodynamic and excluded volume interactions) yield comparable
exponents~\cite{schr:05_1,hur:00}. However, simulations in Ref.
\cite{schr:05_1} for even larger Weissenberg numbers seem to
produce an exponent closer to the theoretically expected value of
$2/3$ \cite{wink:06_1,wink:10}. According to theory, there is a
broad crossover regime before the asymptotic behavior at large
Weissenberg numbers is assumed, and the considered $\mathrm{Wi}$ fall into
that crossover regime.
\begin{figure}[t]
\begin{center}
\includegraphics*[width=.4\textwidth,angle=0]{angle_Rg_test14_1.eps}
\caption{Dependence of $\tan(2\chi_G)$ on shear rate.
Open symbols correspond to
systems with $N_m=50$ for $c/c^*=0.16$ ($\circ$) and $c/c^*=2.08$
(${\color{blue}\square}$). Filled symbols denote results for
$N_m=250$ with the concentrations $c/c^*=0.17$ ($\bullet$),
$c/c^*=0.35$ ($\blacktriangleleft$), $c/c^*=0.69$
($\blacktriangledown$), $c/c^*=2.77$
(${\color{green}\blacktriangle}$), $c/c^*=5.19$
(${\color{blue}\blacklozenge}$), and $c/c^*=10.38$
(${\color{red}\blacksquare}$). The solid and dashed lines are
theoretical results \cite{wink:06_1,wink:10}.
In the inset, the same data are shown as function of $\mathrm{Wi}$. Lines are
guides for the eye only.}
\label{angle_G_1}
\end{center}
\end{figure}
\subsubsection{Alignment}
The alignment of the polymers is characterized by the angle
$\chi_G$, which is the angle between the eigenvector of the
gyration tensor with the largest eigenvalue and the flow
direction. It is obtained from the components of the radius of
gyration tensor via \cite{aust:99}
\begin{equation} \label{alignment}
\tan(2\chi_G)=\frac{2G_{xy}}{G_{xx}-G_{yy}}.
\end{equation}
The dependence of $\tan(2\chi_G)$ on shear rate and concentration
is shown in Fig.~\ref{angle_G_1}. Again, a universal curve is
obtained for the different concentrations at a given polymer
length. Moreover, $\tan(2\chi_G)$ seems to be independent of
polymer length for $\mathrm{Wi}_c<100$, whereas we find a length
dependence for larger Weissenberg numbers. In this high shear
rate regime, we find $\tan(2\chi_G) \sim (\mathrm{Wi}_c)^{-1/3}$.
We like to emphasize that only the shear rate can be scaled in
order to arrive at a universal function. The angle, or
$\tan(2\chi_G)$, cannot be scaled to absorb flow or polymer
properties in an effective variable. Hence, the universal behavior
of the alignment angle for various concentrations confirms that
the Weissenberg number $\mathrm{Wi}_c$ is the correct scaling variable and
that the alignment of polymers at different concentrations depends
on the combination $\mathrm{Wi}_c= \dot \gamma \tau$ of shear rate and
relaxation time only.
The analytical description of Refs.~\cite{wink:06_1,wink:10}
predicts the dependence
\begin{align} \label{align_theo}
\tan(2\chi_G)\sim \left(\frac{l_p}{L \mathrm{Wi}^*}\right)^{1/3}
\end{align}
for semiflexible polymers in dilute solution in the limit $\mathrm{Wi}^*
\to \infty$. Here, we introduce the Weissenberg number $\mathrm{Wi}^* =
\dot \gamma \tau_{\mathrm{th}}$ for the theoretical result,
because the relaxation times from theory and simulation might not
be the same; $L$ is the length and $l_p$ the persistence length of
the polymer. The analytical result describes the simulation data
well at large shear rates, when the Weissenberg number of the
theoretical model is set to $\mathrm{Wi}^*= \mathrm{Wi}_c/2$. To compare the
predicted length dependence with that of the simulation, we apply
the relation $\langle R^2_{e0} \rangle = 2 l_pL$ to obtain a
persistence length, with $\langle R^2_{e0} \rangle$ the polymer
mean square end-to-end distance in dilute solution at equilibrium,
which yields $l_p/L \approx 0.025$ for $N_m=50$ and $l_p/L \approx
0.008$ for $N_m=250$. With these values, the ratio of
$\tan(2\chi_G)$ of the polymer of length $L=50a$ and $L=250a$ is
$1.5$. This compares well with the factor $1.33$ following from
the simulation results, which suggests that excluded volume
interactions are of minor importance for intermediate flow rates.
In the limit $\mathrm{Wi}_c \to 0$, theory predicts $\tan(2\chi_G) \sim
\mathrm{Wi}_c^{-1}$. The simulation data do not show this dependence on
the considered range of Weissenberg numbers, which might be due to
excluded volume interactions.
The inset of Fig.~\ref{angle_G_1} displays a strong dependence of
$\chi_G$ on concentration. The higher the concentration, the more
the chains are orientated along the flow direction. Such a
concentration effect has also been reported in light scattering
experiments~\cite{link:93}, where dilute polymer solutions are
considered. A comparison of the experimental data with the
simulation results is presented in Fig.~\ref{fig:cmp_chi_G}. Two
data sets are presented, a dilute solution, with the concentration
$0.113g/l$, and a semidilute solution with the approximately ten
times higher concentration $1.094g/l$, both taken from Fig.~8 of
Ref. \cite{link:93}. Evidently, the experimental data fit well
with our simulation results. Moreover, both, experiments and
simulations, yield a shift of the curves for the higher
concentrations to smaller Weissenberg numbers. In Ref.
\cite{link:93}, a factor $\beta_e \sim [\eta] \dot\gamma$, where
$[\eta]$ is the intrinsic viscosity, is used to present the data.
This quantity is proportional to $\mathrm{Wi}$, since $[\eta]$ is
proportional to the longest relaxation time $\tau$; however,
$\beta_e$ and $\mathrm{Wi}$ are not identical. No adjustment parameter is
used in Fig.~\ref{fig:cmp_chi_G}, which suggests that the ratio
$\beta_e/\mathrm{Wi}$ is close to unity.
\begin{figure}[t]
\begin{center}
\includegraphics*[width=.4\textwidth,angle=0]{angle_Rg_test20.eps}
\caption{Comparison of experimental and simulation data
for the average orientation angle $\chi_G$. The concentrations are
$c=0.016/l^{-3}$ ($\bullet$), $c=0.205/l^{-3}$
(${\color{red} \blacksquare }$) for our simulations (with $N_m=50$),
and $0.113g/l$ ($\circ$), $1.094g/l$
(${\color{red}\square}$) for the experiments \cite{link:93}.}
\label{fig:cmp_chi_G}
\end{center}
\end{figure}
\section{Rheology}
\subsection{Shear viscosity}
Under shear flow, the viscosity $\eta(\dot{\gamma})$ is obtained
from the relation
\begin{equation}
\eta(\dot{\gamma})=\frac{\sigma_{xy}(\dot{\gamma})}{\dot{\gamma}},
\label{equ:vis}
\end{equation}
where $\sigma_{xy}$ is the shear stress~\cite{bird:76,bird:87}. In
our simulations, $\sigma_{xy}$ is calculated using the virial
formulation of the stress tensor \cite{,alle:87,wink:09,wink:92}.
For sufficiently weak flow, the polymer solution is in the
Newtonian regime, i.e., $\sigma_{xy} \sim \dot{\gamma}$ and the
viscosity is independent of shear rate. Thus, the viscosity
obtained in this low shear rate regime is equal to the zero-shear
viscosity denoted by $\eta_0$. The latter depends on the monomer
concentration, which is often presented in the form
\begin{equation}
\eta_0=\eta_s[1+[\eta]c+k_H([\eta]c)^2+\ldots],
\label{equ:vis_0_d}
\end{equation}
where $\eta_s$ is the solvent viscosity and $k_H$ the Huggins
constant~\cite{bird:76,taka:85}.
We determine the intrinsic viscosity by a linear extrapolation to
zero concentration of both, $(\eta_0 -\eta_s)/ c \eta_s$ and the
inherent specific viscosity $[\ln(\eta_0/\eta_s)]/c$. The common
intercept of these two functions gives $[\eta]$
\cite{taka:85,pami:08}.
The intrinsic viscosity is proportional to $R_g^3/N_m$
\cite{doi:86} and is therefore proportional to the inverse of the
overlap concentration $c^*$~\cite{weil:79, rasp:95}. We find
$[\eta] c^* \approx 0.9$ and $[\eta] c^* \approx 1$ for the
polymer of length $N_m=40$ and $N_m=50$, respectively, which means
that the proportionality coefficient is close to unity for the
considered model systems.
The Einstein relation
\begin{align} \label{einstein}
\eta = \eta_s \left( 1 + 2.5 \phi \right) ,
\end{align}
where $\phi$ is the volume fraction, captures the concentration
dependence of hard sphere suspensions for $\phi \ll 1$. This
relation should also apply for dilute polymer solutions, when the
hydrodynamic radius $R_H$ is used to define the volume fraction,
i.e., $\phi=(4\pi/3) R_H^3 N_p/V$. Equations (\ref{equ:vis_0_d})
and (\ref{einstein}) are consistent if $[\eta] c^*= 2.5
(R_H/R_g)^3$. Since $[\eta] c^* \approx 1$, as explained above,
consistency requires $R_g/R_H \approx 1.36$. From our simulations,
we find the hydrodynamic radii $R_H \approx 3.8 l$ and $R_H
\approx 9.4 l$ for the polymers of length $N_m=50$ and $N_m=250$,
respectively, which yields the ratios $1.3$ and $1.36$. These
values are very close to the value necessary to match the Einstein
relation. The ratios are somewhat smaller than the asymptotic
value $R_g/R_H \approx 1.59$ for $N_m \to \infty$ obtained in
Ref.~[82], which is a consequence of the fact that we consider
insufficiently long chains.
The term $k_H([\eta]c)^2$ depends on hydrodynamic interactions.
The value of the Huggins constant $k_H$ of flexible polymers is in
the range of $0.2-0.8$ and depends on solvent quality
\cite{taka:85,pami:08}. In good solvent, typically the value $0.3$
is found experimentally~\cite{pami:08}. Expressing
Eq.~(\ref{equ:vis_0_d}) in terms of the dimensionless parameter
$[\eta]c$ as
\begin{equation}
\eta_R = \frac{(\eta_0-\eta_s)}{\eta_s[\eta]c}=1+k_H[\eta]c+\ldots ,
\label{equ:vis_0_d2}
\end{equation}
which is denoted as relative viscosity, allows us to determine
$k_H$. The inset of Fig.~\ref{fig:cmp_vis_0} shows $\eta_R-1$ as
function of $[\eta]c$ for polymers of length $N_m=40$ and
$N_m=50$. The slope of the solid line is $k_H=0.35$, in close
agreement with experiments \cite{taka:85,pami:08}.
For semidilute unentangled polymer solutions, the viscosity is
proportional to the number of blobs per chain and can be expressed
by the scaling relation~\cite{dege:79, taka:85, rasp:95, heo:05}
\begin{equation}
\eta_0=\eta_s \left( \frac{c}{c^*} \right)^{1/(3\nu-1)} .
\label{equ:vis_0_semi}
\end{equation}
Figure~\ref{fig:cmp_vis_0} displays zero-shear viscosities as
function of concentration for various polymer lengths. For
$c/c^*\gtrsim 3$, the data are close to the power-law of
Eq.~(\ref{equ:vis_0_semi}).
\begin{figure}[t]
\begin{center}
\includegraphics*[width=.4\textwidth,angle=0]{viscosity_zero_shear_c4.eps}
\caption{Dependence of the zero shear viscosity on the scaled concentration
$c/c^*$ for the polymer lengths $N_m=40$ (${\color{blue} \blacktriangle }$),
$N_m=50$ ($\bullet$), and $N_m=250$ (${\color{red} \blacksquare }$).
The solid line indicates the power-law $(c/c^*)^{1/(3\nu-1)}$
with $\nu=0.6$. In the inset, $\eta_R-1$, Eq~(\ref{equ:vis_0_d2}), is shown
as function of $[\eta]c$ for $N_m=40$ (${\color{blue} \blacktriangle
}$) and $N_m=50$ ($\bullet$); the slope of the solid line is $0.35$, which
corresponds to the Huggins constant of polymers in good solvent.
\label{fig:cmp_vis_0}}
\end{center}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics*[width=.4\textwidth,angle=0]{shear_strong_L250_L50_new3.eps}
\caption{Dependence of the polymer contribution to shear viscosity on shear rate.
Open symbols correspond to
systems with $N_m=50$ for $c/c^*=0.16$ ($\circ$), $c/c^*=1.6$
(${\color{red}\Diamond}$), and $c/c^*=2.08$
(${\color{blue}\square}$). Filled symbols denote results for
$N_m=250$ with the concentration $c/c^*=0.17$ ($\bullet$),
$c/c^*=0.35$ ($\blacktriangleleft$), $c/c^*=1.38$
($\blacktriangleright$), $c/c^*=2.77$
(${\color{green}\blacktriangle}$), $c/c^*=5.19$
(${\color{blue}\blacklozenge}$), and $c/c^*=10.38$
(${\color{red}\blacksquare}$). } \label{vis_strong_shear}
\end{center}
\end{figure}
At sufficiently large shear rates, the polymers are aligned and
deformed, which implies shear
thinning~\cite{bird:87,aust:99,wink:06_1,wink:10}.
Figure~\ref{vis_strong_shear} shows the polymer contribution
$\eta^p$ to the shear viscosity. Similar to the alignment angle,
the viscosity is a universal function of the Weissenberg number
$\mathrm{Wi}_c$ and shows a weak dependence on polymer length. It is
independent of shear rate for $\mathrm{Wi}_c \ll 1$, decrease
approximately as $\mathrm{Wi}_c^{-0.3}$ for $1 <\mathrm{Wi}_c <10^2$, and
$\mathrm{Wi}_c^{-0.45}$ for higher shear rates. This behavior is
consistent with other simulation results
\cite{schr:05_1,liu:89,pete:99, hsie:04,galu:10}. However, an even
stronger decay of the viscosity is observed in simulations at
larger shear rates in Refs. \cite{schr:05_1,aust:99}. Experiments
of dilute polymer solutions reported exponents ranging from $-0.4$
to $-0.85$~\cite{bird:87,schr:05_1}. Theoretical calculations for
dumbbells and finite extendable polymers predict the dependence
$\eta_p \sim \mathrm{Wi}^{-2/3}$ in the limit $\mathrm{Wi} \to \infty$
\cite{bird:87,oett:96,rubi:03,wink:06_1,wink:10}. The differences
in the observed behavior can be explained by a broad crossover
regime before the asymptotic behavior is reached.
The ratio of the viscosities of the two lengths is approximately
$1.33$ for the large Weissenberg-number regime, as for the
alignment angle, which compares well with the theoretically
predicted length dependence in Eq.~(\ref{align_theo}) (cf. Sec. IV
A. 2).
\subsection{Normal stress coefficient}
The concentration and shear rate dependencies of the first and
second normal stress difference~\cite{bird:87,oett:96}
\begin{align} \label{stress_diff}
\Psi_1 = & (\sigma_{xx}-\sigma_{yy})/\dot{\gamma}^2 , \\
\Psi_2 = & (\sigma_{yy}-\sigma_{zz})/\dot{\gamma}^2
\end{align}
are displayed in Fig.~\ref{Psi}. Within the accuracy of the
simulations, the ratio $\Psi_1/\Psi_1^0$, where $\Psi_1^0$ is the
stress difference at zero shear rate, is an universal function of
$\mathrm{Wi}_c$ for various concentrations and decreases as $\Psi_1 \sim
\dot{\gamma}^{-4/3}$ for large shear rates. This is consistent
with analytical calculations~\cite{bird:87,oett:96,wink:10},
various computer
simulations~\cite{zylk:91,oett:96,schr:05_1,liu:89,pete:99,
hsie:04}, and experiments~\cite{teix:05, schr:05} for dilute
solutions. Similar to the viscosity, the decrease is related to
the finite polymer extensibility. Both, hydrodynamic and excluded
volume interactions contribute to $\Psi_1$, which is shown in Ref.
\cite{pete:99} for single polymers. Here, we find the power law
\begin{align}
\Psi_1^0 \sim \left( \frac{c}{c^*} \right)^{1.3},
\end{align}
as shown in the inset of Fig.~\ref{Psi}(a). Hence, the first
normal stress difference exhibits a significant dependence on
excluded volume interactions. At large concentrations, $\Psi_1^0$
might saturate; at least, we cannot exclude such a saturation at
the upper limit of the considered concentration range.
Second normal stress differences are presented in
Fig.~\ref{Psi}(b) for various concentrations. At low
concentrations, their values are much smaller than those of
$\Psi_1$, and hence cannot be calculated within the same accuracy,
and the values $\Psi_2^0$ are difficult to find. We therefore
present the simulations results for $\Psi_2$ directly rather than
in scaled form. Similar to $\Psi_1$, the second normal stress
difference decreases as $\Psi_2\sim\dot{\gamma}^{-4/3}$ with
increasing shear rate. Again, excluded volume and hydrodynamic
interactions contribute to $\Psi_2$
\cite{bird:87,oett:96,pete:99}. The ratio $\Psi_2/\Psi_1$ is
concentration dependent, as shown in Fig.~\ref{Psi}(b). At small
$\mathrm{Wi}_c$ and large concentrations, the ratio is close to unity,
decreases with increasing shear rate and assumes a constant value
above a certain $\mathrm{Wi}_c$, which seems to depend on concentration.
The plateau value itself increases with increasing concentration.
A similar plateau has been found in simulations of dilute
solutions in Ref.~\cite{hsie:04}. The concentration dependence of
the plateau value suggests that excluded volume interactions
determine the behavior of the normal stress differences. If
hydrodynamic interactions would be dominant, we would expect a
decrease of the plateau with increasing concentration due to
screening of hydrodynamic interactions by polymer overlap.
\begin{figure}[t]
\includegraphics*[width=0.4\textwidth]{normal_str_1_2_rescal_1.eps}\\
\includegraphics*[width=0.4\textwidth]{normal_str_1_2_rescal_2.eps}
\caption{
First and second normal stress coefficients $\Psi_1$ (a) and $\Psi_2$ (b) for
polymers of length $N_m=50$ and
the concentrations $c/c^*=0.16$ ($\circ$), $c/c^*=0.41$
($\times$), $c/c^*=0.81$ (${\color{red}+}$), $c/c^*=1.63$
(${\color{red}\Diamond}$) and $c/c^*=2.08$
(${\color{blue}\square}$). The solid lines indicate the power-law decay
$\Psi_i \sim \mathrm{Wi}_c^{-4/3}$. Inset in (a): Concentration
dependence of the zero-shear-rate first normal stress coefficient $\Psi_1^0$.
Inset in (b): Ratio of $\Psi_2/\Psi_1$ for the various concentrations.
\label{Psi}}
\end{figure}
\section{Conclusions}
We have calculated conformational, dynamical, and rheological
properties of polymers in dilute and semidilute solution under
shear flow by mesoscale hydrodynamic simulations. At equilibrium,
our simulations confirm the scaling predictions for the
concentration dependence of the radius of gyration and the longest
relaxation time. Moreover, we find signatures for the screening of
excluded volume interactions in the static structure factor.
In shear flow, the polymers exhibit deformation---the polymer is
stretched in flow direction and shrinks in the transverse
directions---and alignment, which depend on shear rate and
concentration. As one of the the main results of the paper, we
have shown that the relative deformation $\delta G_{xx}$ in the
flow direction, the alignment $\tan(2 \chi_G)$, and the viscosity
$\eta/\eta_0$ are universal functions of the
concentration-dependent Weissenberg number $\mathrm{Wi}_c = \dot \gamma
\tau(c)$ \cite{galu:10,hur:01}. This is surprising because $\tau$
increases rapidly with increasing concentration. Moreover, it
indicates that the dynamics under shear flow is still governed by
the relaxation time at equilibrium despite the anisotropic
deformation of a polymer. This is not evident a priori, as
expressed by the scaling behavior of the radius of gyration tensor
components $G_{yy}$ and $G_{zz}$. Here, we find a
concentration-independent scaling behavior at large $\mathrm{Wi}_c$ only
when these values are scaled by their equilibrium values in {\em
dilute solution}. Hence, the deformations transverse to the flow
directions seem to exhibit a scaling behavior corresponding to a
dilute solution, however, with the relaxation time of the
concentrated system.
In addition, the zero-shear viscosity obeys the scaling
predictions with respect to the concentration dependence.
Moreover, for the first time, we show by simulations that the
Huggins constant is equal to $k_H=0.35$ for a flexible polymer in
good solvent, which is in close agreement with experimental
results \cite{pami:08}.
Finally, we find a strong concentration dependence of the normal
stress differences. Their ratio shows that intermolecular excluded
volume interactions determine their behavior at all shear rates.
Our simulations reveal a complex interplay between shear rate,
deformation, and intramolecular excluded volume interactions,
which is difficult to grasp by analytical theory.
\begin{acknowledgments}
The financial support by the Deutsche Forschungsgemeinschaft (DFG)
within SFB TR6 (project A4) is gratefully acknowledged. We are
grateful to the J\"ulich Supercomputer Centre (JSC) for allocation
of a special CPU-time grant.
\end{acknowledgments}
|
\section{Introduction}
The discovery of iron-based arsenide superconductor
LaFeAsO$_{1-y}$F$_{y}$\cite{Hosono-LaF} has triggered enormous
enthusiasm in searching new high transition temperature ($T_c$)
superconductors in Fe or Ni-based
pnictides\cite{ChenXH-SmOF,WangNL-CeOF,WenHH-LaSr,Hosono-LaNi,LiZ-LaNiAsOF}.
$T_c$ has been raised up to above 50 K when La is replaced by
other rare earth elements in the so-called 1111 type family
\cite{ZhaoZX-LnOD,WangC-GdTh}. The parent compound of the 1111
type pnictides, $LnTPn$O ($Ln =$ rare earths, $T =$ transition
metals, $Pn $= As, P), is of the prototype of ZrCuSiAs structure,
where the conducting $TPn$ layer is sandwiched by the insulating
$Ln$O layer. So far superconductivity (SC) can be induced in the
Fe(Ni)As-based pnictides by many different ways of chemical doping
or by applying high pressure.
Among all these $LnTPn$O's, the members $Ln =$ Ce show extremely
interesting electronic properties. CeFeAsO, as a prototype parent
compound of Fe-based superconductor, is an itinerant bad metal. A
spin-density-wave(SDW)-like antiferromagnetic (AFM) transition of
$d$ electrons was observed around $\sim 140$ K, and the AFM
transition of the Ce$^{3+}$ local moments was observed below 4 K
\cite{Dai-CeNeutron}. SC has been induced by electron doping (i.e.
F-for-O doping) when the SDW long range order of the Fe
3$d$-electrons is suppressed\cite{WangNL-CeOF}. On the other hand,
CeFePO is a paramagnetic (PM) heavy-fermion (HF) system which is
near the ferromagnetic (FM) instability\cite{Geibel-CeP,XuZA-CeP}.
By comparison, CeOsPO shows the Ce$^{3+}$ AFM order at 4.5 K while
CeRuPO is a rare example of the FM Kondo lattice with $T_{C} =$ 15
K\cite{Geibel-CeRu/OsPO}. For the case of $T =$ Co, both CeCoPO and
CeCoAsO show the FM correlated Co-$3d$
magnetism\cite{Geibel-CeCoPO,Ohta-LnCoAsO,Geibel-CeCoAsO}.
It is noted that the nickel arsenide LaNiAsO is a low $T_c$
superconductor and $T_c$ can be enhanced upon F-doping
\cite{Hosono-LaNi,LiZ-LaNiAsOF}. Interestingly, the partial
substitution of Fe by Ni in LaFeAsO, which introduces two more
3$d$ electrons by each Ni$^{2+}$ dopant, can induce SC in a very
narrow doping regime \cite{CaoGH-LaFeNi}. Moreover, the normal
state resistivity shows an upturn behavior at low temperatures,
suggesting a possible Kondo-effect induced semiconductor behavior.
However, there are very few reports on the physical properties of
1111 type NiAs-based pnictides other than LaNiAsO. The early
report did not observe SC above 2 K and suggested a suspicious FM
order in CeNiAsO \cite{WangNL-CeNiAsO}, although its structure and
the valence of Ce ion has been studied \cite{Balnchard-CeNiAsO}.
In this paper, we report the systematic investigation on the
physical properties of CeNiAsO. No evidence of SC is observed for
temperature down to 30 mK. While there is no local magnetic moment
on Ni ions, two successive antiferromagnetic transitions related
to Ce ions are observed. We propose that the Ce moments could
align in "G" type AFM order first at $T_{N1}$ = 9.3 K, and then
transform into "C" type at a lower temperature $T_{N2}$ = 7.3 K.
The Sommerfeld coefficient is about 203 mJ/mol$\cdot$K$^{2}$,
indicating that CeNiAsO is an enhanced correlated Kondo lattice
compound. All these results imply a strong hybridization between
4$f$- and 3$d$-electrons and highlight the important role of the
3$d$-4$f$ interlayer Kondo physics in nickel based cerium
containing pnictide.
\section{Experimental}
Poly-crystalline CeNiAsO sample of high purity was synthesized by
solid state reaction. Ce, Ni, As, and CeO$_{2}$ of high
purity($\geq$ 99.95\%) were used as starting materials. Firstly,
CeAs was presynthesized by reacting Ce discs and As powders at
1320 K for 72 h. NiAs was presynthesized by reacting Ni and As
powders at 970 K for 20 h. Secondly, powders of CeAs, CeO$_{2}$,
Ni, and NiAs were weighted according to the stoichiometric ratio,
thoroughly ground, and pressed into a pellet under a pressure of
600 MPa in an Argon filled glove box. The pellet was packed in an
alumina crucible and sealed into an evacuated quartz tube, which
was then slowly heated to 1450 K and kept at that temperature for
40 h. For comparison, polycrystalline sample of LaNiAsO was also
synthesized, in the similar process, where La$_{2}$O$_{3}$ was
used as the oxygen source.
Powder X-ray diffraction (XRD) was performed at room temperature
using a D/Max-rA diffractometer with Cu-K$_{\alpha}$ radiation and
a graphite monochromator. Lattice parameters were derived by
Rietveld refinement using the programme RIETAN 2000
\cite{Rietan-2000}. X-ray Photoelectron Spectrum (XPS) analysis
was carried out by using a VG ESCALAB MARK \Rmnum{2} device with
Mg-K$\alpha$ ($h\nu =$ 1253.6 eV) nonmonochromatized source
operating in Constant Analysis Energy (CAE, 50 eV) mode, and data
were collected in a step of 0.2 eV. Before taking the XPS data,
the CeNiAsO sample was pre-polished by the argon beam in the
vacuum to make a newly polished sample surface. Electrical
resistivity was measured with a standard four-probe method in a
Quantum Design physical property measurement system (PPMS-9, for
$T \geq$ 2 K) and a dilution refrigerator (for $T <$ 2 K), while
Hall coefficient was measured by scanning field from -5 T to 5 T.
Thermopower was measured by a steady-state technique and a pair of
differential type E thermocouples was used to measure the
temperature gradient. Specific heat was measured by heat pulse
relaxation method in PPMS-9. The dc magnetization measurement was
carried out in a Quantum Design magnetic property measurement
system (MPMS-5) in zero-field-cooling (ZFC) and field-cooling (FC)
protocals under a magnetic field $H$ of 1000 Oe.
\section{Results and discussion}
Fig.~1 shows Rietveld refinement of XRD patterns of CeNiAsO. All
the XRD peaks can be well indexed in to the tetragonal
ZrCuSiAs-type structure with P4/$nmm$ (No.129) space group, and no
obvious impurity trace can be found. The quality factor ($\chi^2$)
of this refinement reaches to as low as 0.88, guaranteeing the
goodness of sample quality. The refined structural parameters are
listed in Tab.~1 with a comparison to CeFeAsO, from which we can
find that CeNiAsO has a slightly longer $a$-axis but much shorter
$c$-axis compared to its iron based counterpart. Because arsenic
is drawn closer to the Ni plane, the thickness of NiAs layer,
$d_{NiAs}$, is much smaller than $d_{FeAs}$, while the angle of
As-$T$-As ($T =$ Fe or Ni, see the inset of Fig.~1) is enlarged.
Fig.~2 shows temperature dependence of dc magnetic susceptibility
of CeNiAsO measured under $H =$ 1000 Oe. The data above 150 K
exhibit good Curie-Weiss behavior, and can be well fitted to
$\chi=\chi_0+C/(T-\theta_W)$, with $\chi_0$ being the temperature
independent item, and $\theta_W$ the Weiss temperature. The
fitting reports that $\theta_W$ = -31.8 K. The obtained effective
moment $\mu_{eff} =$ 2.24 $\mu_B$ is slightly smaller than that of
a free Ce$^{3+}$ ion, 2.54 $\mu_{B}$. For $T <$ 100 K, we observed
a change in the slope of the temperature dependence of inverse
susceptibility (right axis of Fig.~2), and the fitting parameters
are $\mu_{eff} =$ 2.16 $\mu_B$, $\theta_W =$ -25.6 K. The change
in the slope should be ascribed to the crystal electric field
(CEF) effect. It has been demonstrated both theoretically and
experimentally that Ni ions do not exhibit magnetic ordering in
the nickel based
pnictides\cite{LiZ-LaNiAsOF,JiangS-LaFeNi,XuG-LaOMAs,McQueen-LaNiPO,Ronning-Ni2X2}.
Therefore, the Curie-Weiss analysis suggests that the observed
effective moments in CeNiAsO should come from Ce ions. When cooled
to low temperature, a peak in susceptibility is observed around
9.3 K, followed by an upturn when further cooled. Since the Weiss
temperature derived from the Curie-Weiss analysis is negative, it
is reasonable to conclude that Ce moments are ordered
antiferromagnetically below 9.3 K. The AFM ordering of Ce ions is
compatible with the linear field dependence of magnetization (data
not shown here), as well as the small magnitude of susceptibility
in comparison with CeFeAsO \cite{XuZA-CeP}.
One possible interpretation to the noticeable deviation of the
obtained effective moment from that of the free Ce$^{3+}$ ion is
the presence of valence state Ce$^{4+}$ which has no 4$f$-electron
and local moment. To investigate the valence of Ce in this
compound, we thus carried out the XPS experiment. The XPS spectrum
of CeNiAsO is shown in Fig.~3. After subtracting the background,
the XPS spectrum can be well fitted by a combination of ten
Gaussian peaks, i.e., Ce(\Rmnum{3}) = $v^0$+$v'$+$u^0$+$u'$, and
Ce(\Rmnum{4}) = $v$+$v''$+$v'''$+$u$+$u''$+$u'''$, with $u$ and
$v$ corresponding to the spin-orbit split 3$d_{3/2}$ and
3$d_{5/2}$ core holes, respectively. The positions of these peaks
are taken from Ref.\cite{Beche-CeXPS}, where XPS spectrum of
CePO$_4$ (Ce$^{3+}$) and CeO$_2$ (Ce$^{4+}$) were presented. We
found that Ce$^{3+}$ is dominant in CeNiAsO, but there is a small
trace of Ce$^{4+}$ which should account for the small peak around
916 eV \cite{CeRhSb}. An early report on the Ce 3$d$ XPS and Ce
$L_3$-edge XANES analysis of CeNiAsO and CeFeAsO by Balnchard et
al suggested that the valence of Ce should be close to $3+$
\cite{Balnchard-CeNiAsO}. In other Ce containing compounds with
strong Kondo interaction, for example, in CePd$_2$Al$_3$
\cite{CePdAl}, it is common that Ce has a valence very close to
3+. Therefore, although the valence effect may play some role in
the reduction of effective magnetic moment of Ce, a more likely
interpretation to this reduction could mainly come from the CEF
effect, which will be discussed hereinafter.
The transport properties of CeNiAsO are presented in Fig.~4. The
resistivity exhibits several prominent features. Firstly, the
magnitude of resistivity at room temperature, i.e., $\rho_{300
K}$, has a small value of 3.9 $\mu\Omega\cdot$m, which is about
two orders of magnitude smaller than that of CeFeAsO (see in
Tab.~1). The good metallicity of CeNiAsO can be also confirmed by
Hall coefficient $R_H$ measurement, as shown in the lower inset.
$R_H$ remains as a nearly temperature independent constant above
150 K, and converges to 0.68$\times$10$^{-4}$cm$^{3}$/C. A rough
estimate from single band model leads to a carrier density of
$\sim$10$^{24}$cm$^{-3}$, which is an upper limit of charge
carrier density in a multi-band system, about 3 orders of
magnitude larger than those of
$Ln$FeAsO\cite{McGuire-CePrNd,Tao-Nernst,Ishida-carrier}.
Secondly, in contrast to LaNiAsO, the resistivity in CeNiAsO shows
a hump around 100 K. This could be ascribed to the f-electron
contribution via Kondo scattering between the localized 4$f$- and
the conduction 3$d$-electrons. We subtract the resistivity of
LaNiAsO from CeNiAsO, and the difference is shown in the upper
inset. A broad maximum at 92 K becomes obvious. Similar broad peak
is also observed in temperature dependent thermopower $S(T)$. Such
a phenomenon is a distinct feature in Kondo lattice system, and
can be explained by the Kondo scattering from different CEF
levels\cite{Amato-S,Zlatic-S}. However, we can not exclude the
possibility of onset of coherent Kondo scattering, provided that
both have a comparable temperature scale. Thirdly, when
temperature decreases down to around 9.3 K, a pronounced drop is
observed, which is reminiscent of the reduction of spin-flip
scattering when Ce moments become AFM ordered. Such a drop in
resistivity was also observed in previous report
\cite{WangNL-CeNiAsO}. Moreover, as further cooled, a small kink
emerges around 7.3 K. We associate this kink with the
transformation of Ce magnetic structure proposed hereinbefore.
Finally, while LaNiAsO is a superconductor with $T_c \sim$ 2.75 K
\cite{LiZ-LaNiAsOF}, no SC in CeNiAsO is observed down to 30 mK,
implying the influence of Kondo coupling on the conduction
carriers. A small residual resistivity of $\rho_0=$0.24
$\mu\Omega\cdot$m is derived, which again ensures the high purity
of the sample. It should be noted that both $R_{H}$ and $S$ are
positive above 150 K, and become negative below 100 K, manifesting
a multi-band nature of the system in which hole-type carrier
dominates at high temperature while electron-type charge carrier
becomes dominant at low temperature. Whether this change of
carrier type is CEF effect related needs more investigations.
The result of specific heat measurement is shown in Fig.~5. For
high temperature, $C(T)$ follows good Dulong-Petit law and
saturates to the classical limit of 4$\times$3$R\sim$100
J/mol$\cdot$K (data not shown here). Considering the contributions
from the phonons and the Schottky anomaly, the specific heat can
be written as:
\begin{eqnarray}
C=\gamma_0 T+\beta T^3+C_{Sch}
\end{eqnarray}
for temperatures below $\Theta_D$/10, where $\Theta_D$ is the
Debye temperature, $\gamma_0$ and $\beta$ are coefficients of the
electron and phonon contributions, while $C_{Sch}$ is the Schottky
anomaly item. We first subtract the specific heat of LaNiAsO from
the total specific heat, based on the assumption that both of them
have the similar phonon contribution, to estimate the contribution
from $4f$ electrons in CeNiAsO, and the difference ($C^{4f}$) is
shown in the inset of Fig.5(a). $C^{4f}/T$ does not go to zero at
higher temperature ($T
>$ 30 K), but shows a broad peak centering at 50 K. This broad
peak should be attributed to the Schottky anomaly caused by the
excitations between different CEF levels. A Schottky anomaly
formula with three Kramers doublets (one doublet ground state and
two excited doublets)\cite{Geibel-CeRu/OsPO, Dai-CeCEF} is used to
fit this broad peak, i.e.,
\begin{eqnarray}
C_{Sch}(T) = \frac{R}{g_0+g_1\exp(-\Delta_1/T)+g_2\exp(-\Delta_2/T)} \nonumber \\
~~~~~~~~~~~~\times\{g_0g_1(\Delta_1/T)^2\exp(-\Delta_1/T)+g_0g_2(\Delta_2/T)^2\exp(-\Delta_2/T) \nonumber \\
~~~~~~~~~~~~+g_1g_2[(\Delta_2-\Delta_1)/T]^2\exp[-(\Delta_2-\Delta_1)/T]\}
\end{eqnarray}
where $g_{i} $= 2 is the degeneracy of the $i$th doublet state,
and $\Delta_i$ is the energy difference between ground state and
the $i$th excited state. The obtained energy differences from the
fitting are $\Delta_1 =$ 10.5 meV ($\sim$ 120 K) and $\Delta_2 =$
32.8 meV ($\sim$ 380 K). This result is consistent with the slope
change in $1/\chi(T)$, as well as the observed pronounced broad
maximum observed in both $\rho_{Ce}-\rho_{La}$ and $S$. Due to the
large $\Delta_2$, it is likely that the reduction of effective Ce
moment is caused by the CEF effect. We then subtract the fitted
Schottky anomaly item from the total specific heat, and the result
shows a good linear dependence in the $(C-C_{Sch})/T-T^2$ plot
(see the inset of Fig.~5(b)) for the temperatures below 30 K. The
extrapolated Sommerfeld coefficient $\gamma_0 \sim$ 203
mJ/mol$\cdot$K$^2$, which is more than 40 times of that of LaNiAsO
(see in Tab.~1), manifesting the correlated effect contributed
from the Ce 4$f$-electrons. Although the phonon contribution and
the CEF effect is hard to be removed completely, our analysis
should provide a good estimation of the Sommerfeld coefficient of
CeNiAsO. The linear fit also produces that $\beta =$ 0.2511
mJ/mol$\cdot$K$^4$ and Debye temperature $\Theta_D =$ 314 K, which
indicates the above analysis is quite self-consistent. For $T <$
10 K, two $\lambda$-shaped peaks are observed on the $C/T$ curve,
implying two successive phase transitions. This can be further
confirmed by the derivative of susceptibility ($Td\chi/dT$), the
derivative of resistivity ($d\rho/dT$), the magnetoresistivity
($\rho_{5T}-\rho_0$), and the Hall coefficient ($R_H$) shown in
Fig.~5(b,c). The excellent agreement among $C/T$, $d\rho/dT$ and
$Td\chi/dT$ leads to a definition of two characteristic
temperatures, i.e., $T_{N1} =$ 9.3 K and $T_{N2} =$ 7.3 K.
Actually, similar analysis has been widely applied in other nickel
containing compounds, e.g., CeNiGe$_3$ \cite{Mun-CeNiGe3} and
$Ln$Ni$_2$B$_2$C \cite{Ribeiro-RNiBC}. Note that the magnetic
entropy gain derived by integrating $C^{4f}/T$ over temperature
reaches 70\% of $R ln2$ at 10 K, and recovers the full doublet
ground state ($R ln2$) at 30 K. All these indicate that both
specific heat peaks are due to Ce-$4f$ electrons related magnetic
transitions, though the ordered moment is partially screened by
the Kondo coupling in the ground state doublet. The Kondo scale is
estimated to be $T_K \sim$ 15 K by the entropy.
A ratio of $T_K/T_{N1} \sim$ 1 can also be obtained by judging
from the specific heat jump $\Delta (C^{4f}-C_{Sch})$ at $T_{N1}$
\cite{Besnus-TK/TN}, thus it is reasonable that the Kondo scale
$T_K \sim$ 15 K. We propose that the transition at 9.3 K is
originated from the Ce-AFM transition based on the magnetization
measurement. We also notice that the similar phenomenon was also
found in other Ce-based Kondo lattice compounds such as
CeCu$_2$(Si$_{1-x}$Ge$_{x}$)$_2$, where the two peaks were
explained by two incipient instabilities in the magnetic structure
of pure CeCu$_2$Ge$_2$ \cite{Trovarelli-CeCuSi,HQYuan}: the first
one at a higher temperature, related to a reorientation of the
moments, and a second one at lower temperature, related to a
lock-in of the propagation vector.
Now we turn to discuss the possible magnetic configurations of
CeNiAsO. First of all, we should emphasize that all the observed
magnetism should arise from the Ce moments. The negative value of
the Weiss temperature ($\theta_W$) indicates that Ce moments are
AFM correlated. The most simplified model involves the intra-layer
and inter-layer magnetic interactions between the nearest neighbor
Ce moments, denoted by $J^{intra} = J_{0} + J^{intra}_{RKKY}$ and
$J^{inter} = J^{inter}_{RKKY}$, respectively. Where, $J_{0}$ is
the superexchange interaction via O$^{2-}$ and As$^{3-}$ anions in
the absence of $d$-$f$ coupling \cite{DaiJH-CeFe}, while
$J^{intra}_{RKKY}$ and $J^{inter}_{RKKY}$ are the
Ruderman-Kittel-Kasuya-Yosida (RKKY) interactions mediated by
conduction electrons via $d$-$f$ coupling. Generally, $J_{0}$ is
always negative ($ J_{0}<0$) and favors AFM ordered ground
state\cite{DaiJH-CeFe}, while $J^{intra}_{RKKY}$ and
$J^{inter}_{RKKY}$ are functions of Kondo coupling ($J_K$) and
density of state at Fermi energy($N_{E_F}$) \cite{Doniach}. We
can assume $J^{inter}_{RKKY}\sim J^{intra}_{RKKY}=J_{RKKY}$, since
they have almost the same $k_F$ and the same nearest neighbor
distances \cite{note2}. $J_{RKKY}$ may be either negative
(antiferromagnetic) or positive (ferromagnetic) depends on
$J_KN_{E_F}$, which may account for either a G-type or C-type
magnetic order respectively when cooled down (see the lower inset
of Fig.~1). Actually, the changes in the Hall coefficient and
thermopower at low temperatures indeed imply a temperature-induced
change, either quantitatively or qualitatively, in the either
density or relaxation time of different types of charge carriers.
Thus it is plausible that a sign change occurs in $J_{RKKY}$ when
temperature is lowered down to $7K\sim 9K$, resulting in the
magnetic transition from the G-type to C-type magnetic
configuration within the Ce-AFM phase. This scenario provides a
possible explanation on the two magnetic phase transitions, and it
is consistent with the drop in the magnetic susceptibility
($\chi(T)$) in the G-type state and then an upturn as it enters
C-type state at $T < T_{N2}$. Further experimental studies on the
single crystalline sample of CeNiAsO and neutron diffraction
measurement are necessary to testify this scenario.
Finally, it is interesting to compare CeNiAsO with its neighbors,
CeFeAsO and CeCoAsO. Previous first principles calculation on
La$T$AsO \cite{XuG-LaOMAs} suggested that Fe and Co ions show
SDW-like and FM instabilities at low temperatures respectively,
while Ni ions display nonmagnetic behavior. For Ce$T$AsO, Fe ions
in CeFeAsO indeed show SDW-like AFM instability and Co ions in
CeCoAsO are FM ordered\cite{Ohta-LnCoAsO,Geibel-CeCoAsO},
meanwhile neither local magnetic moment nor long range magnetic
order of Ni ions is observed in CeNiAsO. However, this does not
mean that the Kondo coupling between Ce moment and conduction
carriers in the NiAs layer is weaker than that in the FeAs layer.
The situation could be totally on the contrary. We expect the
Ce-4$f$ level is much closer to the conduction band in CeNiAsO
than those in CeFeAsO and CeCoAsO. Thus an enhanced hybridization
between 4$f$- and 3$d$-electrons is possible in CeNiAsO due to the
smaller $c$-axis lattice constant (see Tab.~1), as evidenced by
the moderately large Sommerfeld coefficient and the absence of SC.
\section{Conclusion}
To summarize, the magnetic properties and specific heat of
poly-crystalline CeNiAsO sample with high purity have been
studied. Although CeNiAsO is much more metallic compared to its
neighbor CeFeAsO, no superconductivity is observed down to 30 mK.
Two successive antiferromagnetic transitions of Ce 4f electrons
are observed. We propose that the Ce moments could align in G-type
AFM at $T_{N1}$=9.3 K first, and then modify into C-type at a
lower temperature $T_{N2}$=7.3 K. Regarding to the large
Sommerfeld coefficient of about 203 mJ/mol$\cdot$K$^{2}$, CeNiAsO
represents a new example of AFM dense Kondo lattice with Kondo
scale $T_K \sim$ 15 K in a wide class of the rare earth pnictides
of the ZrCuSiAs-type \cite{Rottgen}. A strong hybridization
between 4$f$- and 3$d$-electrons is proposed and its influence on
the ground state is discussed.
\section*{Acknowledgments}
The authors would like to thank Qimiao Si, Huiqiu Yuan and Shuang
Jia for helpful discussion. This work is supported by the National
Science Foundation of China (Grant Nos 10634030 and 10931160425),
the Fundamental Research Funds for the Central Universities of
China (Program No. 2010QNA3026), and the National Basic Research
Program of China (Grant Nos. 2007CB925001 and 2010CB923003).
\section*{References and Notes}
|
\section{Introduction.}
Variational inequalities in probability, ergodic theory and harmonic analysis
have been the subject of many recent research papers. One important character of these
inequalities is the fact that they can be used to measure the speed of convergence
for the family of operators in consideration. To be more precise, consider, for instance,
a measure space $(\Omega,\mu)$ and an operator $T$ on $L^1(\Omega)+ L^\infty(\Omega)$.
Form the ergodic averages of $T$:
$$
M_n(T)\,=\,\frac{1}{n+1}\,\sum_{k=0}^{n} T^k\,,\qquad n\geq 0.
$$
A fundamental theorem in ergodic theory states that if $T$ is a contraction on
$L^p(\Omega)$ for every $1\le p\le\infty$, then the limit
$\lim_{n\to\infty}M_n(T)x$ exists a.e. for every $x\in L^p(\Omega)$.
One can naturally ask what is the speed of convergence of this limit.
A classical tool for measuring that speed is the following square function
$$
S(x)=\Bigl(\sum_{n\ge0} n\bigl|M_{n+1}(T)x -M_n(T)x \bigr|^2\Bigr)^{\frac12},
$$
the problem being to estimate its norm $\|S(x)\|_p$. This issue goes back to Stein \cite{S2},
who proved that if $T$ as above is positive on $L^2(\Omega)$ (in the Hilbertian sense),
then a square function inequality
$$
\|S(x)\|_p\le C_p\|x \|_p,\quad x\in L^p(\Omega),
$$
holds for any $1<p<\infty$.
Stein's inequality is closely related to Dunford-Schwartz's maximal ergodic inequality,
$$
\big\|\sup_{n\ge0}|M_n(T)x|\,\big\|_p\le C_p\|x\|_p,\quad x\in L^p(\Omega), \; 1<p\le\infty.
$$
This maximal inequality and its weak type $(1,1)$ substitute for $p=1$ are key
ingredients in the proof of the previous pointwise ergodic theorem.
The strong $q$-variation is another (better) tool to measure
the speed of $\lim_{n\to\infty}M_n(T)x$.
Bourgain was the first to consider variational inequalities in ergodic theory.
To state his inequality we need to recall the definition of the strong $q$-variation.
Given a sequence $(a_n)_{n\geq 0}$ of complex numbers and a number $1\leq q<\infty$,
the strong $q$-variation norm is defined as
$$
\norm{(a_n)_{n\geq 0}}_{v^q}\,=\, \sup\Bigl\{\bigl(\vert a_0\vert^q\, +\,\sum_{k\geq 1}
\vert a_{n_k}-a_{n_{k-1}}\vert^q \bigr)^{\frac{1}{q}}\Bigr\},
$$
where the supremum runs over all increasing sequences $(n_k)_{k\geq 0}$
of integers such that $k_0=0$.
It is clear that the set $v^q$ of all sequences with a finite strong $q$-variation is a Banach space
for the norm $\norm{\ }_{v^q}$.
Bourgain \cite{B} proved that if $T$ is induced by a measure preserving transformation
on $(\Omega,\mu)$, then for any $ 2<q<\infty$,
$$
\bignorm{(M_n(T)x)_{n\geq 0}}_{L^2(v^q)}\,
\leq\, C_{q}\,\norm{x}_2, \quad x\in L^2(\Omega).
$$
This inequality was then extended to $L^p(\Omega)$ for any $1<p<\infty$ by Jones,
Kaufman, Rosenblatt and Wierdl \cite{JKRW}. The latter paper contains many other
interesting results on the subject. Note that the predecessor of Bourgain's inequality
is L\'epingle's variational inequality for martingales \cite{Le}. The latter says
that if $(\mathbb E_n)_{n\geq 0}$ is an increasing sequence of conditional expectations
on a probability space $\Omega$, then we have an estimate
$\bignorm{(\mathbb E_n(x))_{n\geq 0}}_{L^p(v^q)}\,
\leq\, C\,\norm{x}_p$
(see also \cite{PX2} for further results on that theme).
Since \cite{B}, variational type inequalities have been extensively studied
in ergodic theory and harmonic analysis. Many classical sequences of operators
and semigroups have been proved to satisfy strong variational bounds, see in particular
\cite{CMMTV, JR, JSW, JW, OSTTW} and references therein.
The main purpose of this paper is to exhibit a large class of operators
$T$ on $L^p(\Omega)$ for a fixed $1<p<\infty$ with the following property:
for any $2<q<\infty$, there exists a constant $C>0$ (which may depend on $q$ and $T$)
such that
for any $x\in L^p(\Omega)$, the sequence $(T^n(x))_{n\geq 0}$
belongs to $L^p(\Omega;v^q)$, and
\begin{equation}\label{1q1}
\bignorm{(T^n(x))_{n\geq 0}}_{L^p(v^q)}\,
\leq\, C\,\norm{x}_p.
\end{equation}
We show that this holds true provided that $T$ is a positive contraction (more generally,
a contractively regular operator) and $T$ is analytic, in the sense that
$$
\sup_{n\geq 1} n\norm{T^n-T^{n-1}}\,<\,\infty.
$$
Inequality \eqref{1q1} implies, of course, a similar variational inequality
for the ergodic averages $M_n(T)$. However, in this latter case,
the analytic assumption above can be removed.
Namely, for a positive contraction $T$ on $L^p(\Omega)$ with $1<p<\infty$ we have
an estimate
\begin{equation}\label{1q3}
\bignorm{(M_n(T)x)_{n\geq 0}}_{L^p(v^q)}\,\leq\,
C\,\norm{x}_p,\qquad x\in L^p(\Omega),
\end{equation}
for any $2<q<\infty$. This result extends those of \cite{B} and \cite{JKRW} quoted previously.
Note that inequality \eqref{1q1} for positive analytic contractions
considerably improves our previous maximal ergodic inequality for such operators $T$ proved in \cite{LMX}.
In this sense, this paper is a continuation of \cite{LMX}. On the other hand,
our proof of \eqref{1q1} heavily relies on the square function inequality of \cite{LMX}.
We also establish results similar to (\ref{1q1}) and (\ref{1q3})
for strongly continuous semigroups.
This requires the following continuous analog of $v^q$.
Given a complex family $(a_t)_{t>0}$, define
$$
\norm{(a_t)_{t>0}}_{V^q}\,=\, \sup\Bigl\{\bigl(\vert a_{t_0}\vert^q\, +\,\sum_{k\geq 1}
\vert a_{t_k}-a_{t_{k-1}}\vert^q \bigr)^{\frac{1}{q}}\Bigr\},
$$
where the supremum runs over all increasing sequences $(t_k)_{k\geq 0}$
of positive real numbers. Then we let
$V^q$ be the resulting Banach space of
all $(a_t)_{t>0}$ such that
$\norm{(a_t)_{t> 0}}_{V^q}\,<\,\infty\,$.
Consider a bounded analytic semigroup $(T_t)_{t\geq 0}$ on $L^p(\Omega)$, with
$1<p<\infty$. We show that if $T_t$ is a positive contraction (more generally,
a contractively regular operator) for any $t\geq 0$, then for any $2<q<\infty$
and any $x\in L^p(\Omega)$,
the family $\bigl([T_t(x)](\lambda)\bigr)_{t>0}$ belongs to $V^q$ for almost every $\lambda\in
\Omega$, and we have an estimate
\begin{equation}\label{1q2}
\Bignorm{\lambda\mapsto \bignorm{\bigl([T_t(x)](\lambda)\bigr)_{t>0}}_{V^q}}_p
\,\leq\,C\,\norm{x}_p,\qquad x\in L^p(\Omega).
\end{equation}
We mention that Jones and Reinhold \cite{JR} proved (\ref{1q2}) for positive, unital
symmetric diffusion semigroups and (\ref{1q1}) for a certain class of convolution
operators (only in the case $p=2$). Our results turn out
to extend these contributions in various directions.
As in the discrete case, we obtain similar results for the averages
of the semigroup if we remove the analyticity assumption.
Inequalities for ergodic averages, such as (\ref{1q3}), will be established in Section 3.
Then our main results leading to (\ref{1q1}) and (\ref{1q2}) will be established in
Section 4, using the above mentioned results from Section 3, as well as square
function estimates from \cite{LMX}.
Section 5 is devoted to various complements. On the one hand, we establish individual (= pointwise)
ergodic theorems in our context. For example, if $T\colon L^p(\Omega)\to L^p(\Omega)$
is a positive analytic contraction, then $\lim_{n\to\infty} T^n(x)$ exists a.e. for
every $x\in L^p(\Omega)$. On the other hand,
it is well-known that (\ref{1q3}) cannot be extended
to the case $q=2$. Then we show analogs of (\ref{1q3}) and (\ref{1q1}) when $v^q$
is replaced by the oscillation space $o^2$. We give examples and applications
in Section 6.
\bigskip
We end this introduction with a few notation. If $X$ is a Banach space, we let
$B(X)$ denote the algebra of all bounded operators on $X$ and we let
$I_X$ denote the identity operator on $X$ (or simply $I$ if there is no ambiguity
on $X$). For any $T\in B(X)$, we let
$\sigma(T)$ denote the spectrum of $T$. Also we let
$\Ddb=\{z\in\Cdb\, :\, \vert z\vert <1\}$ be the open unit disc of $\Cdb$.
For a measurable function $x\colon\Omega\to\Cdb$ acting on
a measure space $(\Omega,\mu)$, $\norm{x}_p$ denotes the $L^p$-norm of $x$.
We refer to \cite{Pa} and \cite{Go} for background on strongly continuous
and analytic semigroups on Banach space.
\medskip
\section{Preliminaries on $q$-variation.}
The aim of this section is to provide some elementary background on the spaces $v^q$ and
$V^q$ defined above. We fix some $1\leq q<\infty$.
Let $(a_n)_{n\geq 0}$ be an element of $v^q$. Then the sequence $(a_n)_{n\geq 0}$ is bounded, with
$$
\norm{(a_n)_{n\geq 0}}_{\infty} \leq 2
\norm{(a_n)_{n\geq 0}}_{v^q},
$$
and
$$
\lim_{m\to \infty}\bignorm{(0,\ldots,0,a_{m+1} - a_m,\ldots, a_n-a_m,\ldots)}_{v^q}\,=0.
$$
Thus the space of eventually constant sequences is dense in $v^q$.
Consequently any element of $v^q$ is a converging sequence.
For any integer $m\geq 1$, let $v_m^q$ be the space of $(m+1)$-tuples $(a_0,a_1,\ldots,a_m)$
of complex numbers, equipped with the norm
$$
\bignorm{(a_0,a_1,\ldots,a_m)}_{v^q_m}\,=\,\bignorm{(a_0,a_1,\ldots,a_m,a_m,\ldots)}_{v^q}.
$$
It follows from above that an infinite sequence $(a_n)_{n\geq 0}$ belongs to $v^q$ if and only if there is a constant
$C\geq 0$ such that $\bignorm{(a_0,a_1,\ldots,a_m)}_{v^q_m}\,\leq C$ for any $m\geq 1$ and in this case,
\begin{equation}\label{6Approx}
\norm{(a_n)_{n\geq 0}}_{v^q}\,=\,\lim_{m\to \infty} \bignorm{(a_0,a_1,\ldots,a_m)}_{v^q_m}.
\end{equation}
Let $(\Omega,\mu)$ be a measure space, let
$1<p<\infty$ and let $L^p(\Omega;v^q)$ denote the
corresponding Bochner space. Any element of that space can be naturally regarded
as a sequence of $L^p(\Omega)$. The following is a direct consequence
of the above approximation properties.
\begin{lemma}\label{6Lem1}
Let $(x_n)_{n\geq 0}$ be a sequence of $L^p(\Omega)$, the following assertions
are equivalent.
\begin{itemize}
\item [(i)]
The sequence $(x_n)_{n\geq 0}$ belongs to $L^p(\Omega;v^q)$.
\item [(ii)]
The sequence $(x_n(\lambda))_{n\geq 0}$ belongs to $v^q$ for almost every $\lambda\in\Omega$ and the
function $\lambda\mapsto \norm{(x_n(\lambda))_{n\geq 0}}_{v^q}$ belongs to $L^p(\Omega)$.
\item [(iii)]
There is a constant $C\geq 0$ such that
$$
\bignorm{(x_0,x_1,\ldots,x_m)}_{L^p(v_m^q)}\,\leq C
$$
for any $m\geq 1$.
\end{itemize}
In this case,
$$
\bignorm{\lambda\mapsto \norm{(x_n(\lambda))_{n\geq 0}}_{v^q}}_p\,=\,
\norm{(x_n)_{n\geq 0}}_{L^p(v^q)}\,=\,
\lim_{m\to \infty} \bignorm{(x_0,x_1,\ldots,x_m)}_{L^p(v^q_m)}.
$$
\end{lemma}
\bigskip
We now consider the continuous case. We note for further use that any element $(a_t)_{t>0}$ of $V^q$ admits limits
$$
\lim_{t\to 0^+} a_t\qquad\hbox{and}\qquad \lim_{t\to\infty} a_t.
$$
The following is a continuous analog of Lemma \ref{6Lem1}. Note however that families satisfying the next statement
do not necessarily belong to the Bochner space $L^p(\Omega; V^q)$.
\begin{lemma}\label{6Lem2}
Let $(x_t)_{t>0}$ be a family of $L^p(\Omega)$ and assume that:
\begin{itemize}
\item [(1)] For a.e. $\lambda\in\Omega$, the function $t\mapsto x_t(\lambda)$ is continuous on $(0,\infty)$.
\item [(2)] There exists a constant $C\geq 0$ such that whenever $t_0<t_1<\cdots<t_m$
is a finite increasing sequence of positive real numbers, we have
$$
\bignorm{(x_{t_0},x_{t_1},\ldots,x_{t_m})}_{L^p(\Omega; v^q_m)}\,\leq C.
$$
\end{itemize}
Then $(x_t(\lambda))_{t>0}$ belongs to $V^q$ for a.e. $\lambda\in\Omega$, the function
$\lambda\mapsto \bignorm{(x_t(\lambda))_{t>0}}_{V^q}$ belongs to $L^p(\Omega)$ and
$$
\bignorm{\lambda\mapsto \norm{(x_t(\lambda))_{t>0}}_{V^q}}_{p}\,\leq C.
$$
\end{lemma}
\begin{proof}
For any integer $N\geq 1$, define
$$
\varphi_N(\lambda) \,=\, \sup_{k\geq 1}\bignorm{\bigl(x_{n2^{-N}} (\lambda)\bigr)_{n\geq k}}_{v^q},
\qquad \lambda\in \Omega.
$$
It follows from (\ref{6Approx}) that $\varphi_N$ is measurable. Moreover the sequence
$(\varphi_N)_{N\geq 1}$ is nondecreasing, and we may therefore define
$$
\varphi(\lambda)=\lim_{N\to\infty}\varphi_N(\lambda), \qquad \lambda\in \Omega.
$$
By construction, $\varphi$ is measurable and by the monotone convergence theorem,
its $L^p$-norm is equal to $\lim_N\norm{\varphi_N}_p$.
According to the assumption (2) and the approximation property
(\ref{6Approx}), the $L^p$-norm of $\varphi_N$ is $\leq C$ for any
$N\geq 1$. Hence
$$
\int_{\Omega}\varphi(\lambda)^p\, d\mu(\lambda)\ \leq\, C^p.
$$
This implies that $\varphi(\lambda)<\infty$ for a.e. $\lambda\in\Omega$.
If $\lambda\in \Omega$ is such that $t\mapsto x_t(\lambda)$ is continuous on $(0,\infty)$, then
$$
\varphi(\lambda)=\bignorm{(x_t(\lambda))_{t>0}}_{V^q}.
$$
According to the assumption (1), this holds true almost everywhere. Hence
the quantity $\bignorm{(x_t(\lambda))_{t>0}}_{V^q}$ is finite for a.e. $\lambda\in\Omega$.
The lemma clearly follows from these properties.
\end{proof}
\medskip
\section{Variation of ergodic averages.}
Throughout we let $(\Omega,\mu)$ be a measure space and we let $1<p<\infty$.
We first recall the notion of regular operators on $L^p(\Omega)$ and some of their basic properties
which will be used in this paper. We refer e.g. to \cite[Chap. 1]{MN} and to \cite{Pe,Pi}
for more details and complements.
An operator $T\colon L^p(\Omega)\to L^p(\Omega)$ is called regular if there is a constant
$C\geq 0$ such that
$$
\bignorm{\sup_{k\geq 1}\vert T(x_k)\vert}_p\,\leq\, C\bignorm{\sup_{k\geq 1}\vert x_k\vert}_p
$$
for any finite sequence $(x_k)_{k\geq 1}$ in $L^p(\Omega)$. Then we let $\norm{T}_r$ denote the
smallest $C$ for which this holds. The set of all regular operators on $L^p(\Omega)$ is a
vector space on which $\norm{\ }_r$ is a norm.
We say that $T$ is contractively regular if $\norm{T}_r\leq 1$.
Let $E$ be a Banach space. If an operator $T\colon L^p(\Omega)\to
L^p(\Omega)$ is regular, then the operator $T\otimes I_E\colon L^p(\Omega)\otimes E\to
L^p(\Omega)\otimes E$ extends to a bounded operator on the Bochner space $L^p(\Omega; E)$, and
\begin{equation}\label{2Reg}
\bignorm{T\otimes I_E\colon L^p(\Omega;E)\longrightarrow L^p(\Omega;E)}\,\leq\,\norm{T}_r.
\end{equation}
Indeed by definition this holds true when $E=\ell^{\infty}_n$ for any $n\geq 1$, and the general
case follows from the fact that for any $\varepsilon >0$,
any finite dimensional Banach space is $(1+\varepsilon)$-isomorphic to
a subspace of $\ell^\infty_n$ for some large enough $n\geq 1$.
Any positive operator $T$ (in the lattice sense) is regular and $\norm{T}_r=\norm{T}$ in this case.
Thus all statements given for
contractively regular operators apply to positive contractions. It is well-known that conversely,
$T$ is regular with $\norm{T}_r\leq C$ if and only if there is a positive operator
$S\colon L^p(\Omega)\to L^p(\Omega)$ with $\norm{S}\leq C$, such that
$\vert T(x)\vert \leq S(\vert x\vert)$ for any $x\in L^p(\Omega)$.
Finally, following \cite{Pe}, we say that an operator $T\colon L^{1}(\Omega)+L^{\infty}(\Omega) \to
L^{1}(\Omega)+L^{\infty}(\Omega)$ is an absolute contraction if it induces two contractions
$$
T\colon L^{1}(\Omega) \longrightarrow
L^{1}(\Omega) \qquad\hbox{and}\qquad T\colon L^{\infty}(\Omega) \longrightarrow
L^{\infty}(\Omega).
$$
Then the resulting operator $T\colon L^{p}(\Omega) \to
L^{p}(\Omega)$ is contractively regular.
The main result of this section is the following theorem, which might be known to experts.
Its proof relies on the transference principle, already used in \cite{B}.
\begin{theorem}\label{2Main} Let $T\colon L^p(\Omega)\to L^p(\Omega)$ be a contractively regular operator,
with $1<p<\infty$, and let $2<q<\infty$. Then we have
$$
\bignorm{\bigl(M_n(T)x\bigr)_{n\geq 0}}_{L^p(v^q)}\,\leq\,C_{p,q}\,\norm{x}_p,\qquad x\in L^p(\Omega),
$$
for some constant $C_{p,q}$ only depending on $p$ and $q$.
\end{theorem}
Let
$$
s_p \colon \ell^p_{\footnotesize{\Zdb}} \longrightarrow \ell^p_{\footnotesize{\Zdb}},
\qquad s_p\bigl((c_n)_n\bigr)=(c_{n-1})_{n},
$$
denote the shift operator on $\ell^p_{\footnotesize{\Zdb}}$. According to \cite[Thm. B]{JKRW},
$s_p$ satisfies Theorem \ref{2Main} (the crucial case $p=2$ going back to \cite{B}). Thus
for any $1<p<\infty$ and $2<q<\infty$ we have
a constant
\begin{equation}\label{2Cpq}
C_{p,q}\,=\,\bignorm{c\mapsto \bigl(M_n(s_p)c\bigr)_{n\geq 0}}_{\ell^p_{\Zdb}\to \ell^p_{\Zdb}(v^q)}.
\end{equation}
The following lemma is a variant of the well-known Coifman-Weiss
transference Theorem \cite[Thm. 2.4]{CW} and
is closely related to \cite{DLT} and \cite{ABG}.
\begin{lemma}\label{2Transfer}
Let $U\colon L^p(\Omega)\to L^p(\Omega)$ be an invertible operator such that $U^j$ is regular
for any $j\in \Zdb$ and suppose that $C=\sup\{\norm{U^j}_r\, :\, j\in\Zdb\}\,<\,\infty\,$.
Let $(e_1,\ldots,e_N)$ be a basis of a finite dimensional Banach space $E$, and
let $a(1),\ldots, a(N)$ be $N$ elements of $\ell^1_{\footnotesize{\Zdb}}$.
Consider the two operators
$$
K\colon \ell^p_{\footnotesize{\Zdb}}\longrightarrow \ell^p_{\footnotesize{\Zdb}}(E),\qquad K(c)\,=\,\sum_{k=1}^{N}
\Bigl(\sum_{j\in\footnotesize{\Zdb}} a(k)_j\, s_p^j(c)\Bigr)\otimes e_k,
$$
and
$$
R\colon L^p(\Omega) \longrightarrow L^p(\Omega;E),\qquad R(x)\,=\,\sum_{k=1}^{N}
\Bigl(\sum_{j\in\footnotesize{\Zdb}} a(k)_j \,U^j(x)\Bigr)\otimes e_k.
$$
Then
$$
\norm{R}\leq C^2\norm{K}.
$$
\end{lemma}
\begin{proof}
The operator $I_{L^p(\Omega)}\otimes K$ extends to a bounded operator
$L^p\bigl(\Omega;\ell^p_{\footnotesize{\Zdb}}\bigr)\to L^p\bigl(\Omega;\ell^p_{\footnotesize{\Zdb}}(E)\bigr)$,
whose norm is equal to $\norm{K}$. (Nothing special about $K$ is required for this tensor extension
property.) By Fubini's Theorem,
$$
L^p\bigl(\Omega;\ell^p_{\footnotesize{\Zdb}}\bigr)\simeq \ell^p_{\footnotesize{\Zdb}}
\bigl(L^p(\Omega)\bigr)\qquad\hbox{and}\qquad
L^p\bigl(\Omega;\ell^p_{\footnotesize{\Zdb}}(E)\bigr)\simeq \ell^p_{\footnotesize{\Zdb}}\bigl(L^p(\Omega;E)\bigr)
$$
isometrically. Further,
under these identifications, the extension of
$I_{L^p(\Omega)}\otimes K$ corresponds to the operator
$$
\widetilde{K}\colon \ell^p_{\footnotesize{\Zdb}}\bigl(L^p(\Omega)\bigr)\longrightarrow \ell^p_{\footnotesize{\Zdb}}\bigl(L^p(\Omega; E)\bigr),\qquad \widetilde{K}(z)\,=\,\sum_{k=1}^{N}
\Bigl(\sum_{j} a(k)_j\, \bigl(s_p^j\overline{\otimes} I_{L^p(\Omega)}\bigr)(z)\Bigr)\otimes e_k
$$
Thus we have $\norm{\widetilde{K}}=\norm{K}$.
By approximation, we may suppose that $a(1),\ldots,a(N)$ are finitely supported. Let $m\geq 1$
be chosen such that $a(k)_j=0$ for any $k$ and any $\vert j\vert >m$.
Let $x\in L^p(\Omega)$ and let
$$
y_k = \,\sum_{j=-m}^{m} a(k)_j\, U^j(x)
$$
for any $k=1,\ldots,N$. Our aim is to estimate the norm of $\sum_k y_k\otimes e_k\,$
in $L^p(\Omega;E)$.
For any $i\in\Zdb$, we have
$$
\sum_{k=1}^{N} y_k\otimes e_k\, =\,(U^{i}\otimes I_E)\Bigl(
\sum_{k=1}^{N} U^{-i}y_k\otimes e_k\Bigr).
$$
Hence applying (\ref{2Reg}) to $U^{i}$, we derive that
$$
\Bignorm{\sum_{k=1}^{N} y_k\otimes e_k}_{L^p(\Omega;E)}\, \leq\, C\,\Bignorm{
\sum_{k=1}^{N} U^{-i}y_k\otimes e_k}_{L^p(\Omega;E)}
$$
Let $n\geq 1$ be an arbitrary integer and let $\chi$ be the characteristic function
of the interval $[-(n+m),(n+m)]$. We deduce from the above estimate that
\begin{align*}
(2n+1)\Bignorm{\sum_{k=1}^{N} y_k\otimes e_k}_{L^p(\Omega;E)}^{p}\, & \leq\, C^p
\,\sum_{i=-n}^{n}
\Bignorm{
\sum_{k=1}^{N} \sum_{j=-m}^{m} a(k)_j\,U^{j-i}(x)\otimes e_k}_{L^p(\Omega;E)}^{p}
\\
&
\leq\, C^p
\,\sum_{i}
\Bignorm{
\sum_{k=1}^{N} \sum_{j} a(k)_j\,\chi(j-i)\,U^{j-i}(x)\otimes e_k}_{L^p(\Omega;E)}^{p}.
\end{align*}
Since
$$
\sum_i \Bignorm{
\sum_{k=1}^{N} \sum_{j} a(k)_j\,\chi(j-i)\,U^{j-i}(x)\otimes e_k}_{L^p(\Omega;E)}^{p}
\,=\,\Bignorm{\widetilde{K}\Bigl[\bigl(\chi(-j)U^{-j}
(x)\bigr)_j\Bigr]}^p_{\ell^p_{\Zdb}(L^p(\Omega; E))}
$$
and $\norm{\widetilde{K}}=\norm{K}$, this yields
\begin{align*}
(2n+1)\Bignorm{\sum_{k=1}^{N} y_k\otimes e_k}_{L^p(\Omega;E)}^{p}\, &\leq\, C^p
\norm{K}^p \sum_{j=-(n+m)}^{n+m}\norm{U^{-j}(x)}^p_p\\
& \leq\, \bigl(2(n+m)+1\bigr)\, C^{2p} \norm{K}^p \norm{x}^p_p.
\end{align*}
Letting $n\to\infty$, we get the result.
\end{proof}
\begin{proof}[Proof of Theorem \ref{2Main}]
Let $T\colon L^p(\Omega)\to L^p(\Omega)$ be a contractively regular operator.
There exists another measure space $(\widehat{\Omega},\widehat{\mu})$,
two positive contractions $J\colon L^p(\Omega)\to L^p(\widehat{\Omega})$
and $Q\colon L^p(\widehat{\Omega})\to L^p(\Omega)$ and an isometric invertible
operator $U\colon L^p(\widehat{\Omega})\to L^p(\widehat{\Omega})$ such that
$$
T^k=QU^kJ,\qquad k\geq 0.
$$
In the case when $T$ is positive, this is Akcoglu's famous dilation Theorem (see \cite{A}).
The extension to regular operators stated here is from \cite{Pe} or \cite{CRW}. Moreover
$U$ can be chosen so that $U$ and $U^{-1}$ are both contractively regular.
Thus
$$
\forall\, j\in\Zdb,\qquad \norm{U^j}_r=1.
$$
Note that when $1<p\not= 2<\infty$, any isometry on $L^p$ is contractively regular,
so the latter information is relevant only when $p=2$.
We fix an integer $m\geq 1$ and we consider $v^q_m$ as defined in Section 2.
For any $n\geq 0$, we clearly have
$$
M_n(T)=QM_n(U)J.
$$
Since $\norm{Q}_r\leq 1$, it follows from
(\ref{2Reg}) that
\begin{equation}\label{2Dilation}
\bignorm{\bigl(M_n(T)x\bigr)_{0\leq n\leq m}}_{L^p(\Omega;v^q_m)}\,\leq\,
\bignorm{\bigl(M_n(U)J(x)\bigr)_{0\leq n\leq m}}_{L^p(\widehat{\Omega};v^q_m)},\qquad x\in L^p(\Omega).
\end{equation}
For any $n=0,1,\ldots, m$, let $a(n)\in \ell^1_{\footnotesize{\Zdb}}$ be defined
by letting $a(n)_j=(n+1)^{-1}$ if $0\leq j\leq n$ and $a(n)_j=0$ otherwise. Then
$$
\sum_{j\in\footnotesize{\Zdb}}a(n)_j \, U^j\, =M_n(U) \qquad
\hbox{and}\qquad
\sum_{j\in\footnotesize{\Zdb}}a(n)_j \, s_p^j\, =M_n(s_p).
$$
Applying Lemma \ref{2Transfer} with $E=v^q_m$ and recalling (\ref{2Cpq}), we therefore deduce that
$$
\bignorm{\bigl(M_n(U)z\bigr)_{0\leq n\leq m}}_{L^p(\widehat{\Omega};v^q_m)}\,\leq\, C_{p,q}\,\norm{z}_p
$$
for any $z\in L^p(\widehat{\Omega})$. Combining with the inequality (\ref{2Dilation})
we obtain that
$$
\bignorm{\bigl(M_n(T)x\bigr)_{0\leq n\leq m}}_{L^p(\Omega;v^q_m)}\,\leq\, C_{p,q}\,\norm{x}_p
$$
for any $x\in L^p(\Omega)$. Then the result follows from Lemma \ref{6Lem1}.
\end{proof}
\begin{remark}\label{2Amenable}
The above Lemma \ref{2Transfer} extends without any difficulty to amenable groups, as follows.
Let $G$ be a locally compact amenable group, with left Haar measure $dt$,
let $\pi\colon G\to B(L^p(\Omega))$ be a strongly continuous
representation valued in the space of regular operators on $L^p(\Omega)$,
and assume that
$$
C=\sup\{\norm{\pi(t)}_r\, :\, t\in G\}\,<\infty\,.
$$
Next let $h_1,\ldots, h_N$ be $N$ elements of $L^1(G)$ and let
$K\colon L^p(G)\to L^p(G;E)$ be defined by letting $K(f)=\sum_k (h_k*f)\otimes e_k\,$
for any $f\in L^p(G)$.
Then for any $x\in L^p(\Omega)$, we have
$$
\Bignorm{\sum_k\Bigl(\int_G h_k(t)\pi(t)x\, dt\,\Bigr)\otimes e_k}_{L^p(\Omega;E)}\,\leq
\,C^2\,\norm{K}\,\norm{x}_p.
$$
\end{remark}
\bigskip
We conclude this section with a continuous version of Theorem \ref{2Main}. Given a
strongly continuous semigroup $T=(T_t)_{t\geq 0}$ on $L^p(\Omega)$,
we let
$$
M_t(T)=\,\frac{1}{t}\,\int_{0}^{t} T_s\, ds\,,\qquad t>0,
$$
defined in the strong sense.
\begin{corollary}\label{2Continuous}
Let $T=(T_t)_{t\geq 0}$ be a strongly continuous semigroup on $L^p(\Omega)$ and assume that
$T_t\colon L^p(\Omega)\to L^p(\Omega)$ is contractively regular for any $t\geq 0$.
Let $2<q<\infty$ and let $x\in L^p(\Omega)$. Then for a.e. $\lambda\in\Omega$, the family
$\bigl([M_t(T)x](\lambda)\bigr)_{t>0}$ belongs to $V^q$ and
$$
\Bignorm{\lambda\mapsto \bignorm{\bigl([M_t(T)x](\lambda)\bigr)_{t> 0}}_{V^q}}_{p}\,\leq\,
C_{p,q}\,\norm{x}_p.
$$
\end{corollary}
\begin{proof}
Consider $x\in L^p(\Omega)$. According to \cite[Section VIII.7]{DS}, the function $t\mapsto [M_t(T)x](\lambda)$
is continuous for a.e. $\lambda\in\Omega$.
Let $t_0< t_1< \cdots < t_m$ be positive real numbers and let $\varepsilon >0$.
It follows from the strong continuity of $T=(T_t)_{t\geq 0}$ that there exist $\alpha>0$ and integers
$n_0, n_1,\ldots, n_m$ such that
$$
\forall\, k=0,\ldots,m,\qquad\bignorm{M_{t_k}(T)x\, - M_{n_k}(T_\alpha)x}_p\,<\varepsilon.
$$
Hence applying Theorem \ref{2Main} and a limit argument, we
deduce that
\begin{equation}\label{2Key}
\bignorm{\bigl(M_{t_0}(T)x,M_{t_1}(T)x,\ldots,M_{t_m}(T)x \bigr)}_{L^p(\Omega; v^q_m)}\,\leq C_{pq}\norm{x}_p.
\end{equation}
The result therefore follows from Lemma \ref{6Lem2}.
\end{proof}
An alternative proof of (\ref{2Key}) consists in using Fendler's dilation Theorem for semigroups (see \cite{Fe}),
and then arguing as in the proof Theorem \ref{2Main}. This only
requires knowing that the result of Corollary \ref{2Continuous} holds true for
the translation group on $L^p(\Rdb)$, which follows from \cite{B, JKRW, JR},
and using Remark \ref{2Amenable} for $G=\Rdb$ to transfer that result to strongly continuous groups
of contractively regular isometries.
\medskip
\section{The analytic case.}
Let $X$ be an arbitrary Banach space, and let
$(T_t)_{t\geq 0}$ be a strongly continuous semigroup on $X$.
We call it a bounded analytic semigroup if there exists a positive angle
$\omega\in\bigl(0,\frac{\pi}{2}\bigr)$ and a
bounded analytic family $z\in\Sigma_\omega\mapsto
T_z\in B(X)$ extending $(T_t)_{t>0}$,
where
$$
\Sigma_{\omega}\,=\,\bigl\{z\in \Cdb^*\, :\, \vert{\rm Arg}(z)\vert<\omega\bigr\}
$$
is the open sector of angle $2\omega$ around $(0,\infty)$. We refer to \cite{Go,Pa} for various
characterizations and properties of bounded analytic semigroups. We simply recall that
if we let $A$ denote the infinitesimal generator of $(T_t)_{t\geq 0}$, then the latter is
a bounded analytic semigroup if and only if
$T_t$ maps $X$ into the domain of $A$ for any $t>0$ and
there exist two constants $C_0,C_1>0$ such that
\begin{equation}\label{3Anal1}
\forall\, t>0,\qquad \norm{T_t}\leq C_0
\qquad\hbox{and}\qquad
\norm{tAT_t}\leq C_1.
\end{equation}
The definition of analyticity for discrete semigroups parallels (\ref{3Anal1}).
Let $T\in B(X)$. We say that $T$ is power bounded if
$$
\sup_{n\geq 0}\norm{T^n}\,<\infty
$$
and that it is analytic if moreover,
$$
\sup_{n\geq 1} n\norm{T^{n}-T^{n-1}}\,<\infty.
$$
This notion goes back to \cite{CSC} and has been studied in various contexts
so far. We gather here a few spectral properties
of these operators and refer to \cite{Bl2, Ly,NZ,N} for proofs and complements.
The most important result is the following:
an operator $T$ is power bounded and analytic if and only if
\begin{equation}\label{3Ritt}
\sigma(T)\subset \overline{\Ddb}\qquad\hbox{and}\qquad
\bigl\{(z-1)(zI-T)^{-1}\,:\,\vert z \vert>1\bigr\}\ \hbox{is bounded}.
\end{equation}
This property is called the `Ritt condition'.
The key argument for this characterization is due to O. Nevanlinna \cite{N}.
For any angle $\gamma\in\bigl(0,\frac{\pi}{2}\bigr)$, let
$B_\gamma$ be the convex hull of $1$ and the closed disc $\overline{D}(0,\sin\gamma)$.
\begin{figure}[ht]
\vspace*{2ex}
\begin{center}
\includegraphics[scale=0.4]{fig1.eps}
\begin{picture}(0,0)
\put(-2,65){{\footnotesize $1$}}
\put(-68,65){{\footnotesize $0$}}
\put(-32,81){{\footnotesize $\gamma$}}
\put(-55,85){{\small $B_\gamma$}}
\end{picture}
\end{center}
\caption{\label{f1}}
\end{figure}
\noindent
Then (\ref{3Ritt}) implies that
\begin{equation}\label{3Stolz1}
\exists\,\gamma\in\bigl(0,\frac{\pi}{2}\bigr)\ \big\vert\qquad
\sigma(T)\subset B_\gamma.
\end{equation}
Furthermore, (\ref{3Stolz1}) is equivalent to
\begin{equation}\label{3Stolz2}
\exists\,K>0\ \big\vert\ \ \forall\, z\in\sigma(T),\qquad \vert 1-z\vert\leq K\bigl(1-\vert z\vert\bigr).
\end{equation}
The aim of this section is to show that under an analyticity assumption, the ergodic averages
can be replaced by the semigroup itself in either Theorem \ref{2Main} (discrete case) or
Corollary \ref{2Continuous} (continuous case).
\bigskip
As in Section 3, we consider a measure space $(\Omega,\mu)$ and
a number $1<p<\infty$. We will consider an operator
$T\colon L^p(\Omega)\to L^p(\Omega)$ and we let
$$
\Delta_n^m= T^n(T-I)^m
$$
for any integers $n,m\geq 0$. Note that $(\Delta_n^m)_{n\geq 0}$ is the $m$-difference
sequence of $(T^n)_{n\geq 0}$.
We will need the following Littlewood-Paley type
inequalities which were estabished in \cite{LMX}.
\begin{proposition}\label{3SFE}
Let $T\colon L^p(\Omega)\to L^p(\Omega)$ be a contractively regular
operator, with $1<p<\infty$, and assume that $T$ is analytic.
Then for any integer $m\geq 0$, there is a constant $C_m>0$ such that
$$
\Bignorm{\Bigl(\sum_{n=0}^{\infty} (n+1)^{2m+1}\bigl\vert \Delta_{n}^{m+1}(x)
\bigr\vert^2\Bigr)^{\frac{1}{2}}}_p\,\leq\,C_m \,\norm{x}_p,\qquad x\in L^p(\Omega).
$$
\end{proposition}
We will also use the following elementary estimates, whose proofs are left to the reader.
\begin{lemma}\label{3Sum}
For any integer $m\geq 0$, there exists a constant $K_m$ such that for any $n\geq 1$,
$$
\Bigl(\sum_{j=n}^{2n} (j+1)^{1-2m}\Bigr)^{\frac{1}{2}}\,\leq\, K_m n^{-m+1}.
$$
\end{lemma}
\begin{lemma}\label{3Mult}
For any sequences $(\delta_n)_{n\geq 0}\in v^1$ and $(z_n)_{n\geq 0}\in L^p(\Omega;v^q)$, we have
$(\delta_n z_n)_{n\geq 0}\in L^p(\Omega;v^q)$ and
$$
\bignorm{(\delta_n z_n)_{n\geq 0}}_{L^p(v^q)}\,\leq 3 \,
\bignorm{(\delta_n)_{n\geq 0}}_{v^1}\,\bignorm{(z_n)_{n\geq 0}}_{L^p(v^q)}.
$$
\end{lemma}
\bigskip
In the next statements and their proofs, $\lesssim$ will stand for an inequality up to
a constant which may depend on $T$, $q$ and $m$, but not on $x$.
\begin{theorem}\label{3Main} Let $T\colon L^p(\Omega)\to L^p(\Omega)$ be a contractively regular
operator, with $1<p<\infty$, and assume that $T$ is analytic. Then for any $2<q<\infty$, we have an estimate
\begin{equation}\label{3Tn}
\bignorm{\bigl(T^n (x)\bigr)_{n\geq 0}}_{L^p(v^q)}\,\lesssim\,\norm{x}_p,\qquad x\in L^p(\Omega).
\end{equation}
More generally, for any integer $m\geq 0$,
we have an estimate
\begin{equation}\label{3Delta}
\bignorm{\bigl(n^m\Delta_n^m (x)\bigr)_{n\geq 1}}_{L^p(v^q)}\,\lesssim\,\norm{x}_p,\qquad x\in L^p(\Omega).
\end{equation}
\end{theorem}
\begin{proof}
It will be convenient to set
$$
\Delta_n^{-1}\,=\,nM_{n-1}(T)\,=\sum_{j=0}^{n-1} T^j \,,\qquad n\geq 1.
$$
Then for any $n<N$ and any $m\geq -1$, we have
\begin{equation}\label{31}
\Delta_{N}^{m} -\Delta_{n}^{m}\,=\,\sum_{j=n}^{N-1}\Delta_{j}^{m+1}.
\end{equation}
With the above notation, (\ref{3Delta}) holds true for $m=-1$, by Theorem \ref{2Main}.
We will proceed by induction. We fix an integer $m\geq 0$ and assume that
(\ref{3Delta}) holds true for $(m-1)$. Thus using Lemma \ref{3Mult}, we both have
\begin{equation}\label{3Induction1}
\bignorm{\bigl(n^{m-1}\Delta_{n+1}^{m-1}(x)\bigr)_{n\geq 1}}_{L^p(v^q)}\,\lesssim\,\norm{x}_p
\end{equation}
and
\begin{equation}\label{3Induction2}
\bignorm{\bigl(n^{m-1}\Delta_{2n+1}^{m-1} (x)\bigr)_{n\geq 1}}_{L^p(v^q)}\,\lesssim\,\norm{x}_p.
\end{equation}
Next for any $n\geq 1$, we write
\begin{align*}
\sum_{j=n}^{2n} (j+1)\Delta_{j}^{m+1}\, & =\,\sum_{j=n}^{2n}(j+1)(
\Delta_{j+1}^{m} - \Delta_{j}^{m})\\
& =\,\sum_{j=n+1}^{2n+1} j\Delta_{j}^{m}\, -\,\sum_{j=n}^{2n}(j+1)\Delta_{j}^{m}\\
& =\,-\sum_{j=n+1}^{2n} \Delta_{j}^{m}\, +\, (2n+1)\Delta_{2n+1}^{m}\,-\,(n+1) \Delta_{n}^{m}\\
& =\, n\Delta_{2n+1}^{m}\,+ (n+1)(\Delta_{2n+1}^{m} - \Delta_{n}^{m})\, +\, \Delta_{n+1}^{m-1}\, -\,\Delta_{2n+1}^{m-1},
\end{align*}
using (\ref{31}) in due places.
Hence
\begin{align}\label{32}
n^m\Delta_{2n+1}^{m}\,=\,n^{m-1}\sum_{j=n}^{2n} (j+1)\Delta_{j}^{m+1}& \,-\, n^{m-1}(n+1)
(\Delta_{2n+1}^{m} - \Delta_{n}^{m})\\ \notag & +\, \,n^{m-1}\Delta_{2n+1}^{m-1}\, -\, n^{m-1}\Delta_{n+1}^{m-1}.
\end{align}
This identity suggests the introduction of the following two sequences of operators. For any $n\geq 1$, we set
$$
A_n=\,n^{m-1}\sum_{j=n}^{2n} (j+1)\Delta_{j}^{m+1}\qquad\hbox{and}\qquad
B_n=\, n^m(\Delta_{2n+1}^{m} - \Delta_{n}^{m}).
$$
Also for any $x\in L^p(\Omega)$, we set
$$
\Phi_m(x)\,=\,\Bigl(\sum_{j=1}^{\infty}(j+1)^{2m+1}\bigl\vert \Delta_{j}^{m+1}(x)\vert^2\Bigr)^{\frac{1}{2}}.
$$
According to Proposition \ref{3SFE}, this function is an element of $L^p(\Omega)$.
Let $(n_k)_{k\geq 0}$ be an increasing sequence of integers, with $n_0=1$.
For any $k\geq 1$, we set
\begin{align*}
a_k & = \left\{
\begin{array}{ll}\displaystyle
n_k^{m-1}\,\sum_{j=2n_{k-1}+1}^{2n_k} (j+1)\Delta_{j}^{m+1}\qquad & \mbox{if } 2n_{k-1}\geq n_k \\
\displaystyle
n_k^{m-1}\,\sum_{j=n_k}^{2n_k} (j+1)\Delta_{j}^{m+1} \qquad & \mbox{if } 2n_{k-1} < n_k,
\end{array}
\right.\\
\smallskip
b_k &= \left\{
\begin{array}{ll}\displaystyle
-n_{k-1}^{m-1}\,\sum_{j=n_{k-1}}^{n_k -1} (j+1)\Delta_{j}^{m+1} \qquad \ & \mbox{if } 2n_{k-1}\geq n_k \\
\displaystyle
-n_{k-1}^{m-1}\,\sum_{j=n_{k-1}}^{2n_{k -1}} (j+1)\Delta_{j}^{m+1} \qquad \ & \mbox{if } 2n_{k-1} < n_k,
\end{array}
\right.\\
\bigskip
c_k & = \left\{
\begin{array}{ll}\displaystyle
\Bigl(n_k^{m-1} -n_{k-1}^{m-1}\Bigr)\,\sum_{j=n_{k}}^{2n_{k-1}}
(j+1)\Delta_{j}^{m+1}\ & \mbox{if } 2n_{k-1}\geq n_k \\
\quad 0 & \mbox{if } 2n_{k-1} < n_k.
\end{array}
\right.
\end{align*}
This yields a decomposition
\begin{equation}\label{3Decomp}
A_{n_k} -A_{n_{k-1}}\,=\,a_k+ b_k+ c_k.
\end{equation}
Let $x\in L^p(\Omega)$. If $2n_{k-1}\geq n_k$,
we have, using Cauchy-Schwarz,
\begin{align*}
\bigl\vert a_k(x)\bigr\vert\,& \leq\,n_k^{m-1}
\sum_{j=2n_{k-1}+1}^{2n_k} (j+1)\bigl\vert \Delta_{j}^{m+1}(x)\bigr\vert\\
& \leq\, n_k^{m-1}\Bigl(\sum_{j=2n_{k-1}+1}^{2n_k} (j+1)^{1-2m}\Bigr)^{\frac{1}{2}}\,
\Bigl(\sum_{j=2n_{k-1}+1}^{2n_k} (j+1)^{2m+1}\bigl\vert \Delta_{j}^{m+1}(x)\bigr\vert^2
\Bigr)^{\frac{1}{2}}\\
& \leq\, n_k^{m-1}\Bigl(\sum_{j=n_{k}}^{2n_k} (j+1)^{1-2m}\Bigr)^{\frac{1}{2}}\,
\Bigl(\sum_{j=2n_{k-1}+1}^{2n_k} (j+1)^{2m+1}\bigl\vert \Delta_{j}^{m+1}(x)\bigr\vert^2
\Bigr)^{\frac{1}{2}}.
\end{align*}
Similarly if $2n_{k-1}< n_k$, we have
\begin{align*}
\bigl\vert a_k(x)\bigr\vert\,&
\leq\, n_k^{m-1}\Bigl(\sum_{j=n_{k}}^{2n_k} (j+1)^{1-2m}\Bigr)^{\frac{1}{2}}\,
\Bigl(\sum_{j=n_{k}}^{2n_k} (j+1)^{2m+1}\bigl\vert \Delta_{j}^{m+1}(x)\bigr\vert^2
\Bigr)^{\frac{1}{2}}\\
& \leq\, n_k^{m-1}\Bigl(\sum_{j=n_{k}}^{2n_k} (j+1)^{1-2m}\Bigr)^{\frac{1}{2}}\,
\Bigl(\sum_{j=2n_{k-1}+1}^{2n_k} (j+1)^{2m+1}\bigl\vert \Delta_{j}^{m+1}(x)\bigr\vert^2
\Bigr)^{\frac{1}{2}}.
\end{align*}
Hence in both cases,
we have
$$
\bigl\vert a_k(x)\bigr\vert^2\,\leq\,
K_m^2 \sum_{j=2n_{k-1}+1}^{2n_k} (j+1)^{2m+1}\bigl\vert \Delta_{j}^{m+1}(x)\bigr\vert^2
\,,
$$
by Lemma \ref{3Sum}. Summing up, we deduce that
\begin{equation}\label{3Alpha}
\sum_{k=1}^{\infty}\bigl\vert a_k(x)\bigr\vert^2\,\leq\,
K_m^2 \,\Phi_m(x)^2.
\end{equation}
Likewise we have
\begin{equation}\label{3Beta}
\sum_{k=1}^{\infty}\bigl\vert b_k(x)\bigr\vert^2\,\leq\,
K_m^2 \,\Phi_m(x)^2.
\end{equation}
We now turn to $c_k(x)$. Assume that $2n_{k-1}\geq n_k$. Then using again
Cauchy-Schwarz and Lemma \ref{3Sum}, we have
$$
\bigl\vert c_k(x)\bigr\vert\,
\leq\, \bigl\vert n_k^{m-1} -n_{k-1}^{m-1}\bigr\vert\,\Bigl(\sum_{j=n_{k}}^{2n_{k-1}}
(j+1)^{1-2m}\Bigr)^{\frac{1}{2}}\,\Bigl(\sum_{j=n_{k}}^{2n_{k-1}}
(j+1)^{2m+1}\bigl\vert \Delta_{j}^{m+1}(x)\bigr\vert^2
\Bigr)^{\frac{1}{2}},
$$
hence
$$
\bigl\vert c_k(x)\bigr\vert^2\,\leq\, K_m^2\,\biggl(
\frac{n_k^{m-1} -n_{k-1}^{m-1}}{n_k^{m-1}}\biggr)^2\
\sum_{j=n_{k}}^{2n_{k-1}}
(j+1)^{2m+1}\bigl\vert \Delta_{j}^{m+1}(x)\bigr\vert^2.
$$
For any integer $j\geq 1$, define
$$
J_j=\{k\geq 1\, :\, n_{k}\leq j\leq 2n_{k-1}\},
$$
and set
$$
\Lambda_j\,=\,\sum_{k\in J_j}\biggl(
\frac{n_k^{m-1} -n_{k-1}^{m-1}}{n_k^{m-1}}\biggr)^2\,.
$$
Then it follows from the above calculation that
$$
\sum_{k=1}^{\infty}\bigl\vert c_k(x)\bigr\vert^2\,\leq\,
K_m^2\sum_{j=1}^{\infty} \Lambda_j\,(j+1)^{2m+1}\bigl\vert \Delta_{j}^{m+1}(x)\bigr\vert^2\,.
$$
Let us now estimate the $\Lambda_j$'s.
Observe that if $J_j$ is a non empty set, then it is a finite interval of integers. Thus it reads as
$$
J_j=\{k_j-N+1, k_j-N+2,\ldots, k_j-1, k_j\},
$$
where $k_j$ is the biggest element of $J_j$ and $N$ is its cardinal.
Suppose that $m\geq 2$, so that
the sequence $(n_k^{m-1})_k$ is increasing.
Then
$$
\sum_{k\in J_j}
n_k^{m-1} -n_{k-1}^{m-1}\, = \,\sum_{r=0}^{N-1} n_{k_j-r}^{m-1} - n_{k_j -r-1}^{m-1}\,
= \, n_{k_j}^{m-1} - n_{k_j -N}^{m-1}\,\leq n_{k_j}^{m-1}.
$$
Since $k_j\in J_j$, we have $n_{k_j}\leq j$, hence
$$
\sum_{k\in J_j}
n_k^{m-1} -n_{k-1}^{m-1}\,\leq j^{m-1}.
$$
On the other hand, we have $j\leq 2n_k$ for any $k\in J_j$, hence
$$
\sum_{k\in J_j}
\frac{n_k^{m-1} -n_{k-1}^{m-1}}{n_k^{m-1}}\, \leq \, \Bigl(\frac{2}{j}\Bigr)^{m-1}\,
\sum_{k\in J_j}
n_k^{m-1} -n_{k-1}^{m-1}\,
\leq \, 2^{m-1}.
$$
We immediatly deduce that
$$
\Lambda_j\leq 4^{m-1}.
$$
In the case when $m=0$, we have similarly
$$
\sum_{k\in J_j}
\frac{n_{k-1}^{-1} -n_{k}^{-1}}{n_k^{-1}}\, \leq \,j\,
\sum_{k\in J_j} n_{k-1}^{-1} -n_{k}^{-1}\\
\leq \,j\,
n_{k_j-N}^{-1}\leq 2.
$$
Hence $\Lambda_j\leq 4$ in this case. Lastly, it is plain that if $m=1$, we have $\Lambda_j=0$.
This shows that in all cases, we have an estimate
$$
\sum_{k=1}^{\infty}\bigl\vert c_k(x)\bigr\vert^2\,\leq\,
{K'_m}^2\,\Phi_m(x)^2.
$$
Now recall (\ref{3Decomp}). Combining the above estimate with (\ref{3Alpha}) and (\ref{3Beta}), we obtain
that
\begin{align*}
\sum_{k=1}^{\infty}
\bigl\vert A_{n_k}(x) -A_{n_{k-1}}(x)\bigr\vert^2\,\leq &\, 3\Bigl(
\sum_{k=1}^{\infty} \bigl\vert a_k(x)\bigr\vert^2\, +\,
\sum_{k=1}^{\infty} \bigl\vert b_k(x)\bigr\vert^2\, +\,
\sum_{k=1}^{\infty} \bigl\vert c_k(x)\bigr\vert^2\Bigr)\\
&\leq \bigl(6K_m^2+{K'_m}^2\bigr)\,\Phi_m(x)^2.
\end{align*}
Since the upper bound does not depend on the sequence $(n_k)_{k\geq 0}$, this estimate and
Proposition \ref{3SFE} imply that the sequence
$(A_n(x))_{n\geq 0}$ belongs to $L^p(\Omega;v^2)$, and that we have an estimate
\begin{equation}\label{3A}
\bignorm{\bigl(A_n (x)\bigr)_{n\geq 1}}_{L^p(v^2)}\,\lesssim\,\norm{x}_p,\qquad x\in L^p(\Omega).
\end{equation}
We will now apply a similar treatment to the sequence $(B_n)_n$. According to (\ref{31}), we can write
$$
B_n=n^m\,\sum_{j=n}^{2n} \Delta_{j}^{m+1}.
$$
For any $k\geq 1$, we set
\begin{align*}
\alpha_k & = n_k^{m}\,\sum_{j=2n_{k-1}+1}^{2n_k} \Delta_{j}^{m+1},\\
\beta_k & = -n_{k-1}^{m}\,\sum_{j=n_{k-1}}^{n_k -1} \Delta_{j}^{m+1},\\
\gamma_k & = \Bigl(n_k^{m} -n_{k-1}^{m}\Bigr)\,\sum_{j=n_{k}}^{2n_{k-1}}
\Delta_{j}^{m+1}
\end{align*}
if $2n_{k-1}\geq n_k$, and
\begin{align*}
\alpha_k & = n_k^{m}\,\sum_{j=n_k}^{2n_k} \Delta_{j}^{m+1},\\
\beta_k & = -n_{k-1}^{m}\,\sum_{j=n_{k-1}}^{2n_{k -1}} \Delta_{j}^{m+1},\\
\gamma_k & = 0
\end{align*}
if $2n_{k-1} < n_k$.
Arguing as above, we obtain that for any $x\in L^p(\Omega)$, we have
$$
\bigl\vert\alpha_k(x)\bigr\vert\,\leq \,
n_k^{m}\Bigl(\sum_{j=n_{k}}^{2n_k} (j+1)^{-1-2m}\Bigr)^{\frac{1}{2}}\,
\Bigl(\sum_{j=2n_{k-1}+1}^{2n_k} (j+1)^{2m+1}\bigl\vert \Delta_{j}^{m+1}(x)\bigr\vert^2
\Bigr)^{\frac{1}{2}},
$$
and then
$$
\sum_{k=1}^{\infty}\bigl\vert \alpha_k(x)\bigr\vert^2\,\leq\,
K_{m+1}^2 \,\Phi_m(x)^2.
$$
Likewise we have
$$
\sum_{k=1}^{\infty}\bigl\vert \beta_k(x)\bigr\vert^2\,\leq\,
K_{m+1}^2 \,\Phi_m(x)^2,
$$
as well as an estimate
$$
\sum_{k=1}^{\infty}\bigl\vert \gamma_k(x)\bigr\vert^2\,\leq\,
K^{'2}_{m+1}\,\Phi_m(x)^2.
$$
These three inequalities imply that the sequence
$(B_n(x))_{n\geq 0}$ belongs to $L^p(\Omega;v^2)$, and that we have an estimate
\begin{equation}\label{3B}
\bignorm{\bigl(B_n (x)\bigr)_{n\geq 1}}_{L^p(v^2)}\,\lesssim\,\norm{x}_p,\qquad x\in L^p(\Omega).
\end{equation}
We can now conclude our proof. Recall that $q>2$, so that $v^2\subset v^q$.
Then using (\ref{3Induction1}), (\ref{3Induction2}), (\ref{3A}), (\ref{3B})
and Lemma \ref{3Mult}, it follows from the decomposition formula (\ref{32}) that for any $x\in L^p(\Omega)$,
$\bigl(n^m\Delta_{2n+1}^m(x)\bigr)_{n\geq 1}$ belongs to $L^p(\Omega;v^q)$ and that we have an estimate
$$
\bignorm{\bigl(n^m\Delta_{2n+1}^m(x)\bigr)_{n\geq 1}}_{L^p(v^q)}\,\lesssim\,\norm{x}_p.
$$
Since $n\Delta_{n}^{m}= n^{m}\Delta_{2n+1}^{m} - B_{n}$, a second application of (\ref{3B}) yields (\ref{3Delta}).
\end{proof}
The following is an analog of Theorem \ref{3Main} for continuous semigroups.
\begin{corollary}\label{3Continuous}
Let $(T_t)_{t\geq 0}$ be a bounded analytic semigroup on $L^p(\Omega)$
and assume that $T_t\colon L^p(\Omega)\to L^p(\Omega)$ is contractively regular for any
$t\geq 0$. Let $2<q<\infty$ and let $x\in L^p(\Omega)$.
Then for a.e. $\lambda\in\Omega$, the family
$\bigl([T_t (x)](\lambda)\bigr)_{t>0}$ belongs to $V^q$ and we have an estimate
\begin{equation}\label{3Cont1}
\Bignorm{\lambda\mapsto \bignorm{\bigl([T_t(x)](\lambda)\bigr)_{t> 0}}_{V^q}}_{p}\,\lesssim\,\norm{x}_p.
\end{equation}
More generally, for any
integer $m\geq 0$, the family
$\Bigl(t^m \frac{\partial^m}{\partial t^m}\bigl(T_t(x)\bigr)(\lambda)\Bigr)_{t>0}$
belongs to $V^q$ for a.e. $\lambda\in\Omega$ and we have an estimate
\begin{equation}\label{3Cont2}
\Bignorm{\bignorm{\lambda \mapsto \Bigl(t^m
\frac{\partial^m}{\partial t^m}\bigl(T_t(x)\bigr)(\lambda)\Bigr)_{t>0}}_{V^q}}_p
\,\lesssim\,\norm{x}_p,\qquad x\in L^p(\Omega).
\end{equation}
\end{corollary}
\begin{proof} Let $m\geq 0$ be an integer. It follows from \cite[Lemma, p. 72]{S2}
that for any $x\in L^p(\Omega)$, the function
$$
t\mapsto t^m \frac{\partial^m}{\partial t^m}\bigl( T_t(x)\bigr)(\lambda)
$$
is continuous for a.e. $\lambda\in \Omega$.
Next it follows from the proof of \cite[Cor. 4.2]{LMX} that
there exists a constant $C_m>0$ such that
$$
\Bignorm{\Bigl(\sum_{n=0}^{\infty} (n+1)^{2m+1}\bigl\vert T_t^n(T_t-I)^m(x)
\bigr\vert^2\Bigr)^{\frac{1}{2}}}_p\,\leq\,C_m \,\norm{x}_p
$$
for any $t>0$ and any $x\in L^p(\Omega)$. That is, the operators $T_t$ satisfy Proposition
\ref{3SFE} uniformly. Since they also satisfy Theorem \ref{2Main} uniformly,
it follows from the proof
of Theorem \ref{3Main} that they satisfy the estimate (\ref{3Delta}) uniformly.
Hence using an approximation argument as in the proof of \cite[Cor. 4.2]{LMX}, we deduce that
for any $x\in L^p(\Omega)$ and for any $0<t_0<t_1<\cdots<t_m$, we have
$$
\bignorm{\bigl(T_{t_0}(x),T_{t_1}(x),\ldots,T_{t_m}(x) \bigr)}_{L^p(\Omega; v^q_m)}\,\leq C.
$$
The result therefore follows from Lemma \ref{6Lem2}.
\end{proof}
\medskip
\section{Additional properties.}
We give here further properties of contractively regular operators
and contractively regular semigroups, in connection with variational inequalities.
Let $(\Omega,\mu)$ be a measure space and let $1<p<\infty$.
\medskip
{\it 5.1. Individual ergodic theorems}
\smallskip
Let $T\colon L^p(\Omega)\to L^p(\Omega)$ be a contraction. According to the
Mean Ergodic Theorem, we have a direct sum decomposition
$$
L^p(\Omega)\,=\,N(I-T)\oplus\overline{R(I-T)},
$$
where $N(\cdotp)$ and $R(\cdotp)$ denote the kernel and the range, respectively.
Moreover if we let $P_T\colon L^p(\Omega)\to L^p(\Omega)$ denote the
corresponding projection onto $N(I-T)$, then
\begin{equation}\label{7Mean}
M_n(T)x\,\stackrel{L^p}{\longrightarrow}\, P_T(x)
\end{equation}
for any $x\in L^p(\Omega)$. It is well-known that if $T$ is an absolute
contraction, then $M_n(T)x\to P_T(x)$ almost everywhere (see e.g. \cite[Section VIII.6]{DS}).
We extend this classical result, as follows.
\begin{corollary}
Assume that $T\colon L^p(\Omega)\to L^p(\Omega)$ is contractively regular.
Then for any $x\in L^p(\Omega)$,
$$
[M_n(T)x](\lambda)\,\longrightarrow\, [P_T(x)](\lambda)\qquad \hbox{for a.e.}\ \lambda\in\Omega.
$$
\end{corollary}
\begin{proof} Let $x\in L^p(\Omega)$ and let $2<q<\infty$. According to Theorem \ref{2Main}, the sequence
$\bigl([M_n(T)x](\lambda)\bigr)_{n\geq 0}$ belongs to $v^q$ for almost every $\lambda\in\Omega$.
Hence $\bigl([M_n(T)x](\lambda)\bigr)_{n\geq 0}$ converges for almost every $\lambda\in\Omega$.
Combining with (\ref{7Mean}), we obtain the result.
\end{proof}
If a contraction $T\colon L^p(\Omega)\to L^p(\Omega)$ is analytic, then
$\norm{T^n-T^{n+1}}\to 0$, hence $T^n(x)\to 0$ for any $x\in R(I-T)$.
Consequently,
$$
T^n(x)\,\stackrel{L^p}{\longrightarrow}\, P_T(x)
$$
for any $x\in L^p(\Omega)$. Using Theorem \ref{3Main} and arguing as above, we obtain the following.
\begin{corollary}\label{72}
Let $T\colon L^p(\Omega)\to L^p(\Omega)$ be a contractively regular operator and assume that $T$ is analytic.
Then for any $x\in L^p(\Omega)$,
$$
[T^n(x)](\lambda)\,\longrightarrow\, [P_T(x)](\lambda)\qquad \hbox{for a.e.}\ \lambda\in\Omega.
$$
\end{corollary}
\bigskip
We now consider the continuous case. The situation is essentially similar, except that
we can also consider the behaviour when the parameter $t$ tends to $0^+$.
Let $T=(T_t)_{t\geq 0}$ be a strongly continuous semigroup
of contractions. By definition, for any $x\in L^p(\Omega)$, $T_t(x)\to x$ in the $L^p$-norm
when $t\to 0^+$.
This implies that $M_t(T)x\to x$ when $t\to 0^+$.
Let $A$ denote the infinitesimal generator of $T=(T_t)_{t\geq 0}$. As in the discrete case,
we have a direct sum decomposition
$$
L^p(\Omega)\,=N(A)\oplus \overline{R(A)}.
$$
Moreover if we let $P_A\colon L^p(\Omega)\to L^p(\Omega)$ denote the
corresponding projection onto $N(A)$, then
$$
M_t(T)x\,\stackrel{L^p}{\longrightarrow}\, P_A(x)\quad\hbox{when}\ t\to\infty
$$
for any $x\in L^p(\Omega)$. If further $(T_t)_{t\geq 0}$ is a bounded analytic semigroup,
then
$$
T_t(x)\,\stackrel{L^p}{\longrightarrow}\, P_A(x)
$$
for any $x\in L^p(\Omega)$.
Now applying Corollaries \ref{2Continuous} and \ref{3Continuous}, we deduce
the following individual ergodic theorems.
\begin{corollary}\label{74}
Let $T=(T_t)_{t\geq 0}$ be a strongly continuous semigroup of contractively
regular operators on $L^p(\Omega)$, and
let $x\in L^p(\Omega)$. Then for almost every $\lambda\in\Omega$,
$$
[M_t(T)x](\lambda)\,\longrightarrow\, [P_A(x)](\lambda)\quad\hbox{when}\ t \to\infty
$$
and
$$
[M_t(T)x](\lambda)\,\longrightarrow\, x(\lambda)\quad\hbox{when}\ t\, \to 0^+.
$$
\end{corollary}
\begin{corollary}\label{74}
Let $T=(T_t)_{t\geq 0}$ be a bounded analytic semigroup and
assume that $T_t$ is contractively
regular for any $t\geq 0$. Let $x\in L^p(\Omega)$.
Then for almost every $\lambda\in\Omega$,
$$
[T_t(x)](\lambda)\,\longrightarrow\, [P_A(x)](\lambda)\quad\hbox{when}\ t\to\infty
$$
and
$$
[T_t(x)](\lambda)\,\longrightarrow\, x(\lambda)\quad\hbox{when}\ t\to 0^+.
$$
\end{corollary}
\bigskip
{\it 5.2. The case $q=2$}
\smallskip
In this section,
we fix a increasing sequence $(n_k)_{k\geq 0}$ of integers, with $n_0=0$. Given any sequence
$(a_n)_{n\geq 0}$ of complex numbers, we define the so-called oscillation norm
$$
\norm{(a_n)_{n\geq 0}}_{o^2}\,=\, \bigl(\vert a_0\vert^2\, +\,\sum_{k\geq 0}\,\max_{n_{k}\leq n,m\leq n_{k+1}}
\vert a_{n}-a_{m}\vert^2 \bigr)^{\frac{1}{2}},
$$
and we let $o^2$ denote the Banach space of all sequences with a finite oscillation norm, equipped with
$\norm{\ }_{o^2}$. This space (whose definition depends on the sequence $(n_k)_{k\geq 0}$) was used
in \cite{B,JKRW,JR} as a substitute to $v^q$ in the case $q=2$ (see also \cite{Ga}). Indeed,
neither Theorem \ref{2Main} nor Theorem \ref{3Main} holds true for $q=2$, see \cite{JKRW}
and \cite[Section 8]{JW}.
Recall the shift operator $s_p\colon \ell^p_{\footnotesize{\Zdb}}\to \ell^p_{\footnotesize{\Zdb}}$
for any $1<p<\infty$ (see Section 3).
According to \cite[Thm. A]{JKRW}, there is a constant $C_{p,2}$ such that
$$
\bignorm{\bigl(M_n(s_p)c\bigr)_{n\geq 0}}_{L^p(o^2)}\,\leq\,C_{p,2}\,\norm{c}_p
$$
for any $c\in \ell^p_{\footnotesize{\Zdb}}$.
Hence arguing as in the proof of Theorem \ref{2Main}, we obtain the following $o^2$-version of the latter
statement.
\begin{theorem}\label{41} Let $T\colon L^p(\Omega)\to L^p(\Omega)$ be a contractively regular operator,
with $1<p<\infty$. Then we have
$$
\bignorm{\bigl(M_n(T)x\bigr)_{n\geq 0}}_{L^p(o^2)}\,\leq\,C_{p,2}\,\norm{x}_p,\qquad x\in L^p(\Omega).
$$
\end{theorem}
We also have an $o^2$-version of Theorem \ref{3Main}, as follows.
\begin{theorem}\label{42}
Let $T\colon L^p(\Omega)\to L^p(\Omega)$ be a contractively regular
operator, with $1<p<\infty$, and assume that $T$ is analytic. Then we have an estimate
\begin{equation}\label{4Tn}
\bignorm{\bigl(T^n (x)\bigr)_{n\geq 0}}_{L^p(o^2)}\,\lesssim\,\norm{x}_p,\qquad x\in L^p(\Omega).
\end{equation}
More generally, for any integer $m\geq 0$,
we have an estimate
\begin{equation}\label{4Delta}
\bignorm{\bigl(n^m\Delta_n^m (x)\bigr)_{n\geq 1}}_{L^p(o^2)}\,\lesssim\,\norm{x}_p,\qquad x\in L^p(\Omega).
\end{equation}
\end{theorem}
\begin{proof}
The proof is a variant of the one written for Theorem \ref{3Main}, let us explain this briefly.
We use the notation from Section 4. For any $m\geq -1$, consider the following three properties:
\begin{itemize}
\item [(i)$_{m}$] $\ \bignorm{\bigl(n^{m}\Delta_n^{m}(x)\bigr)_{n\geq 1}}_{L^p(o^2)}\,\lesssim\,\norm{x}_p$;
\smallskip
\item [(ii)$_{m}$] $\ \bignorm{\bigl(n^{m}\Delta_{n+1}^{m} (x)\bigr)_{n\geq 1}}_{L^p(o^2)}\,\lesssim\,\norm{x}_p$;
\smallskip
\item [(iii)$_{m}$] $\ \bignorm{\bigl(n^{m}\Delta_{2n+1}^{m} (x)\bigr)_{n\geq 1}}_{L^p(o^2)}\,\lesssim\,\norm{x}_p$.
\end{itemize}
Property (i)$_{m}$ is the result we wish to prove. Our strategy is to
show, by induction, that these three estimates hold true.
First, it is easy to deduce from Proposition \ref{3SFE} that (i)$_{m}$ and (ii)$_{m}$ are equivalent.
Second it follows from (\ref{3A}), (\ref{3B}) and (\ref{32}) that
(ii)$_{m-1}$ and (iii)$_{m-1}$ imply (iii)$_{m}$ and (i)$_{m}$. Indeed $v^2\subset o^2$
and $n^m\Delta_n^m = n^m\Delta_{2n+1}^m -B_n$. Hence it suffices to show
(i)$_{-1}$ and (iii)$_{-1}$. This is obtained by applying Theorem \ref{41} twice, the first time for
the $o^2$-space associated with the sequence $(n_k)_{k\geq 1}$, the second time for
$o^2$-space associated with the sequence $(2n_k+1)_{k\geq 1}$.
\end{proof}
There are also $o^2$-versions of Corollary \ref{2Continuous} and
Corollary \ref{3Continuous}, whose statements are left to the reader.
\bigskip
{\it 5.3. Jump functions}
\smallskip
It is well known that variational inequalities for a sequence of operators have consequences
in terms of jump functions. For any $\tau>0$ and any sequence
$a= (a_n)_{n\geq 0}$ of complex numbers, let $N(a,\tau)$ denote
the number of $\tau$-jumps of $a$, defined as
the supremum of all integers $N\geq 0$ for which there exist integers
$$
0\leq n_1 <m_1\leq n_2 < m_2\leq \cdots\leq n_N < m_N,
$$
such that $\vert a_{m_k}-a_{n_k}\vert >\tau$ for each $k=1,\ldots,N$.
It is clear that for any $1\leq q<\infty$,
$$
\tau^q N(a,\tau)\,\leq\,\norm{a}_{v^q}^q.
$$
Combining with Theorem \ref{2Main} and Theorem \ref{3Main}, we immediately
obtain (as in \cite[Thm. 3.15]{JR}) the following jump estimates.
\begin{corollary}\label{3Jumps}
Consider $1<p<\infty$ and $2<q<\infty$. Let $T\colon L^p(\Omega,\mu)\to L^p(\Omega,\mu)$ be a contractively
regular operator.
\begin{itemize}
\item [(1)] We have an estimate
$$
\Bignorm{\lambda\mapsto N\Bigl(\bigl([M_n(T)x](\lambda)\bigr)_{n\geq 0}\, ,\, \tau\Bigr)^{\frac{1}{q}}}_p\,\lesssim \, \frac{\norm{x}_p}{\tau}\,,
$$
and for any $K>0$, we also have
$$
\mu\Bigl\{\lambda\in\Omega\,\Big\vert\,
N\Bigl(\bigl([M_n(T)x](\lambda)\bigr)_{n\geq 0}\, ,\, \tau\Bigr)\,>\,
K\Bigr\}\,\lesssim\,
\frac{\norm{x}_p^p}{\tau^p K^{\frac{p}{q}}}\,,
$$
\item [(2)] Assume moreover that $T$ is analytic. Then we have similar estimates
$$
\Bignorm{\lambda\mapsto N\Bigl(\bigl([T^n (x)](\lambda)\bigr)_{n\geq 0}
\, ,\, \tau\Bigr)^{\frac{1}{q}}}_p\,\lesssim \, \frac{\norm{x}_p}{\tau}\,,
$$
and
$$
\mu\Bigl\{\lambda\in\Omega\,\Big\vert\,
N\Bigl(\bigl([ T^n (x)](\lambda)\bigr)_{n\geq 0}\, ,\, \tau\Bigr)\,>\,
K\Bigr\}\,\lesssim\,
\frac{\norm{x}_p^p}{\tau^p K^{\frac{p}{q}}}\,,
$$
Furthermore for any integer $m\geq 0$, similar results hold with
$n^m\Delta_n^m$ instead of $T^n$.
\end{itemize}
\end{corollary}
Similar results for the continuous case can be deduced as well from Corollary \ref{2Continuous} and
Corollary \ref{3Continuous}.
\medskip
\section{Examples and applications.}
We will now exhibit various classes of operators or semigroups
to which our results from Section 4 and Section 5 apply. We
focus on statements involving $q$-variation, although statements involving
the oscillation norm from the subsection 5.2 are also possible.
We start with the continuous case.
Let $(T_t)_{t\geq 0}$ be a strongly continuous semigroup on $L^2(\Omega)$ and assume that each
$T_t$ is an absolute contraction, that is,
\begin{equation}\label{3Diff1}
\norm{T_t(x)}_1\leq\norm{x}_1\qquad\hbox{and}\qquad
\norm{T_t(x)}_\infty\leq\norm{x}_\infty
\end{equation}
for any $x\in L^1(\Omega)+L^\infty(\Omega)$ and any $t>0$
(see Section 3). Thus for any $1<p<\infty$,
$(T_t)_{t\geq 0}$ extends to a strongly continuous semigroup of contractively regular operators.
It is well-known that $(T_t)_{t\geq 0}$ is a bounded analytic semigroup on $L^p(\Omega)$
for every $1<p<\infty$ if (and only if) it is a bounded analytic semigroup on $L^p(\Omega)$
for one $1<p<\infty$. Applying Corollary \ref{3Continuous} and Corollary \ref{74}, we derive the following.
\begin{corollary}\label{3Diff2}
Let $(T_t)_{t\geq 0}$ be a bounded analytic semigroup on $L^2(\Omega)$ satisfying
(\ref{3Diff1}). Then it
satisfies estimates (\ref{3Cont1}) and (\ref{3Cont2}) for every $1<p<\infty$ and every $2<q<\infty$.
Moreover for any $x\in L^p(\Omega)$, $T_t(x)$ converges almost everywhere when $t\to 0^+$ and when
$t\to\infty$.
\end{corollary}
Note that if each $T_t\colon L^2(\Omega)\to L^p(\Omega)$ is selfadjoint,
then $(T_t)_{t\geq 0}$ is a bounded analytic semigroup on $L^2(\Omega)$. Hence
the above corollary applies to symmetric diffusion semigroups and extends \cite[Thm. 3.3]{JR}.
We now consider the so-called subordinated semigroups. Let $1<p<\infty$ and let
$(T_t)_{t\geq 0}$ be a strongly continuous bounded semigroup on $L^p(\Omega)$. Let
$A$ denote its infinitesimal generator and let $0<\alpha<1$. Then $-(-A)^\alpha$ generates a
bounded analytic semigroup $(T_{\alpha,t})_{t\geq 0}$ on $L^p(\Omega)$, and for any $t>0$,
there exists a continuous function $f_{\alpha,t}\colon(0,\infty)\to\Rdb\,$ such that
\begin{equation}\label{3Subord1}
\forall\, s>0,\quad f_{\alpha,t}(s)\geq 0; \qquad \int_{0}^{\infty} f_{\alpha,t}(s)\, ds\, =1;
\end{equation}
and
\begin{equation}\label{3Subord2}
T_{\alpha,t}(x)\,=\, \int_{0}^{\infty} f_{\alpha,t}(s)\, T_s(x)\, ds \, ,\qquad
x\in L^p(\Omega).
\end{equation}
See e.g. \cite[IX.11]{Y} for details and complements.
\begin{corollary}\label{3Subord}
Let $(T_t)_{t\geq 0}$ be a strongly continuous semigroup on $L^p(\Omega)$, with $1<p<\infty$,
and assume that $T_t\colon L^p(\Omega)\to L^p(\Omega)$ is contractively regular for any
$t\geq 0$. Then for any $0<\alpha<\infty$ and any $2<q<\infty$, we have an estimate
$$
\Bignorm{\lambda\mapsto \bignorm{\bigl([T_{\alpha,t}(x)](\lambda)\bigr)_{t> 0}}_{V^q}}_{p}\,\lesssim\,\norm{x}_p
$$
for the subordinated semigroup $T_{\alpha,t}=e^{-t(-A)^\alpha}$. Moreover for any $x\in L^p(\Omega)$,
$T_{\alpha,t}(x)$ converges almost everywhere when $t\to 0^+$ and when
$t\to\infty$.
\end{corollary}
\begin{proof}
Let $0<\alpha<\infty$. It follows from (\ref{3Subord1}) and (\ref{3Subord2}) that
for any $t>0$,
$$
\norm{T_{\alpha,t}}_r\leq \int_{0}^{\infty} f_{\alpha,t}(s)\,\norm{T_s}_r\, ds\,\leq 1.
$$
Hence the result follows from Corollary \ref{3Continuous} and Corollary \ref{74}.
\end{proof}
We now turn to the discrete case and consider normal operators on $L^2$.
\begin{lemma}\label{3Normal1}
Let $H$ be a Hilbert space and let $T\in B(H)$ be a normal operator. Then
$T$ is an analytic power bounded operator if and only if it satisfies
(\ref{3Stolz2}).
\end{lemma}
\begin{proof} The `only if' part holds for any operator, as discussed at the beginning of Section 4.
Conversely, assume that a normal operator $T\colon H\to H$ satisfies
(\ref{3Stolz2}).
Applying the Spectral Theorem, we deduce that $T$ is a contraction and that
for some constant $K>0$, we have
\begin{align*}
n\norm{T^n-T^{n-1}}\, & =\,\sup_{z\in\sigma(T)} n\vert z^n -z^{n-1}\vert
\, =\,\sup_{z\in\sigma(T)} n\vert z\vert^{n-1}\,\vert 1-z\vert\,\\ & \leq\,
K\sup_{r\in[0,1]} n(r^{n-1} -r^n)\,=\,K \Bigl(\frac{n-1}{n}\Bigr)^{n-1}
\end{align*}
for any $n\geq 1$. Hence the sequence $\bigl(n(T^n-T^{n-1})\bigr)_{n\geq 1}$ is bounded.
\end{proof}
The following is a straightforward consequence of the above lemma, Theorem \ref{3Main} and Corollary \ref{72}.
\begin{corollary}\label{3Normal2} Let
$T\colon L^2(\Omega)\to L^2(\Omega)$ be a contractively regular normal operator satisfying
(\ref{3Stolz2}). Then it satisfies an estimate
$$
\bignorm{\bigl(T^n (x)\bigr)_{n\geq 0}}_{L^2(v^q)}\,\lesssim\,\norm{x}_2,\qquad x\in L^2(\Omega)
$$
for any $2<q<\infty$. Moreover for any $x\in L^2(\Omega)$, $T^n(x)$ converges almost everywhere
when $n\to\infty$.
\end{corollary}
The following is a discrete analog of Corollary \ref{3Diff2}.
\begin{corollary}\label{3Diff3}
Let $T\colon L^1(\Omega)+L^\infty(\Omega)\to L^1(\Omega)+L^\infty(\Omega)$ be an absolute
contraction and assume that $T\colon L^2(\Omega)\to L^2(\Omega)$ is analytic. Then $T$
satisfies estimates (\ref{3Tn}) and (\ref{3Delta}) for every $1<p<\infty$ and every $2<q<\infty$.
Moreover for any $x\in L^p(\Omega)$, $T^n(x)$ converges almost everywhere
when $n\to\infty$.
\end{corollary}
\begin{proof} If $T\colon L^2(\Omega)\to L^2(\Omega)$ is analytic, then
$T\colon L^p(\Omega)\to L^p(\Omega)$ is analytic as well for any $1<p<\infty$, by
\cite[Thm. 1.1]{Bl2}. Hence the result follows from Theorem \ref{3Main} and Corollary \ref{72}.
\end{proof}
Of course the above corollary applies if $T\colon L^2(\Omega)\to L^2(\Omega)$ is a
positive selfadjoint operator, more generally if $T$
is normal and satisfies (\ref{3Stolz2}). As a consequence, we extend
the main result of \cite{JR}, as follows. The next statement solves a problem raised
in the latter paper.
\begin{corollary} Let $G$ be a locally compact abelian group and let $L^p(G)$ denote the
corresponding $L^p$-spaces with respect to a Haar measure. Let $\nu$ be a probability
measure on $G$ and let
$T\colon L^1(G)+L^\infty(G)\to L^1(G)+L^\infty(G)$ be the associated convolution
operator,
$$
T(x)\,=\, \nu*x.
$$
Assume that there exists a constant $K>0$ such that
$\vert 1-\widehat{\nu}(s)\vert\leq K (1-\vert\widehat{\nu}(s)\vert)\,$ for any $s\in\widehat{G}$.
Then we have an estimate
$$
\bignorm{\bigl(T^n (x)\bigr)_{n\geq 0}}_{L^p(v^q)}\,\lesssim\,\norm{x}_p,\qquad x\in L^p(\Omega),
$$
for any $1<p<\infty$ and any $2<q<\infty$.
\end{corollary}
\begin{proof} Regard $T$ as an $L^2$-operator. By Fourier analysis, its spectrum is equal to the
essential range of $\widehat{\nu}$. It therefore follows from the assumption and Lemma \ref{3Normal1}
that the operator $T\colon L^2(G)\to L^2(G)$ is analytic. Hence $T$ satisfies the assumptions of
Corollary \ref{3Diff3}, which yields the result.
\end{proof}
\begin{remark}
For an operator $T\in B(L^2(\Omega))$,
let $W(T)=\{\langle T(x),x\rangle\, :\, \norm{x}_2=1\}$ denote the numerical range.
We recall that $W(T)$ is a compact convex set and that $\sigma(T)\subset W(T)$.
Assume that there exists $\gamma\in(0,\frac{\pi}{2})$ such that $W(T)\subset B_\gamma$
(which is a stronger condition than (\ref{3Stolz1}) or (\ref{3Stolz2})). Then according
to \cite{DD}, there exists a constant $C>0$ such that
$$
\norm{f(T)}\leq\,C\sup\bigl\{\vert f(z)\vert\, :\, z\in B_\gamma\bigr\}
$$
for any polynomial $f$. Arguing as in the proof of Lemma \ref{3Normal1}, we deduce
that $T$ is an analytic power bounded operator.
Consequently if $T\colon L^2(\Omega)\to L^2(\Omega)$ is contractively regular and
$W(T)\subset B_\gamma$ for some $\gamma\in(0,\frac{\pi}{2})$, then it satisfies
(\ref{3Tn}).
\end{remark}
|
\section{Introduction}
Salient features in an image have the defining property that they instantly
attract our visual attention towards their locations without the need to scan
the entire image systematically---they are image features which have a ``pop
out" effect \cite{koch}. In addition, such features are in some sense rare or
unique---they have a low probability of being mistaken for other features in an
image, and often best distinguish one particular object category from another
\cite{walker,moosmann,sharon}. This property may allow the visual system to
rapidly segment and classify an image based on its most distinctive features, or
to devote further visual processing resources to specific points in an image. In
evolutionary terms a rapid response to salient features could be critical for
biological systems in predator-prey situations. In computer vision, saliency has
been shown to improve performance in object and feature recognition tasks
\cite{walker,moosmann}.
Saliency is thought to be assigned early in the visual pathway, and models of
the primary visual cortex (V1) are often used for this purpose
\cite{pettet,yen,li98,li99,li01}. One reason is that saliency is assumed to
involve only bottom-up (image-based) mechanisms which group simple features
together and process a scene rapidly and in parallel over the entire visual
field \cite{koch}. Later stages of visual attention probably involve top-down
(task-dependent) cues \cite{itti}. Long range horizontal connections in V1
provide a likely mechanism for rapid integration of information across the
visual field \cite{gilbert,kapadia,bosking}. The salient objects we shall mostly
be concerned with in this work are closed contours---such as the
boundaries of objects and surfaces in images. Simple cells in V1 respond best to
oriented bars and edges like those detected using a gabor filter \cite{jones},
and a contour is represented as a discrete set of line segments in V1. However, any given image will result in many line segments in V1, and most of these will be unrelated to
any salient contour. Long-range horizontal connections modulate neural activity,
and provide a mechanism for grouping together line segments that form
salient contours. For example, two tangents are said to be co-circular when they are tangent to the same circle \cite{parent}. If horizontal connections facilitate co-circularity then neural activity
will be higher (and the feature more salient) when line segments form a circular contour rather than a cluttered background of line segments with random orientations and positions, for example.
Experiments on contour detection by the human visual system have shown that contour saliency depends on factors such as contour length, curvature, degree of
fragmentation, whether the contour is open or closed, and the statistical
properties of the background relative to the contour
\cite{kapadia,kovacs,braun}. Line segments which are co-linear (i.e. lie on a
straight line and are orientated in the same direction as the line) are highly
salient, but saliency decreases as the line segments are either separated or
their relative orientations are changed \cite{kapadia}. The saliency of a
contour can be increased either by increasing the contour length or by closing the contour, but decreases with increase in the contour curvature \cite{kovacs}. A final
point is that saliency also depends on the statistical properties of the
background---false contours formed by fortuitously aligned background line
segments will decrease contour saliency \cite{braun}. Field et.~al.~proposed an association
field for grouping line segments that form contours \cite{field}. We will make use of this in Sec.~\ref{sec3} when choosing a form for horizontal connections.
The purpose of this work is to present a neural model with V1 architecture that
can be used to detect visually salient contours in images. We would also like to determine what level of performance is possible in such tasks using a simple V1 architecture. This architecture
includes functionally specific long-range horizontal connections, local feedback
connections, thresholding, and neural orientation preference. A key motivation is to see how a
biophysically-constrained model of V1 can be used to assign saliency in the
10-20 ms period available under certain viewing conditions \cite{keysers}. We would also like to provide a solid framework for including horizontal connections into certain
classes of biologically-inspired hierarchical neural networks used for object
recognition \cite{serre}. We start with a model of neural activity (spike-rate)
dynamics such as used in Refs.~\cite{li98} and \cite{wilson,bressloff,loxley}
for describing populations of V1 neurons. From these dynamical equations a
convolution model is derived under the assumption that horizontal connections
only modulate neural activity rather than drive it. A convolution model has
the advantage that it avoids the stability problems
found in dynamical neural models \cite{li98,li01}. Operating as a feed-forward network, the convolution model increases the activity of line segments
facilitated by the horizontal connections. Alternatively, operating as a feedback network, it increases the activity of facilitated line segments over several iterations through cooperative interactions.
Previous models of salient contour detection can be divided approximately into
two classes: V1-like neural models with horizontal connections
\cite{pettet,yen,li98}, and non-neural models as typified by
Refs.~\cite{zucker,sha,williams,wang,zhu,ren}. Although the non-neural models are interesting in their own right, here our primary interest is
in biologically-inspired visual processing, and we will not further consider
non-neural models. In Ref.~\cite{pettet}, neural elements were represented
by gabor patches, and a dynamical model with facilitatory horizontal connections
and global inhibition was assumed. Horizontal connections involved fitting a
cubic spline between pairs of neighboring patches, then including penalties for
distance, curvature, and change in curvature---such that co-linear and
co-circular neighbors were facilitated. A region was salient when its steady state
response reached a certain out-lier criterion level separating background noise
from signal. In Ref.~\cite{yen}, neural elements were grouped using a model with
facilitatory horizontal connections and various forms of competitive inhibition.
Horizontal connections were formed from an association field similar to that proposed in
Ref.~\cite{field}, and co-circular neighbors were facilitated. Global inhibition was used
to suppress weakly facilitated elements, and the remaining elements were modeled
as phase oscillators. Salient regions exhibited phase synchrony due to strong
facilitatory interactions. In Ref.~\cite{li98}, each edge segment represented a
pair of neural elements, one excitatory and one inhibitory, which were described
using a dynamical model with horizontal connections. Facilitation came from an association field,
and inhibition was given by so-called \textit{flank inhibition} \cite{li98,li01}. Saliency was
related to the activity level at the steady state. In addition, the nonlinear dynamical
properties of this model caused gaps in contours to be filled in, and activity
to leak out across boundaries \cite{li01}.
The
structure of this paper is as follows: In Sec.~\ref{sec2} the basic model is
derived and explained. In Sec.~\ref{sec3} the model is used to find salient
contours in example images which include a line, a circular closed contour, and a non-circular closed contour. In Sec.~\ref{sec4}, a summary and a discussion are given.
\section{Model}\label{sec2}
The model we consider describes the processing of visual information in a sequence of stages which reside in early parts of the visual pathway. Visual input from an image
arrives at V1 simple cells, which respond most
strongly to visual features such as a bar or edge of a particular orientation. Dynamical processes then rapidly redistribute
the neural input among the simple cells via functionally-specific long
range
horizontal connections. Once neural activity reaches equilibrium after
approximately 10 ms it is either
directly output to later stages of the visual pathway, or else iterated several times
through thresholding and
local feedback connections before being output to later stages. The most salient regions in the image correspond to the regions of highest activity in the V1 model output. We now
describe the three stages in greater detail.
\subsection{Simple Cell Classical Receptive fields}
Visual input to the model comes from an image, and is given by the set of intensity values
$I({\bf{r}})$ for each point ${\bf{r}}=(x,y)$ in the image. To construct the
simple cell classical receptive fields, this input is convolved with the gabor
function $g_{\phi}({\bf{r}})$ before thresholding is applied, giving
\begin{equation}
I_{\phi}({\bf{r}})=H\left (\int
g_{\phi}({\bf{r}}-{\bf{r}}^{\prime})I({\bf{r}}^{\prime})d{\bf{r}}^{\prime}\right ),
\label{gabor}
\end{equation}
where the gabor function is
\begin{eqnarray}
g_{\phi}({\bf{r}})&=&g(R_{\phi}{\bf{r}}),\\
R_{\phi}&=&\left(\begin{array}{cc}
\cos{\phi}&-\sin{\phi}\\
\sin{\phi}&\cos{\phi}\\
\end{array}\label{rot}
\right),
\\ \nonumber \\
g({\bf{r}})&=&\exp{\left(-\frac{x^2+\gamma^2y^2}{2\sigma^2}\right)}\cos{
\left(\frac{2\pi}{\lambda}x+\varphi\right)},\label{endgabor}
\end{eqnarray}
and depends on the parameters $\phi$, $\gamma$, $\sigma$, $\lambda$, and
$\varphi$ \cite{jones}. The integral in Eq.~(\ref{gabor}) is taken over the
whole image, and $H$ is a unit step function satisfying: $H(x)=1$ if
$x\geq \kappa$, and $H(x)=0$ if $x< \kappa$. The key parameter here is $\phi$---the neural orientation
preference---which we have made explicit in Eq.~(\ref{gabor}). All other parameters will be considered as being constant for a particular image.
\subsection{Horizontal Connections and Neural Dynamics}
Simple-cell neural populations will be labeled by their position ${\bf{r}}\in R^{2}$, and
orientation preference $\phi\in(0,\pi)$. The dynamics of simple-cell neural populations is
described by the neural activity $u_{\phi}({\bf{r}},t)$, which corresponds to
the local mean cell-body potential. Given a visual input $I_{\phi}({\bf{r}})$
from the previous stage, neural activity is redistributed according to long
range
orientation-specific horizontal connections
$w_{\phi\phi^{\prime}}({\bf{r}}-{\bf{r}}^{\prime})$, which link cell populations
at
${\bf{r}}$ and $\phi$, with cell populations at ${\bf{r}}^{\prime}$ and
$\phi^{\prime}$.
The equation for neural-activity dynamics has a similar form to that used in
coarse-grained models of V1 \cite{li98,wilson,bressloff,loxley}, and is given by
\begin{eqnarray}
&&\left(\tau\frac{\partial }{\partial t}+1\right)u_{\phi}({\bf{r}},t)\nonumber
\\
&&=\sum_{\phi^{\prime}}\int
w_{\phi\phi^{\prime}}({\bf{r}}-{\bf{r}}^{\prime})S\boldsymbol{(}u_{\phi^{\prime}}({\bf{r
}}^{\prime},t)\boldsymbol{)}d{\bf{r}}^{\prime}+I_{\phi}({\bf{r}}),\label{dyn1}
\end{eqnarray}
where $\tau \simeq 10$ ms is the membrane time constant over which incoming
action potential
spikes are smoothed into pulses at the synapses of a neuron, and $S$ is the
population spiking rate, given by
\begin{equation}
S\left(u\right)=\frac{1}{1+\exp(-(u-\theta)/\sigma^{\prime})},\label{dyn2}
\end{equation}
which is a sigmoid-shaped function, with spiking-rate threshold $\theta$, and
width $\sigma$. The sum in Eq.~(\ref{dyn1}) is over all neural orientation
preferences---presumably forming a discrete set, and the integral is over all neural population positions.
Solutions of Eq.~(\ref{dyn1}) are guaranteed to converge to a stable fixed point
if an associated Lyapunov function exists \cite{loxleyB}. Alternatively, we can
avoid the dynamical stability problem by setting the time derivative in
Eq.~(\ref{dyn1}) to zero, and working with an equilibrium solution. We also let
$\sigma^{\prime}\rightarrow 0$ in Eq.~(\ref{dyn2}) so that we can replace $S$ in Eq.~(\ref{dyn1}) with a unit step function $H$ satisfying: $H(x)=1$ if
$x\geq \theta$, and $H(x)=0$ if $x< \theta$. The final step is to assume
$I_{\phi}({\bf{r}})\geq \theta$ are the only points where
$u_{\phi}({\bf{r}})\geq \theta$, i.e. neural activity is only ever above
threshold
where the input is above threshold. This means horizontal connections
modulate neural activity rather than drive it \cite{gilbert,kapadia,bosking}. We can now replace
$u_{\phi}({\bf{r}})$ with $I_{\phi}({\bf{r}})$ in the argument of $H$, and
Eq.~(\ref{dyn1}) reduces to
\begin{equation}
u_{\phi}({\bf{r}})=\sum_{\phi^{\prime}}\int
w_{\phi\phi^{\prime}}({\bf{r}}-{\bf{r}}^{\prime})H\boldsymbol{(}I_{\phi^{\prime}}({\bf{r
}}^{\prime})\boldsymbol{)}d{\bf{r}}^{\prime}+I_{\phi}({\bf{r}}).\label{eqlm}
\end{equation}
This equation is our convolution model, giving the facilitation due to horizontal connections. The horizontal connections have the form
\begin{equation}
w_{\phi\phi^{\prime}}({\bf{r}})=M_{\phi\phi^{\prime}}
w(R_{\phi^{\prime}}{\bf{r}}),\label{connectivity}
\end{equation}
where $R_{\phi}$ is the rotation matrix from Eq.~(\ref{rot}),
$M_{\phi\phi^{\prime}}$ is a matrix for the orientation preference specificity of
the horizontal connections, and $w({\bf{r}})$ is a function that depends on
the spatial distribution of the connections.
Processing time is limited in many
visual tasks where
bottom-up saliency is required. The equilibrium given by Eq.~(\ref{eqlm}) takes approximately 10 ms to attain, and neural activity can then be directly output to later parts of the visual pathway. Summing neural activity over all orientation preferences gives the total output as
\begin{equation}
u({\bf{r}})=\sum_{\phi} u_{\phi}({\bf{r}}),\label{totalact}
\end{equation}
and the regions of largest $u({\bf{r}})$ are the regions which are most salient in the image input. Alternatively, if a longer processing time is available, neural activity can be iterated through thresholding and local feedback connections, as described next.
\subsection{Thresholding and Local Feedback}
This stage of the model involves thresholding $u_{\phi}({\bf{r}})$ from Eq.~(\ref{eqlm}) to select only the regions where it attains its largest values, returning the result to Eq.~(\ref{eqlm}) as $I_{\phi}({\bf{r}})$, then re-calculating $u_{\phi}({\bf{r}})$, and so on. The pooling operation of V1 complex cells has previously been modeled by selecting the maximum from each group of simple cell outputs \cite{serre}. Here, we use inhibitory neural dynamics to perform a global thresholding of simple cell outputs. In a more detailed model, the pooling operation of V1 complex cells could be used to perform a local thresholding. The equilibrium solution for spatially-uniform inhibitory dynamics is given by
\begin{equation}
I^{\prime}_{\phi}({\bf{r}})= H(u_{\phi}({\bf{r}})- u),\label{thresh}
\end{equation}
where $H$ is a unit step function with threshold $\theta$, and $u$ determines the strength of spatially uniform
inhibition. The solution given by Eq.~(\ref{thresh}) is a simple
thresholding step: defining $\Delta\theta\equiv u$, then all
$u_{\phi}({\bf{r}})<\theta+\Delta\theta$ is discarded in Eq.~(\ref{thresh}). This is consistent with selecting regions where $u_{\phi}({\bf{r}})$ attains its largest values.
Next, we include the effect of local feedback connections by replacing $I_{\phi}({\bf{r}})$ in Eq.~(\ref{eqlm}) with $I^{\prime}_{\phi}({\bf{r}})$ from Eq.~(\ref{thresh}), then use it to re-calculate $u_{\phi}({\bf{r}})$. The whole process is repeated as many times as possible. Iterating Eqs.~(\ref{eqlm}) and (\ref{thresh}) will mean any visual input not facilitated by the horizontal
connections will eventually be discarded due to thresholding. The number of possible iterations is determined by the processing time available for the specific visual task being carried out. Assuming local feedback and thresholding act faster than horizontal integration, each iteration will require at least 10 ms---meaning that only a few iterations may be possible in many low level visual tasks involving saliency.
\section{Detecting salient contours in example images}\label{sec3}
The model given by Eqs.~(\ref{gabor})--(\ref{endgabor}) and (\ref{eqlm})--(\ref{thresh}) is now used to find salient contours in the example images shown in Fig.~1. A salient line made from 4 co-linear line segments is shown in Fig.~1(a) and will be treated in Sec.~\ref{sec3.1}, a salient circular closed contour is shown in Fig.~1(b) and will be treated in Sec.~\ref{sec3.2}, while a salient non-circular closed contour is shown in Fig.~1(c) and will be treated in Sec.~\ref{sec3.3}.
\begin{figure}
\includegraphics[width=150pt,bb=247 237 370 631,clip=true]{imagecombined.eps}
\caption{Example images of salient contours within cluttered backgrounds. (a) Salient line made from 4 co-linear line segments. (b) Salient circular closed contour. (c) Salient non-circular closed contour.}
\end{figure}
\subsection{Form of Horizontal Connections}
Horizontal connections are parameterized to link neural populations which are neighbors in both ${\bf{r}}$ and $\phi$---i.e.~position and orientation preference. For any two populations with $\phi$ and $\phi^{\prime}$, the elements of $M_{\phi\phi^{\prime}}$ in Eq.~(\ref{connectivity}) are chosen as
\begin{equation}
M_{\phi\phi^{\prime}}=\left\{
\begin{array}{ll}
1,&\mathrm{if}\ |\phi \pmod{\pi}-\phi^{\prime}|\leq\phi_{max},\\
0,&\mathrm{otherwise},
\end{array}\right.
\end{equation}
where mod $\pi$ denotes that a factor of $\pi$ can be added or subtracted from $\phi$ in order to make the inequality true. The parameter $\phi_{max}$ is the size of the $\phi$-neighborhood of neural populations which are connected. The form of $w({\bf{r}})$ used in Eq.~(\ref{connectivity}) is shown in Fig.~2, and is a modified form of the association field proposed in Ref.~\cite{field}. The spatial distribution of connections is anisotropic: strong excitation along the axis of orientation preference (shown as $\phi=0$ in Fig.~2) means that co-linear line segments are strongly facilitated. Crude facilitation of co-circular line segments also takes place due to the angular dispersion of excitation off the orientation preference axis. The parameters $r_{1}$ and $r_{2}$ determine the minimum and maximum range of connections---$r_{1}$ is a short-range cutoff intended to eliminate self-excitation of short line segments, $\psi$ specifies the angular dispersion of excitation off the orientation preference axis, while $w_{e}>0$ and $w_{i}<0$ are the strengths of excitation and inhibition, respectively.
\begin{figure}
\includegraphics[width=140pt, bb=72 166 589 650, clip=true]{mod2connectplot.eps}
\caption{Plot of the function $w({\bf{r}})$ used in Eq.~(\ref{connectivity}). The parameters $r_{1}$ and $r_{2}$ give the minimum and maximum range (arrows) of horizontal connections, $\psi$ is the angular dispersion of excitation off the orientation preference axis (shown as the horizontal axis), while $w_{e}$ and $w_{i}$ are the strengths of excitation and inhibition, respectively.}
\end{figure}
\subsection{Salient Line}\label{sec3.1}
Now we apply our model to group co-linear line segments which form a salient line. The image in Fig.~1(a) contains a single salient line made of 4 co-linear line segments within a cluttered background of 104 line segments with random orientations and positions. The image is normalized to have side lengths of 1 unit, setting the length scale for parameters describing the receptive field size and horizontal connection range. The size of each line segment is 0.0417$\times$0.0125, and the value of $I({\bf{r}})$ is 1 on a line segment, and 0 elsewhere.
To detect the salient line in Fig.~1(a) the first step is to find $I_{\phi}({\bf{r}})$---that is, for each value of $\phi$ we find the set of line segments which trigger the largest neural activity response. We discretize $\phi$ into $k$ elements between 0 and $\pi$: $\phi=0$, $\pi/k$, $2\pi/k$,..., $(k-1)\pi/k$; and set $k=10$. Applying Eqs.~(\ref{gabor})--(\ref{endgabor}) to the image in Fig.~1(a) yields $I_{\phi}({\bf{r}})$, as shown in Fig.~3(a) for all values of $\phi$. This corresponds to the simple-cell neural activity. Next, applying Eqs.~(\ref{eqlm})--(\ref{totalact}) yields the neural activity $u({\bf{r}})$ from the feed-forward model, as shown in Fig.~3(b). The model parameters are given in the figure caption, and the values of $\psi$ and $r_{1}$ were chosen to facilitate co-linearity and a short-range cutoff (where only line segments separated by distance $d\geq r_{1}$ are facilitated). Several groups of co-linear line segments above the cutoff range $r_{1}$ have been facilitated by horizontal connections in Fig.~3(b), but only one of these corresponds to the salient line in Fig.~1(a)---the rest are fortuitous alignments of background clutter segments. Finally, iterating Eqs.~(\ref{eqlm}) and (\ref{thresh}) 4 times yields $u({\bf{r}})$ from the feedback model, as shown in Fig.~3(c). Cooperative interactions between facilitated line segments take place over the 4 iterations until the only activity that remains corresponds to the salient line in Fig.~1(a).
\begin{figure}
\includegraphics[width=150pt,bb=250 244 370 618,clip=true]{modelfig1.eps}
\caption{Saliency assigned by different stages of the V1 model for the image in Fig.~1(a). (a) Simple cell activity $I_{\phi}({\bf{r}})$ for all $\phi$ from Eqs.~(\ref{gabor})--(\ref{endgabor}) with Fig.~1(a) for $I({\bf{r}})$. (b) Feed-forward activity $u({\bf{r}})$ from Eqs.~(\ref{eqlm})--(\ref{totalact}). Only highest activity is shown. (c) Feedback activity $u({\bf{r}})$ from iterating Eqs.~(\ref{eqlm}) and (\ref{thresh}) 4 times. Parameters are: $\kappa=0.06$, $\gamma=0.5$, $\sigma=0.008$, $\lambda=0.025$, $\varphi=0$, $r_{1}=0.125$, $r_{2}=0.175$, $\psi=\pi/12$, $w_{e}=0.5$, $w_{i}=0$, $\phi_{max}=\pi/12$, $\theta=0.6$, and $u=0.44$.}
\end{figure}
\subsection{Circular Closed Contour}\label{sec3.2}
The model is now used to group line segments which form a closed contour that is approximately circular. The image in Fig.~1(b) contains a salient closed contour made from 10 line segments within a cluttered background of 97 line segments with random orientations and positions. Image details are the same as in Sec.~\ref{sec3.1}.
Applying Eqs.~(\ref{gabor})--(\ref{endgabor}) to the image in Fig.~1(b) yields $I_{\phi}({\bf{r}})$, as shown in Fig.~4(a) for all values of $\phi$. Next, applying Eqs.~(\ref{eqlm})--(\ref{totalact}) yields the neural activity $u({\bf{r}})$ from the feed-forward model, as shown in Fig.~4(b). The model parameters are given in the figure caption, and the values of $\psi$ and $\phi_{max}$ were chosen to facilitate line segments which form circular arcs as well as lines. Several groups of line segments forming circular arcs and lines have been facilitated by horizontal connections in Fig.~4(b), however, only one of these corresponds to the closed contour in Fig.~1(b). As shown in the previous subsection, the feed-forward model filters out much of the background clutter from the image, while retaining the salient feature.
\begin{figure}
\includegraphics[width=150pt, bb=254 256 363 600,clip=true]{newmodelfig2.eps}
\caption{Saliency assigned by different stages of the V1 model for the image in Fig.~1(b). (a) Simple cell activity using Fig.~1(b) for $I({\bf{r}})$. (b) Feed-forward activity from Eqs.~(\ref{eqlm})--(\ref{totalact}). Only highest activity is shown. Parameters are: $r_{1}=0.042$, $r_{2}=0.117$, $\psi=\pi/6$, $w_{i}=-0.3$, $\phi_{max}=\pi/5$, and others as in Fig.~3. (c) Feedback activity from iterating Eqs.~(\ref{eqlm}) and (\ref{thresh}) 7 times with simple cell activity from Fig.~5(a). Parameters are now: $r_{2}=0.133$, $\psi=\pi/4$, $\phi_{max}=\pi/3$, $u=0.473+0.017n$ (for iterate $n$) and others as in (b).}
\end{figure}
The feedback model activity is shown in Fig.~4(c), and results from iterating Eqs.~(\ref{eqlm}) and (\ref{thresh}) 7 times with the simple cell activity from Fig.~4(a) and a different set of parameter values. It is seen that the only activity remaining corresponds to the salient closed contour in Fig.~1(b). The model parameters are given in the figure caption, and the values of $r_{2}$, $\psi$, and $\phi_{max}$ have been increased relative to their values in Fig.~4(b). The reason for this was to increase the size of the excitatory component of the horizontal connections in order to increase facilitation between neighboring line segments in the closed contour. This can be seen in Fig.~5, where the bottom line segment of the closed contour is completely facilitated by its left and right neighbors. A general principal now applies: Each line segment that is part of a closed contour is facilitated by two neighbors, while line segments at the ends of an open contour are facilitated by only one neighbor. This principal allows cooperative interactions in the feedback model to group line segments according to whether or not they form closed contours. Open contours get shorter and shorter with each iteration of feedback as line segments are discarded from the ends. A closed contour will also become inhibited less and less as nearby clutter segments are iteratively discarded. In order to maintain a balance between excitation and inhibition in this model, the strength of inhibition in the thresholding step needs to be increased as inhibition due to horizontal connections gradually decreases. This is done linearly in the number of iterations using the form for $u$ given in Fig.~4.
\begin{figure}
\includegraphics[width=150pt,bb=250 293 366 537]{newmodelfig3.eps}
\caption{Facilitation of the bottom line segment in Fig.4(c) by its left and right neighbors. (a) and (b) show an enlargement of the excitatory component of horizontal connections (Grey) due to each neighbor. Parameters are as in Fig.~4(c).}
\end{figure}
\subsection{Non-Circular Closed Contour}\label{sec3.3}
The model is now used to group line segments which form a closed contour that is not even approximately circular, but has an arbitrary shape. The image in Fig.~1(c) consists of several curved segments, some making up the closed contour and the rest forming background clutter. Applying Eqs.~(\ref{gabor})--(\ref{endgabor}) to the image in Fig.~1(c) yields $I_{\phi}({\bf{r}})$, as shown in Fig.~6(a) for all values of $\phi$. Next, applying Eqs.~(\ref{eqlm})--(\ref{totalact}) yields the neural activity $u({\bf{r}})$ from the feed-forward model, as shown in Fig.~6(b). The feedback model activity is shown in Fig.~6(c), and results from iterating Eqs.~(\ref{eqlm}) and (\ref{thresh}) 9 times. It is seen that this activity approximates the salient closed contour in Fig.~1(c).
The model parameters are given in the figure caption, and the value of $\phi_{max}$ indicates that horizontal connections are not orientation specific---line segments of all orientations are connected together. The reason is that the closed contour is not even close to being smooth or circular, and has several regions where line segments actually meet at right angles instead of having smooth circular joins. Models which are limited to co-circular connections would not be able to detect a closed contour in this case.
\begin{figure}
\includegraphics[width=150pt, bb=246 236 371 635,clip=true]{newmodelfig4.eps}
\caption{(a) Simple cell activity using Fig.~1(c) for $I({\bf{r}})$. (b) Feed-forward activity from Eqs.~(\ref{eqlm})--(\ref{totalact}). Only highest activity is shown. (c) Feedback activity from iterating Eqs.~(\ref{eqlm}) and (\ref{thresh}) 9 times. Parameters are: $k=8$, $\sigma=0.017$, $\lambda=0.048$, $r_{1}=0.05$, $r_{2}=0.117$, $\psi=\pi/3$, $w_{i}=0$, $\phi_{max}=\pi/2$, $u=0.66$, and others as in Fig.~3.}
\end{figure}
\section{Summary and Discussion}\label{sec4}
In this work, our primary aim has been to establish a solid foundation for modeling horizontal connections in biological neural networks. Our model takes into account the known classical receptive fields of simple cells in V1, the activity dynamics of neural populations, and some of the basic properties of horizontal connections found in physiological and psychophysical studies.
We have applied our model to the detection of salient contours in three different cases: a line, a circular closed contour, and a non-circular closed contour. Using different values of the model parameters the model performed well in all cases: assigning higher neural activity to line segments forming a salient contour, and lower activity to line segments forming background clutter. Values of the model parameters for each case indicated different facilitation mechanisms taking place. Horizontal connections facilitating co-linear line segments and a short-range cutoff were useful for detecting salient lines. Facilitation of line segments forming circular arcs was found to be useful for detecting salient closed contours which were approximately circular. In addition, cooperative interactions were able to distinguish between open and closed contours---open contours getting shorter and shorter with each iteration as line segments were discarded from the ends, while closed contours remained unchanged. For the detection of non-circular closed contours facilitation of line segments of all orientations was found to be necessary. The reason was that there was no guarantee of having smooth circular joins between different line segments in this contour.
Finally, we showed how our model can operate as either a feed-forward or a feedback network. Operating as a feed-forward network, the model assigns saliency within 10 ms, and increases the neural activity of line segments facilitated by the horizontal connections. In Figs.~3(b) and 4(b), the feed-forward network filtered out much of the background clutter while retaining the salient contour. Operating as a feedback network, further improvement is achieved through cooperative interactions over several iterations. Further, our model does not suffer from stability problems as dynamical neural models sometimes do. This makes it suitable for use in assigning saliency in biologically-inspired neural networks performing object recognition.
Several further questions remain. Firstly, it is difficult to know what model parameter values are actually used in computations involving V1 horizontal connections. It seems unlikely that precise definitions such as co-circularity are implemented by horizontal connections---this property would only apply to a small number of all salient contours. We favor a less stringent construction involving a modified association field. Also, scale invariance has been observed in natural images. To include this in our model would require a range of receptive field sizes and horizontal connection ranges to be considered.
The performance of our model in the successful detection of a salient feature depends on details of the background clutter relative to the feature. For example, operating as a feedback network, the number of iterations taken to distinguish between open and closed contours will depend directly on contour length. A long open contour will take a larger number of iterations to resolve than a short one. Similarly, detecting the most salient feature from a set of salient features could also take a large number of iterations. Addressing these questions, and using bottom-up saliency to robustly improve performance in object recognition tasks, will be the subject of future work.
\begin{acknowledgments}
We would like to thank S.~Barr for the image in Fig.~1(c), and G.~Kenyon and S.~Brumby for useful related discussions. We gratefully acknowledge the support of the U.S. Department of Energy through
the LANL/LDRD Program project 20090006DR for this work.
\end{acknowledgments}
|
\section{Introduction}
\subsection{Goursat distributions and small growth vectors}
Let $M$ be a smooth manifold of real dimension $n \geq 4$. Let $D \subset TM$ be a smooth distribution (subbundle) of corank $s$. Let $D^2 = [D, D] + D$, called the $\textit{Lie square}$ of $D$. Iterate this squaring to obtain a chain
\[ D_s \subseteq D_{s-1} \subseteq \cdots \subseteq D_i \subseteq D_{i-1} \subseteq \cdots \]
where $D_s=D$ and $D_{i-1}=D_i^2$ for $i\leq s$. Note that $D_i$ may not, in general, have constant rank, and thus fail to be a distribution on $M$. \emph{D} is called \emph{Goursat} if corank $D_i =i$.
In this case, one has a $\textit{Goursat flag} \ \mathcal{F}$:
\[ D_s \subset D_{s-1} \subset \cdots \subset D_1 \subset D_0=TM \]
Note that when $D$ is Goursat, each member $D_i$ of the flag is itself a distribution, and a hyperplane in $D_{i-1}$.
Given a Goursat distribution $D$, one can alternatively form the sequence $D^{(i)}=[D, D^{(i-1)}]+D^{(i-1)}$, where $D^{(0)}=D$ and $i \geq 1$. It is not hard to show that this sequence will also eventually terminate. That is, there exists an $r$ such that $D^{(r)}=TM$. Thus, Goursat distributions are \emph{completely nonholonomic}. The least such $r$ is call the \emph{degree of nonholonomy}. For each $p \in M$, we define the \emph{small growth vector at p} to be the integer valued vector
\[sgv(p)= (\text{dim} D^{(0)}(p), \text{dim} D^{(1)}(p), \dots , \text{dim} D^{(r)}(p)=n) \]
While the small growth vector is the traditional object of interest in the theory of nonholonomic distributions, for us it is more convenient to work with the \emph{derived vector}, which is equivalent in our setting.
\begin{definition}
\label{derdef}
The \emph{derived vector} of a small growth vector consists of the multiplicities of the entries in the small growth vector.
\end{definition}
For a Goursat distribution, the dimensions of the sequence $D^{(i)}$ grow by at most one at a time, so from the list of multiplicities we may recover the original small growth vector. By convention, we omit the last multiplicity 1 from the derived vector. For example, if we are given a small growth vector (2, 3, 4, 4, 5), the associated derived vector is (1, 1, 2). Similarly, given a derived vector (1, 1, 1, 3, 3), the associated small growth vector is (2, 3, 4, 5, 5, 5, 6, 6, 6, 7). It is not obvious, but follows from Jean's work, that the derived vector is always increasing.
\subsection{History}
Goursat distributions are the antithesis of integrable distributions, as they are \emph{bracket-generating}. However, they grow the slowest of all such distributions. Cartan (\cite{C}) studied the model of the canonical distribution on the jet space $J^{n-2}(\mathbb{R}, \mathbb{R})$. All Goursat distributions were believed to be equivalent to Cartan's until Giaro, Kumpera, and Ruiz discovered the first singularity in 1978 (\cite{GKR}). A generic Goursat distribution is in fact equivalent to Cartan's (modulo trivial factors).
Montgomery and Zhitomirskii (\cite{MZ2}) introduced the Monster tower, a sequence of manifolds with distributions in which every Goursat germ occurs, along with a geometric coding in the letters RVT which is an invariant of the distribution germ with respect to diffeomorphic equivalence. Jean (\cite{J}) studied the kinematic model of a car pulling $N$ trailers and derived recurrence relations enabling one to compute the small growth vector of a distribution germ from the RVT code. Mormul (\cite{Mo1}) solved Jean's relations, allowing for the calculation of an RVT code from the small growth vector.
In \cite{MZ1}, Montgomery and Zhitomirskii showed that Goursat germs correspond to finite jets of Legendrian curve germs, and that the RVT coding corresponds to the classical invariant in the singularity theory of planar curves: the Puiseux characteristic (see next section). They gave an explicit algorithm for computing the Puiseux characteristic from the RVT code.
\subsection{The Puiseux characteristic}
Suppose $\gamma \colon (\mathbb{R}, 0) \to \mathbb{R}^2$ is the parametrization of an analytic plane curve germ. We say $\gamma$ is \textit{badly-parametrized} if there exist analytic germs $\mu \colon (\mathbb{R},0) \to \mathbb{R}^2, \phi \colon (\mathbb{R},0)\to (\mathbb{R},0)$ such that $d\phi/ dt(0)=0$ and $\gamma=\mu \circ \phi$. Otherwise, $\gamma$ is called \textit{well-parametrized}. If $\gamma$ is well-parameterized and not immersed then we may define its Puiseux characteristic, an invariant with respect the RL-equivalence of curve germs. Then, up to RL-equivalence, $\gamma$ has the form
\[ \gamma(t)=(t^m, \sum_{k \geq m} a_k t^k) \]
where $m \geq 2$.
The definition of the Puiseux characteristic is the following.
Let $\lambda_0=e_0=m$. Then define inductively for $i \geq 0$
\[ \lambda_{i+1} = \text{min} \{ k \ | \ a_k \neq 0,\ e_i \nmid k\}, \quad \quad e_{i+1}= \text{gcd}(e_i, \lambda_{i+1}) \]
until we first obtain a $g$ with $e_g=1$. Then the vector $[\lambda_0 ; \lambda_1, \dots, \lambda_g]$ is called the \textit{Puiseux characteristic} of $\gamma$.
The Puisuex characteristic is the fundamental invariant in the singularity theory of plane curves. In \cite{W}, Proposition 4.3.8 shows that it is equivalent to at least seven other classical invariants.
\section{Main Result}
In the present contribution we effectively compose the algorithms presented in \cite{Mo1} and \cite{MZ1}, yielding a formula for the Puiseux characteristic of the curve corresponding to a Goursat germ with given small growth vector. This formula turns out to be simpler than either of the two from which it was derived, suggesting a deeper geometric link between singularities of plane curves and singular Goursat distributions. The problem solved herein was first proposed in \cite{MZ1} as Question 9.19, part 3, and was asked again in \cite{Mo2} in the Afterword.
\subsection{Main Theorem}
Suppose we are given a small growth vector whose derived vector (see Definition \ref{derdef}) is
\[ der=(d_1, \dots d_N). \]
Consider the set $S=\{ d_i | \ d_{i-1} \text{properly divides}\ d_i \}$. Let $g=|S|$. For $1\leq j\leq g$ denote by $d_{k_j}$ the entries of $S$ in decreasing order. That is, write $S=\{ d_{k_1}, \dots d_{k_g} \}$, with $d_{k_1}> \dots> d_{k_g}$.
\begin{theorem}
\label{maintheorem}
The corresponding Puiseux characteristic is
\[ [\lambda_0; \lambda_1, \dots, \lambda_{g}] \]
where
\begin{align*}
\lambda_0&= d_N \\
\lambda_j &=\sum_{i\geq k_j} d_i+d_{k_j} +d_{k_j-1}
\end{align*}
for $1\leq j\leq g$.
\end{theorem}
\subsection{Example}
Suppose $der=(1, 1, 2, 2, 2, 2, 2, 2, 4, 6, 6, 6, 18, 24, 24)$. Note that $d_N=d_{15}=24$. We also have
$S=\{18, 4, 2\}$, and therefore $g=3$. Then write $S=\{18, 4, 2 \}=\{d_{13}, d_9, d_3 \}$ so that $k_1=13,\ k_2=9,$ and $k_3=3$.
Finally, we compute
\begin{align*}
\lambda_1&= \sum_{i\geq 13} d_i+d_{13} +d_{12}=90 \\
\lambda_2&= \sum_{i\geq 9} d_i+d_{9} +d_{8}=94 \\
\lambda_3&= \sum_{i\geq 3} d_i+d_{3} +d_{2}=103 \\
\end{align*}
The Puiseux characteristic is thus
\[ [24; 90, 94, 103] \]
\subsection{Remark}
For the proof of the theorem, we need alternative notation which captures the multiplicities of repeated entries in the derived vector. To this end, we rewrite the derived vector as
\[ der=(\underbrace{M_1, \ M_1, \dots, M_1}_{m_1}, \ \underbrace{M_2, \ M_2, \dots, M_2}_{m_2}, \dots, \underbrace{M_{v+1}, \ M_{v+1}, \dots, M_{v+1}}_{m_{v+1}}). \]
Assume that $m_1=M_2$. This ensures that the vector represents a critical RVT code (that is, one which ends with the letter V or T). We can only discuss the Puiseux characteristic for critical (non-immersed) plane curves, since any immersed curve germ has normal form $(t, 0)$. Critical RVT codes correspond to critical curves. Then $S=\{ M_i | \ M_{i-1} \text{divides}\ M_i \}$. Let $N_1, N_2, \dots, N_g$ denote the elements of $S$ in decreasing order. We always have $N_g=M_2$, since $M_1=1$. For $1\leq j\leq g$ let $M_{k_j}=N_j$. In this notation, Theorem \ref{maintheorem} says that the corresponding Puiseux characteristic is
\[ [\lambda_0; \lambda_1, \dots, \lambda_{g}] \]
where
\begin{align}
\label{eqn1}
\lambda_0&= M_{v+1} \\
\label{eqn2}
\lambda_j &=\sum_{i\geq k_j} m_i M_i+M_{k_j} +M_{k_j-1}
\end{align}
for $1\leq j\leq g$.
\subsection{Mormul's results}
Before proving the Theorem, we recall Mormul's construction of the RVT code from the derived vector (see \cite{Mo1}), with slightly modified notation.
We have that $v$ is the number of letters V in the RVT code $\alpha$. We have thus (as in section 3.1) partitioned our code in $v+1$ pieces seperated by the letters V. We write that the \textit{last} letter V in the code is followed by $t_1$ many letters T, and then $r_1$ many letters R. We continue for $1 \leq i \leq v$ letting $t_i$ denote the number of letters T following the $i$th V \textit{from the right}, and $r_i$ denote the number of letters R following those letters T. Let $r_{v+1}$ denote the number of letters R preceding the first letter V. Mormul derived the following relations: \\
$r_{v+1}=m_{v+1}+1 $ \\
$t_1=M_2 -2$ \\
$r_1=m_1-M_2 $ \\
For $2\leq j \leq v $ we have
\textbf{Case 1}: $M_j$ divides $M_{j+1}$. Then \\
$\indent t_j= \frac{M_{j+1}}{M_j} -2$ \\
$\indent r_j= m_j-t_j -1$
\textbf{Case 2}: $M_j$ does not divide $M_{j+1}$. Then \\
$\indent t_j= m_j-1$ \\
$\indent r_j= 0$ \\
We now prove the theorem.
\section{Proof of Theorem}
\subsection{Proof of Theorem; v=1}
Assume $v=1$, so we need only compute the multiplicities $r_2, t_1$, and $r_1$. Recall that $m_1=M_2$ here as well. From Mormul's relations, we have
\begin{align*}
r_2&= m_2+1 \\
t_1&= M_2-2 \\
r_1&= m_1-t_1-2=0
\end{align*}
Thus, the given derived vector corresponds to the RVT code $(\alpha)= R^{m_2+1}VT^{M_2-2}$. We now compute the Puiseux characteristic following Montgomery and Zhitomirskii (section 3.8.4 in \cite{MZ1}).
\[(a, b)= \mathbb{E}_{VT^{M_2-2}}(1, 2)=\mathbb{E}_{VT^{M_2-3}}(1, 3)=\cdots=\mathbb{E}_{V}(1, M_2)=(M_2, M_2+1)
\]
Thus, the Puiseux characteristic is Pc($der)=[\lambda_0; \lambda_1]$ where
\begin{align*}
\lambda_0 &= a = M_2 \\
\lambda_1 &= (m_2-1)a + a +b \\
&=m_2M_2 + (M_2 +1) \\
&=(m_2+1)M_2+M_1
\end{align*}
This proves \eqref{eqn1} and \eqref{eqn2} for $v=1$.
\begin{flushright}
$\blacksquare$
\end{flushright}
\subsection{Proof of Theorem; Case A}
This is the case when $g=1$. That is, we assume $M_i$ does not divide $M_{i+1}$ for all $i \geq 2$. Mormul's relations give
\begin{align*}
r_{v+1}&=m_{v+1}+1 \\
t_1&=M_2-2 \\
r_1&=m_1-M_2=0
\end{align*}
and for $j=2, \dots, v$ we are in ``Case 2" so that
\[ t_j=m_j-1, \quad \quad r_j=0.
\]
Thus our RVT code is
\[ (\alpha)={R^{m_{v+1}+1}VT^{m_v-1}VT^{m_{v-1}-1}\cdots VT^{m_3-1}VT^{m_2-1}VT^{M_2-2}}
\]
We can now see that this case corresponding to ``Case A" in \cite{MZ1}. That is, our RVT code is of the form
\[ (\alpha)=R^l\omega \]
where $\omega$ is entirely critical. Here we have $l=m_{v+1}+1$. Our task is now to compute $(a,b)=\mathbb{E}_\omega (1,2)$. We make the following 2 claims: \\
\textit{Claim 1:} $a=M_{v+1}$ \\
\textit{Claim 2:} $b= \sum_{i=2}^{v} m_i M_i +M_2 +M_1$ \\
It then follows from \cite{MZ1} that the Puiseux characteristic is $[\lambda_0; \lambda_1]$, where
\begin{align*}
\lambda_0&=a=M_{v+1} \\
\lambda_1&=(l-2)a+a+b \\
&=(m_{v+1}-1)M_{v+1} +M_{v+1} +\sum_{i=2}^{v} m_i M_i +M_2 +M_1 \\
&= \sum_{i=2}^{v+1} m_i M_i +M_2 +M_1
\end{align*}
which is the precisely the content of \eqref{eqn1} and \eqref{eqn2} in this case.
It remains to prove the claims. We note that the value $l=m_{v+1}+1$ is irrelevant here, and we assume without loss of generality that $l=2$. We will prove the claims by induction on $v$. The case $v=1$ is done above. Now suppose $v=2$, so that
\[ (\alpha)=RRVT^{m_2-1}VT^{M_2-2}. \]
We now compute $(a,b)$, following \cite{MZ1}:
\begin{align*}
(a,b)=\mathbb{E}_\omega(1,2) &= \mathbb{E}_{VT^{m_2-1}VT^{M_2-2}} (1,2) \\
&=\mathbb{E}_{VT^{m_2-1}VT^{M_2-3}} (1,3) \\
&=\mathbb{E}_{VT^{m_2-1}VT^{M_2-4}} (1,4) \\
&\quad \vdots \\
&=\mathbb{E}_{VT^{m_2-1}V} (1,M_2) \\
&=\mathbb{E}_{VT^{m_2-1}} (M_2,M_2+1) \\
&=\mathbb{E}_{VT^{m_2-2}} (M_2,2M_2+1) \\
&=\mathbb{E}_{VT^{m_2-3}} (M_2,3M_2+1) \\
&\quad \vdots \\
&=\mathbb{E}_{V} (M_2,m_2M_2+1) \\
&=(m_2M_2+1, m_2M_2+M_2+1) \\
&=(m_2M_2+1, m_2M_2+M_2+M_1)
\end{align*}
This proves claim 2 for $v=2$, and gives $a=m_2M_2+1$.
We next compute $der(\alpha)$ following Mormul:
\begin{align*}
(1,1) &\overset{V}{\rightarrow} (1,1,2) \\
&\overset{T}{\rightarrow} (1,1,1,3) \\
&\overset{T}{\rightarrow} (1,1,1,4) \\
&\quad \vdots \\
&\overset{T}{\rightarrow} (1^{m_2}, m_2) \\
&\overset{T}{\rightarrow} (1^{m_2+1},m_2+1) \\
&\overset{V}{\rightarrow} (1,1,2^{m_2}, 2m_2+1) \\
&\overset{T}{\rightarrow} (1,1,1,3^{m_2},3m_2+1) \\
&\overset{T}{\rightarrow} (1,1,1,1,4^{m_2},4m_2+1) \\
&\overset{T}{\rightarrow} (1^5,5^{m_2},5m_2+1) \\
&\quad \vdots \\
&\overset{T}{\rightarrow} (1^{M_2},M_2^{m_2}, m_2M_2+1) \\
\end{align*}
So $M_3=m_2M_2+1=a$, proving Claim 1 for $v=1$.
Both claims have been shown for $v=1$, so we continue to the inductive step.
Now let
\begin{align*} der&=(M_1^{m_1},M_2^{m_2}, \dots, M_v^{m_v}, M_{v+1}^{m_{v+1}}) \\
der'&=(M_1^{m_1},M_2^{m_2}, \dots, M_v^{m_v})
\end{align*}
with associated RVT codes
\begin{align*}
(\alpha)&=R^{m_{v+1}+1}VT^{m_v-1}VT^{m_{v-1}-1}\cdots VT^{m_3-1}VT^{m_2-1}VT^{M_2-2} \\
(\alpha')&=R^{m_{v}+1}VT^{m_{v-1}-1}\cdots VT^{m_3-1}VT^{m_2-1}VT^{M_2-2}
\end{align*}
so that
\begin{align*}
(\omega)&=VT^{m_v-1}VT^{m_{v-1}-1}\cdots VT^{m_3-1}VT^{m_2-1}VT^{M_2-2} \\
(\omega')&=VT^{m_{v-1}-1}\cdots VT^{m_3-1}VT^{m_2-1}VT^{M_2-2}.
\end{align*}
Then by induction, we have
\[(a',b')= \mathbb{E}_{\omega'}(1,2)=\big(M_v, \sum_{i=2}^{v-1} m_i M_i +M_2 +M_1\big).\]
So we finally have
\begin{align*}
(a,b)=\mathbb{E}_\omega(1,2) &=\mathbb{E}_{VT^{m_v-1}}(a',b')\\
&=\mathbb{E}_{VT^{m_v-1}}(M_v, b') \\
&=\mathbb{E}_{VT^{m_v-2}}(M_v, b'+M_v) \\
& \quad \vdots \\
&=\mathbb{E}_{V}(M_v, b'+(m_v-1)M_v) \\
&=(b'+(m_v-1)M_v, b'+m_vM_v) \\
&=\Bigg(\sum_{i=2}^{v} m_i M_i +M_2 +M_1-M_v,\ \sum_{i=2}^{v} m_i M_i +M_2 +M_1\Bigg)
\end{align*}
This proves Claim 2, and reduces Claim 1 to showing
\begin{align}
\label{eqn3}
M_{v+1}=\sum_{i=2}^{v} m_i M_i +M_2 +M_1-M_v
\end{align}
To show \eqref{eqn3}, we apply induction on $v$. That is, we assume
\[M_{v}=\sum_{i=2}^{v-1} m_i M_i +M_2 +M_1-M_{v-1}.
\]
Then
\begin{align*}
M_{v+1}+M_v&=M_{v+1}+ \sum_{i=2}^{v-1} m_i M_i +M_2 +M_1-M_{v-1} \\
\end{align*}
so that \eqref{eqn3} reduces to
\[ M_{v+1}-M_{v-1}=m_vM_v. \]
This last equality, however, follows from Proposition 3.2 and Theorem 3.3 of \cite{Mo1} (as does the base case $v=2$ in the above induction).
This completes the proof of Case A.
\begin{flushright}
$\blacksquare$
\end{flushright}
\subsection{Proof of Theorem; Case B}
This is the case when $g>1$. Mormul's results imply that this corresponds to multiple critical strings in the associated RVT code $(\alpha)$. The proof will be by induction on $g$, where the base case was completed in the previous section. The idea is to truncate the code $(\alpha)$ after the last occurring letter R. To this end, we need only consider the entry $N_{g-1}=M_{k_g-1}$ in $der$. For sake of cleaner notation, we set $r=k_{g-1}$. Then by assumption we have that $M_{j-1}$ divides $M_j$, and $M_{j-1}$ is the smallest such entry (besides $M_1=1$). Then Mormul's relations imply that the RVT code has the form
\[(\alpha)=(\beta R^s \omega)
\]
where
\begin{align*}
\omega &= VT^{m_{r-1}-1}VT^{m_{r-2}-1} \cdots VT^{m_{2}-1}VT^{M_{2}-2} \\
s&= m_r-\frac{M_{r+1}}{M_r}+1 \\
\beta &= \text{some critical RVT code}
\end{align*}
In fact, we can say slightly more about ($\beta$) - it ends with $VT^{M_{r+1}/M_r-2}$ - but we will not use this fact. We will adorn all data concerning ($\beta$) with a tilde to distinguish it from that of ($\alpha$). In particular, we write $[\tilde{\lambda}_0; \tilde{\lambda}_1, \dots, \tilde{\lambda}_{g-1}]$ for the Puiseux characteristic of $(\beta)$, and
\[(\tilde{M}_1^{\tilde{m}_1}, \tilde{M}_2^{\tilde{m}_2}, \dots, \tilde{M}_{\tilde{v}+1}^{\tilde{m}_{\tilde{v}+1}})\]
for the derived vector. Note that $\tilde{g}$ is indeed equal to $g-1$ by construction. While not obvious from Mormul's relations listed above, in ``Case 1" one always has $r_j>0$. This essentially follows from a result of Luca and Risler (see \cite{LR}) which states that the degree of nonholonomy of a Goursat distribution on an $n$-manifold cannot exceed the $n$th Fibonacci number. (Here the degree of nonholonomy is equal to the sum of the entries in the derived vector plus one).
Now a result of Mormul (Proposition 1, \cite{Mo2}) allows us to compute the derived vector for ($\beta$) in terms of the derived vector for $(\alpha)$:
\begin{align*}
\tilde{m}_1 &= \frac{M_{r+1}}{M_r} \\
\tilde{v}&=v+r-1 \\
\tilde{M}_i&=\frac{M_{r+i-1}}{M_r} \quad \text{for} \ i=2, \dots, \tilde{v}+1 \\
\tilde{m}_{i+1}&=m_{r+i} \quad \text{for} \ i=1, \dots, v-r+1
\end{align*}
Now by induction we may assume
\begin{align*}
\tilde{\lambda}_0&= \tilde{M}_{\tilde{v}+1} \\
\tilde{\lambda}_j &=\sum_{i\geq \tilde{k}_j} \tilde{m}_i \tilde{M}_i+\tilde{M}_{\tilde{k}_j} +\tilde{M}_{\tilde{k}_j-1}
\end{align*}
for $1\leq j\leq g-1$. So the above relations imply
\[\tilde{\lambda}_0= \frac{M_{v+1}}{M_r}.
\]
Now the two claims from the previous section imply
\[(a,b)=\mathbb{E}_\omega (1,2)=(M_{r}, \sum_{i=2}^{r-1} m_i M_i +M_2 +M_1).
\]
Then the algorithm described in \cite{MZ1} gives
\[\lambda_0= a\tilde{\lambda}_0 = M_r \frac{M_{v+1}}{M_r}= M_{v+1},
\]
agreeing with \eqref{eqn1} as desired.
We now compute $\lambda_g$. We always have $M_1=1$ so $N_g=M_2$ so $k_g=2$. We first observe that
\begin{align*}
M_r \tilde{\lambda}_{g-1} &= \sum_{i\geq2} m_{r+i-1} M_{r+i-1}+M_{r+1}+M_{r} \\
&= \sum_{i\geq r+1} m_{i} M_{i}+M_{r+1}+M_{r}
\end{align*}
Thus, following \cite{MZ1} we obtain
\begin{align*}
\lambda_g&= a (\tilde{\lambda}_{g-1} + s -1) +b -a \\
&=M_r \Big(\tilde{\lambda}_{g-1} +m_r - \frac{M_{r+1}}{M_r} \Big) + \sum_{i=2}^{r-1} m_i M_i +M_2+M_1 -M_r \\
&=M_r \tilde{\lambda}_{g-1} +M_r m_r + \sum_{i=2}^{r-1} m_i M_i +M_2+M_1- M_{r+1} -M_r \\
&=\sum_{i=2}^{r} m_i M_i+M_2+M_1 + M_r \tilde{\lambda}_{g-1} - M_{r+1} -M_r \\
&=\sum_{i\geq2} m_i M_i+M_2+M_1
\end{align*}
agreeing with \eqref{eqn2} for $j=g$.
Finally, we compute the remaining entries in the Puisuex characteristic: $\lambda_j$ for $j=1, \dots, g-1$. First we recall that $k_j$ is defined so that $N_j=M_{k_j}$ is the $j$th smallest entry in $der$ which is divisible by the preceding entry. Since this observation applies to the derived vectors of both $(\alpha)$ and $(\beta)$, we find that
\begin{align*}
\tilde{M}_{\tilde{k}_j-1}\ \text{divides}\ \tilde{M}_{\tilde{k}_j} &\Leftrightarrow \frac{M_{r+\tilde{k}_j-2}}{M_r}\ \text{divides} \ \frac{M_{r+\tilde{k}_j-1}}{M_r} \\
&\Rightarrow M_{r+\tilde{k}_j-2}\ \text{divides}\ M_{r+\tilde{k}_j-1} \\
&\Rightarrow M_{r+\tilde{k}_j-1} =M_{k_j} \\
&\Rightarrow r+\tilde{k}_j-1 =k_j.
\end{align*}
Whence we obtain our result \eqref{eqn2}:
\begin{align*}
\lambda_j &= a \tilde{\lambda}_j \\
&= M_r \sum_{i\geq \tilde{k}_j} \tilde{m}_i \tilde{M}_i+\tilde{M}_{\tilde{k}_j} +\tilde{M}_{\tilde{k}_j-1} \\
&= \sum_{i\geq \tilde{k}_j} m_{r+i-1} M_{r+i-1}+ M_{r+\tilde{k}_j-1} + M_{r+\tilde{k}_j-2} \\
&=\sum_{i\geq k_j} m_i M_i+M_{k_j} +M_{k_j-1}
\end{align*}
\begin{flushright}
$\blacksquare$
\end{flushright}
\subsection*{Acknowledgments}
The author would like to thank Richard Montgomery, Piotr Mormul, Wyatt Howard and Alex Castro for helpful discussion, advice, and most importantly, motivation.
|
\section{Introduction}
It has always been accepted that a star cluster (SC) is composed of
stars belonging to a single, simple stellar population with a
uniform age and chemical composition. In recent years, however, due
to the increase in spatial resolution and photometric accuracy, it
has been discovered that some SCs have unusual structures in their
observed color-magnitude diagrams (CMDs). For example, NGC 2173 in
the Large Magellanic Cloud (LMC) has an unusually large spread in
color about its main-sequence turn-off (MSTO) \citep{bert03}; Omega
Centauri exhibits a main sequence (MS) bifurcation \citep{piot05};
NGC 2808 possesses a triple MS split \citep{piot07}; and other
clusters also contain mutiple stellar populations \citep{milo09,
milo10}. More interesting discoveries are that the intermediate-age
massive clusters in the LMC, such as NGC 1846, 1806, and 1783,
possess double main-sequence turn-offs (DMSTOs) on their CMDs
\citep{mack07, mack08, milo09}. Moreover, \cite{goud09} found that
the spread of the MSTO is fairly continuous rather than strongly
bimodal. \cite{mack08} and \cite{goud09} argued that these clusters
do not possess a significant line-of-sigth depth or internal
dispersion in [Fe/H], or suffer from significant differential
extinction. \cite{mucc08} showed that spreads in [Fe/H] in SCs like
NGC 1783 are very small. The apparent homogeneity in [Fe/H] of all
stars in NGC 1846 indicates that capture of pre-existing field stars
during cluster formation seems to hard to explain the DMSTOs
\citep{mack08, goud09}. The DMSTOs are interpreted as that the
clusters have two distinct stellar populations with differences in
age of $\sim$ 200-300 Myr but with similar metal abundance
\citep{mack07, mack08}. Moreover, \cite{glat08} found multiple MSTOs
(MMSTOs) in NGC 419 in the Small Magellanic Cloud, and \cite{gira09}
found two distinct red clumps (RC) in this cluster. These
observational results are challenging the traditional picture.
In order to understand how a cluster can contain multiple
populations with different ages, many scenarios were proposed.
\cite{mack07} put forward the scenario of merger of two (or more)
SCs. However, \cite{goud09} pointed out that this scenario seems
unlikely. \cite{bekk09} simulated the merger of a SC with a giant
molecular cloud (GMC). They found that this merger can produce a
second-generation of stars that are required to explain the DMSTOs.
However, in order to obtain the differences in age of $\sim$ 200 -
300 Myr, this scenario require rather strongly constrained ranges of
GMC parameters such as their spatial distribution, mass function,
and chemical composition \citep{goud09, bast09}. \cite{derc08} and
\cite{goud09} put forward the scenario of formation of a second
generation of stars from the ejecta of first generation asymptotic
giant branch stars. In this scenario, the young second generation is
usually less than the first one \citep{bast09}, which is not
consistent with observations \citep{milo09}. Recently, using two
scaling relations to mimic the effects of rotation, \cite{bast09}
considered the effect of rotation on stellar evolutions. They found
that stellar rotation in stars with masses between 1.2 and 1.7
$\mathrm{M}_\odot${} can mimic the effect of a double population when rotation
rates reach 20-50 per cent of the critical rotation, whereas in
actuality only a single population exists. However, \cite{rube10}
and \cite{gira11} argued that rotational effect could not explain
the presence of MMSTOs. They proposed a prolonged star formation
(PSF) to explain the MMSTOs and the dual RC. In \cite{rube10}
models, in order to reproduce the dual RC of NGC 419, a PSF of about
700 Myr is required. The dual RC was also found in NGC 1751
\citep{rube11}. In order to explain this dual RC, a PSF of about 460
Myr is required \citep{rube11}. This age spread is twice longer than
$\sim$ 200 Myr estimated by \cite{milo09} using the method of
isochrone fitting. Furthermore, recently, \cite{kell11} argued that
the isolation of extended MSTO clusters to intermediate ages could
be the consequence of observational selection effects.
In addition, binary stars of LMC clusters are unresolved even with
Hubble Space Telescope \citep{milo09}. The binary system should
appear as a single point-source object. Thus the unresolved binary
systems may have contributions to the spread of MSTO. However,
several investigators \citep{mack08, goud09, milo09} had found that
unresolved binaries cannot alone reproduce the peculiar MSTO
structures. Nevertheless, after a star in a binary system accretes
mass from its companion, it will become younger apparently. Hence,
interactive binaries may have contributions to the spread of MSTO.
\cite{goud09} found that the younger populations in clusters are
more centrally concentrated than the older populations. Thus the
interactive binaries may be important in interpreting the observed
CMDs. In this paper, using the Hurley rapid single and binary
evolution codes \citep{hurl00,hurl02} and a stellar
population-synthesis method, we investigated the contributions of
interactive binaries to the DMSTOs and dual RC.
The paper is organized as follow. We show our stellar
population-synthesis method in section 2. We present the results in
section 3 and discuss and summarize them in section 4.
\section{Stellar population synthesis}
To investigate DMSTOs and dual RC of SCs, we calculated a
binary-star stellar population (BSP). Stellar samples are generated
by Monte Carlo simulations. The basic assumptions for the
simulations are as follows. (i) The lognormal initial mass function
(IMF) of \cite{chab01} is adopted. (ii) We generate the mass of the
primary, $M_{1}$, according to the IMF. The ratio ($q$) of the mass
of the secondary to that of the primary is assumed to be a uniform
distribution within 0-1 for simplicity. The mass of the secondary
star is then determined by $qM_{1}$. (iii) We assume that all stars
are members of binary systems and that the distribution of
separations ($a$) is constant in $\log a$ for wide binaries and
falls off smoothly at close separation:
\begin{equation}
an(a)=\left\{
\begin{array}{lc}
\alpha_{\rm sep}(a/a_{\rm 0})^{\rm m} & a\leq a_{\rm 0};\\
\alpha_{\rm sep}, & a_{\rm 0}<a<a_{\rm 1},\\
\end{array}\right.
\end{equation}
where $\alpha_{\rm sep}\approx0.070$, $a_{\rm 0}=10R_{\odot}$,
$a_{\rm 1}=5.75\times 10^{\rm 6}R_{\odot}=0.13{\rm pc}$ and
$m\approx1.2$. This distribution implies that the numbers of wide
binary system per logarithmic interval are equal, and that
approximately 50\% of the stellar systems are binary systems with
orbital periods less than 100 yr \citep{han95}. \textbf{However,
binaries account for typically $\sim$30-40\% of all stars in the
clusters of LMC, such as in NGC 1818 and 1806 \citep{elson98,
milo09}.} (iv) The eccentricity (e) of each binary system is assumed
to be a uniform distribution within 0-1. With these assumptions, we
calculated the evolutions of 5$\times 10^{4}$ binaries with $M_{1}$
between 0.8 and 5.0 $\mathrm{M}_\odot${}.
The metal abundance Z of evolutionary models was converted firstly
into [Fe/H]. Then the theoretical properties ([Fe/H], T$_{eff}$,
$\log g$, $\log L$) have been transformed into colors and magnitudes
using the color transformation tables of \cite{leje98}. The binaries
with $a \leq 10^6$ $\mathrm{R}_\odot${} were treated as unresolved ones, while
others were treated as resolved ones when we computed their colors
and magnitudes.
\section{Calculation results}
Figure \ref{fig1} shows the CMDs of the simulated BSP with Z = 0.008
and age = 1.8 Gyr. The interactive binaries are shown in red, while
non-interactive binaries are shown in green. The population of the
brighter MSTO (bMSTO) is mainly from the unresolved binaries with
$q>$ 0.7 and no interactions. The isochrones of single-star stellar
populations (SSP) with the same Z but different ages are overplotted
on the CMD in the right panel of Fig. \ref{fig1}. The stars that
experienced binary interactions clearly deviate from the isochrone
of non-interactive binaries in the region around MSTO, which leads
to a large spread in color. The interactive binaries clearly produce
an extended MSTO rather than bimodal MSTO. For main sequence
\textbf{binary} stars with $\mathrm{m_{V}} <$ 22.5 mag,
\textbf{about 10\% of the binary stars lie between the 1.8 and 1.5
Gyr isochrones in our models, not including the binaries on the 1.8
Gyr isochrone}. However, interactive binaries hardly affect the rest
of MS, subgiant branch and first giant branch (FGB). Moreover,
interactive binaries result in the appearance of a secondary RC
(SRC) at about 0.5 mag below the main RC. The SRC is slightly bluer
than the ridgeline of FGB stars. The contribution of non-interactive
binaries to the SRC is negligible. The right panel of Fig.
\ref{fig1} shows that the ridge of the SRC is almost coincident with
the isochrone of RC stars of the SSP with age = 1.2 Gyr. About 12\%
of core-helium burning (CHeB) \textbf{binary} stars are located
between the 1.8 and 1.5 Gyr isochrones. However, about 30\% of CHeB
\textbf{binary} stars are located between the main RC and 1.2 Gyr
isochrone, not including \textbf{the binary} stars on the main RC
but including those on 1.2 Gyr isochrone. The mass of the SRC stars
is mainly located between 1.85 and 2.1 $\mathrm{M}_\odot${}, while that of the
main RC stars is less than 1.7 $\mathrm{M}_\odot${}.
The dual RC was discovered in NGC 752, 7789 and 419 \citep{gira00,
gira09} and NGC 1751 \citep{rube11}. Cluster NGC 419 has an age of
about 1.35 Gyr \citep{gira09}. We computed a cluster with age = 1.3
Gyr and Z = 0.004. Its CMDs are shown in Fig. \ref{fig2}. Just as
above results, the interactive binaries lead to a large spread in
color in the region around MSTO, but they do not reproduce DMSTOs.
However, the interactive binaries lead to the presence of a distinct
SRC. The ridge of the SRC is almost coincident with the isochrone of
RC stars of the SSP with age = 1.1 Gyr. This implies that the SRC
produced by binary interactions has an apparent age of about 1.1 Gyr
in our models, whereas actually it has an age of 1.3 Gyr.
\textbf{For this dual RC, about 15\% of binary stars} are located
between the 1.3 and 1.1 Gyr isochrones, including the
\textbf{binaries} on the 1.1 Gyr isochrone but not including those
on 1.3 Gyr isochrone. The magnitude extension between main RC and
SRC is about 0.35 mag. The mass of the SRC stars is mainly located
between 1.85 and 2.05 $\mathrm{M}_\odot${}, while that of the main RC stars is
mainly located between 1.78 and 1.81 $\mathrm{M}_\odot${}. For Z = 0.004, in stars
more massive than about 1.88 $\mathrm{M}_\odot${} (this value is about 2.0 $\mathrm{M}_\odot${}
for Z = 0.02), helium-burning temperatures are reached at the center
before electrons become degenerate there. For these stars, when they
are on zero-age horizontal branch (ZAHB), their magnitudes decrease
with decreasing mass. However, for the stars with $M <$ 1.88 $\mathrm{M}_\odot${},
electrons in the hydrogen-exhausted core are highly degenerate
before helium ignition occurs. When they are on ZAHB, their
magnitudes increase rapidly with decreasing mass except for stars
with $M <$ 1.5 $\mathrm{M}_\odot${}. The magnitudes of ZAHB stars with $M <$ 1.5
$\mathrm{M}_\odot${} increase very slowly with decreasing mass. For Z = 0.004, the
evolutionary tracks of RC stars with mass between about 1.85 and 2.0
$\mathrm{M}_\odot${} are very near each other in CMDs. Thus the interactive
binaries with mass between about 1.85 and 2.0 $\mathrm{M}_\odot${} can gather to
form a SRC. Our calculations show that the SRC caused by binary
interactions appears mainly in clusters with 1.2 Gyr $<$ age $<$ 3.0
Gyr.
\section{Discussion and Conclusions}
\cite{milo09} showed that the fraction of younger (brighter)
population is about 75\% in the case of NGC 1846. In our models, the
`younger' population are mainly from the merged binary systems and
interactive binaries with mass transfer. The merging and mass
transfer in binary systems can be affected by the distributions of
separation ($a$), eccentricity ($e$) and mass-ratio ($q$) of
systems. We computed the BSP by using the distributions $n(\log a)$
= constant \citep{hurl02}, $n(e) = 2e$ and $n(q) = 2q$ and the IMF
of \cite{salp55}. However, the fraction of the interactive binaries
are not obviously enhanced. We also computed the BSP using the
distribution $n(\log a)$ = constant and Gaussian distributions for
$e$ and $q$. The mean value and standard deviation of the Gaussian
distributions is 0.5 and 0.13 for $e$, 0.6 and 0.1 for $q$,
respectively. The values of the standard deviation are chosen in
order to make the values of $e$ and $q$ are located between 0 and 1.
In this simulation, \textbf{about 20\% of binary stars lie between
the 1.8 and 1.5 Gyr isochrones for the cluster with Z = 0.008 and
age = 1.8 Gyr; while for the RC of the cluster with Z = 0.004 and
age = 1.3 Gyr, about 30\% of binary stars lie between the 1.3 and
1.1 Gyr isochrones.}
The mass of turn-off stars is about 1.46 $\mathrm{M}_\odot${} for the cluster with
Z = 0.008 and age = 1.8 Gyr in our models. The evolutions of a wide
binary system without mass transfer are similar to those of single
stars with the same mass. However, the evolution of stars whose mass
increased via mass accretion or merging is always slower than that
of single stars with the same mass. For example, a mass-accreted
star with $M$ = 1.55 $\mathrm{M}_\odot${}, which was reached by accretion, just
arrives at around MSTO at the age of 1.8 Gyr. However, a single star
with an initial mass of 1.55 $\mathrm{M}_\odot${} has left the MSTO at the same
age. Thus the mass-accreted stars clearly deviate from the ridge of
the isochrone in the region around MSTO and have an apparent younger
age. The more the accreted mass, the more the deviation. Thus the
spread of MSTO of BSP is continuous rather than bimodal. However,
for the mass-accreted stars with $M <$ 1.25 $\mathrm{M}_\odot${}, an increase in
mass leads to a shift of their position in CMDs almost along MS.
Therefore, the most of MS are hardly spread by binary interactions.
The SRC is almost made of the mass-accreted or merged stars whose
mass is slightly more massive than the critical mass that is just
enough to avoid electron-degeneracy occurring in stellar H-exhausted
cores in our models. The stars with the critical mass usually evolve
into RC at the age of about 1.1-1.2 Gyr (For the stars with $Z \geq
0.04$, this age can increase to about 1.3 Gyr). Hence, the SRC
caused by interactive binaries should have an apparent age of about
1.1-1.2 Gyr, whereas actually it has an age as old as that of main
RC. The apparent age of about 1.1-1.2 Gyr is consistent with the
estimates of \cite{rube11, rube10} for NGC 1751 and 419. The main RC
stars passed through electron degeneracy. The magnitude of these
stars increases with decreasing mass. The older the age of clusters,
the lower the mass of main RC stars. Thus, the magnitude extension
between main RC and SRC increases with age until the mass of the
main RC is less than about 1.5 $\mathrm{M}_\odot${}. Because the magnitude of ZAHB
stars with $M <$ 1.5 $\mathrm{M}_\odot${} increases very slowly with decreasing
mass, there is an upper limit of magnitude extension of about 0.8
mag between main RC and SRC in our models. Moreover, with increasing
age, the SRC becomes slightly bluer than FGB. For the clusters with
age $<$ 1.2 Gyr, the mass of all RC stars is almost larger than the
critical mass. Therefore, the SRC caused by interactive binaries
should not appear in these clusters. In addition, in old clusters,
the initial mass of RC stars is low. If the mass is not enough to be
increased to above the critical mass by accretion or merging, the
SRC should not be produced by binary interactions in the old
clusters. Moreover, if the fraction of binaries is too low, the SRC
also could not be produced. \cite{gira09} argued that the MS+red
clump binaries cannot mimic the dual RC in NGC 419. In our models,
the SRC stars are from the merged binaries and binary systems with
mass transfer.
In this paper, we showed that binary interactions such as mass
transfer and binary merging can produce an extended MSTO and dual RC
in CMDs of intermediate-age clusters, whereas in actuality only a
single population exists. Despite these, the rest of MS, subgiant
branch and FGB are not clearly spread by the binary interactions.
Interactive binaries can lead to an extension of MSTO rather than
DMSTOs. For a cluster with Z = 0.008 and age = 1.8 Gyr, its
isochrone can be spread down to $\sim$ 1.5 Gyr by interactive
binaries. However, \textbf{only about 10\% of binary stars lie
between the 1.8 and 1.5 Gyr isochrones} in our models. Moreover, the
SRC that is caused by interactive binaries should have an apparent
age of about 1.1-1.2 Gyr. Although binary interactions cannot lead
to the bimodal MSTO, the SRC in NGC 419 and extended MSTO of NGC
1846 may be partly from the mass-accreted or merged binary stars.
\acknowledgments We thank the anonymous referee for his/her helpful
comments and acknowledge support from the CPSF 20100480222, NSFC
11003003, 10773003, 10933002 and the Ministry of Science and
Technology of the People's republic of China through grant
2007CB815406.
|
\section{Introduction}
In this paper we present a general approach to structured diagram spaces and the relation to symmetric spectra. We begin by discussing the motivating examples of ${\mathcal I}$- and ${\mathcal J}$-spaces. Here the underlying category of spaces ${\mathcal S}$ may be interpreted either as the category of (compactly generated weak Hausdorff) topological spaces or the category of simplicial sets.
\subsection{${\mathcal I}$-spaces and $E_{\infty}$ spaces}
Let ${\mathcal I}$ be the category whose objects are the finite sets
$\bld n=\{1,\dots,n\}$ (including the empty set $\bld 0$) and whose morphisms are the injective maps. The usual ordered concatenation of ordered sets makes ${\mathcal I}$ a symmetric monoidal category. An ${\mathcal I}$-space is a functor $X\colon {\mathcal I}\to{\mathcal S}$ and we write ${\mathcal S}^{{\mathcal I}}$ for the category of ${\mathcal I}$-spaces. As it is generally the case for a diagram category indexed by a small symmetric monoidal category, ${\mathcal S}^{{\mathcal I}}$ inherits a symmetric monoidal structure from
${\mathcal I}$. A map of ${\mathcal I}$-spaces $X\to Y$ is said to be an
\emph{${\mathcal I}$-equivalence} if the induced map of homotopy colimits
$X_{h{\mathcal I}}\to Y_{h{\mathcal I}}$ is a weak homotopy equivalence. The
${\mathcal I}$-equivalences are the weak equivalences in a model structure on
${\mathcal S}^{{\mathcal I}}$, the \emph{projective ${\mathcal I}$-model structure}, with the property that the usual adjunction
\[
\xymatrix{
\colim_{{\mathcal I}} \colon {\mathcal S}^{{\mathcal I}} \ar@<0.5ex>[r] &
{\mathcal S} \thinspace \colon \! \! \const_{{\mathcal I}} \ar@<0.5ex>[l]
}
\]
defines a Quillen equivalence with respect to the standard model structure on ${\mathcal S}$. Thus, the homotopy category of ${\mathcal S}^{{\mathcal I}}$ is equivalent to the usual homotopy category of spaces. We think of $X_{h{\mathcal I}}$ as the underlying space of the ${\mathcal I}$-space $X$.
The main advantage of ${\mathcal S}^{{\mathcal I}}$ compared to ${\mathcal S}$ is that it provides a more flexible setting for working with structured objects, in particular $E_{\infty}$ structures. Let ${\mathcal C}{\mathcal S}^{{\mathcal I}}$ be the category of commutative
${\mathcal I}$-space monoids, that is, commutative monoids in ${\mathcal S}^{{\mathcal I}}$. Utilizing an idea of Jeff Smith first implemented in the category of symmetric spectra \cite{MMSS} we show that there is a \emph{positive projective ${\mathcal I}$-model structure} on ${\mathcal S}^{{\mathcal I}}$ that lifts to a model structure on ${\mathcal C}{\mathcal S}^{{\mathcal I}}$. In the following theorem we consider an $E_{\infty}$ operad ${\mathcal D}$ with associated monad $\mathbb D$ and ${\mathcal S}[\mathbb D]$ denotes the category of $\mathbb D$-algebras in ${\mathcal S}$.
\begin{theorem}\label{thm:intro-I-theorem}
Let ${\mathcal D}$ be an $E_{\infty}$ operad. Then there is a chain of Quillen equivalences ${\mathcal C}{\mathcal S}^{{\mathcal I}}\simeq{\mathcal S}[\mathbb D]$ relating the positive projective ${\mathcal I}$-model structure on ${\mathcal C}{\mathcal S}^{{\mathcal I}}$ to the standard model structure on
${\mathcal S}[\mathbb D]$ lifted from ${\mathcal S}$.
\end{theorem}
In the topological setting this result applies for instance to the little $\infty$-cubes operad and the linear isometries operad. Specializing to the Barratt-Eccles operad ${\mathcal E}$ we can give an explicit description of the induced equivalence of homotopy categories: A commutative ${\mathcal I}$-space monoid $A$ is mapped to the homotopy colimit $A_{h\mathcal I}$ with its canonical action of the monad $\mathbb E$ associated to ${\mathcal E}$, see
\cite[Proposition~6.5]{Schlichtkrull-Thom_symm}.
As a consequence of the theorem any $E_{\infty}$ homotopy type can be represented by a commutative ${\mathcal I}$-space monoid. It is well-known that in general such a rectification cannot be carried out in ${\mathcal S}$ since a grouplike commutative monoid in ${\mathcal S}$ is equivalent to a product of
Eilenberg-Mac\ Lane spaces.
\begin{example}\label{ex:X-bullet-example}
A based space $X$ gives rise to a commutative ${\mathcal I}$-space monoid $X^{\bullet}\colon \bld n\mapsto X^n$ and it is proved in \cite{Schlichtkrull-infinite} that the underlying space $X^{\bullet}_{h{\mathcal I}}$ is equivalent to the Barratt-Eccles construction $\Gamma^+(X)$ (in the topological setting we need the extra assumption that $X$ be well-based). Thus, for connected $X$ it follows from the Barratt-Priddy-Quillen Theorem that $X^{\bullet}$ represents the infinite loop space $Q(X)$.
\end{example}
\begin{example}\label{ex:BO-example}
Writing $O(n)$ for the orthogonal groups we have the commutative ${\mathcal I}$-space monoid $BO\colon\bld n\mapsto BO(n)$ in the topological setting. In this case $BO_{h{\mathcal I}}$ is equivalent to the classifying space $BO(\infty)$ for stable vector bundles.
Thus, we have represented the $E_{\infty}$ space $BO(\infty)$ as a commutative ${\mathcal I}$-space monoid.
\end{example}
\begin{example}\label{ex:BGL(R)-example}
Let $R$ be an ordinary ring and consider the commutative ${\mathcal I}$-space monoid
$B\mathrm{GL}(R)\colon \bld n\mapsto B\mathrm{GL}_n(R)$. In this case the underlying space $B\mathrm{GL}(R)_{h{\mathcal I}}$ is equivalent to Quillen's plus construction
$B\mathrm{GL}_{\infty}(R)^+$. Thus, we have represented the higher algebraic
$K$-theory of $R$ by a commutative $\mathcal I$-space monoid.
\end{example}
For a commutative ${\mathcal I}$-space monoid $A$, we know from \cite[Theorem 4.1]{Schwede_S-algebras} that the projective model structure on ${\mathcal S}^{{\mathcal I}}$ lifts to a monoidal model structure on the category
of $A$-modules with the symmetric monoidal product inherited from
${\mathcal S}^{{\mathcal I}}$. This symmetric monoidal structure on $A$-modules is one of the benefits of working with a strictly commutative monoid and is hard to come by in the operadic context.
The category of ${\mathcal I}$-spaces is closely related to the category of symmetric spectra ${\mathrm{\textrm{Sp}^{\Sigma}}}$ and in particular we have an adjunction
${\mathbb S}^{{\mathcal I}}[-] \colon{\mathcal S}^{{\mathcal I}} \rightleftarrows {\mathrm{\textrm{Sp}^{\Sigma}}} \colon\Omega^{{\mathcal I}}$,
where $\mathbb S^{{\mathcal I}}[-]$ is an ${\mathcal I}$-space version of the (unbased) suspension spectrum functor and $\Omega^{{\mathcal I}}$ takes a symmetric spectrum $E$ to the $\mathcal I$-space $\Omega^{{\mathcal I}}(E)\colon\bld n\mapsto \Omega^n(E_n)$.
As demonstrated in \cite{Schlichtkrull-units} this can be used to give an ${\mathcal I}$-space model $\GL^{\cI}_1(R)$ for the units of a symmetric ring spectrum $R$. In this paper we analyze the homotopical properties of these constructions and we lay the foundation for the applications in \cite{Rognes-TLS} to the theory of topological logarithmic structures. Commutative ${\mathcal I}$-space monoids as models for $E_{\infty}$ spaces have also proved useful for the study of algebraic K-theory of
structured ring spectra~\cite{Schlichtkrull-units}, Thom
spectra~\cite{Schlichtkrull-Thom_symm}, and the topological Hochschild
homology of Thom spectra~\cite{Blumberg-C-S_THH-Thom,Schlichtkrull-higher-THH}. In particular, the present paper provides results which were referred to (and, in some cases, also used) in
\cite{Blumberg-C-S_THH-Thom,Rognes-TLS,
Schlichtkrull-Thom_symm,Schlichtkrull-higher-THH}.
In~\cite{Sagave-S_group-completion}, the authors examine group completion of commutative ${\mathcal I}$-space monoids, express it in terms of model structures, relate it to the notion of units, and explain the
connection to $\Gamma$-spaces.
It is in order to relate this work to the framework for structured spectra developed by Elmendorf, Kriz, Mandell, and May \cite{EKMM}. In this framework the analogues of symmetric spectra are the so-called \emph{$S$-modules}. Just as one may view ${\mathcal I}$-spaces as space level analogues of symmetric spectra, there is a space level analogue of $S$-modules known as \emph{$*$-modules}, see
\cite[Section 4]{Blumberg-C-S_THH-Thom}. It is proved by Lind \cite{Lind-diagram} that the $(\mathbb S^{{\mathcal I}}[-],\Omega^{{\mathcal I}})$-adjunction discussed above has an $S$-module analogue which on the level of homotopy categories agrees with the latter up to natural isomorphism. Furthermore, Lind goes on to establish a Quillen equivalence between the category ${\mathcal C}{\mathcal S}^{{\mathcal I}}$ and the category of commutative monoids in $*$-modules which in turn can be identified with algebras for the linear isometries operad. This gives a way to relate commutative ${\mathcal I}$-space monoids with $E_{\infty}$ spaces which is quite different from the approach taken in this paper.
\subsection{${\mathcal J}$-spaces and graded units}
Whereas the ${\mathcal I}$-space monoid $\GL^{\cI}_1(R)$ is a useful model for the units of a connective symmetric ring spectrum, this construction is of limited value for symmetric ring spectra that are not connective: If $R\to R'$ is a map of positive fibrant symmetric ring spectra which induces an isomorphism of homotopy groups in non-negative degrees, then the induced map $\GL^{\cI}_1(R)\to \GL^{\cI}_1(R')$ is an
${\mathcal I}$-equivalence. Consequently, the ${\mathcal I}$-space units do not distinguish between a symmetric ring spectrum and its connective cover and
cannot detect periodicity phenomena in stable homotopy. This defect of the
${\mathcal I}$-space units is shared by any other previous definitions of the units of a structured ring spectrum.
What we seek instead is a notion of the units which takes into account all the possible stable $R$-module equivalences $\Sigma^{n_2}R\to \Sigma^{n_1}R$ between suspended copies of $R$. Motivated by the definition of
$\Omega^{{\mathcal I}}(R)$ it is natural to try organizing the collection of spaces
$\Omega^{n_2}(R_{n_1})$ into a ${\mathcal J}$-diagram for a suitable small symmetric monoidal category ${\mathcal J}$. One of the main features of the paper is to show that Quillen's localization construction $\Sigma^{-1}\Sigma$ on the category of finite sets and bijections $\Sigma$ is a natural choice for such a category ${\mathcal J}$.
The objects of $\Sigma^{-1}\Sigma$ are pairs $(\bld n_1,\bld n_2)$ of objects in ${\mathcal I}$ and the morphisms $(\bld m_1,\bld m_2)\to(\bld n_1,\bld n_2)$ can be described explicitly as triples $(\alpha_1,\alpha_2,\rho)$ given by a pair of morphisms $\alpha_1\colon\bld m_1\to\bld n_1$ and $\alpha_2\colon\bld m_2\to\bld n_2$ in ${\mathcal I}$, together with a bijection $\rho$ identifying the complements of the images of these morphisms. The point is that the extra connecting tissue provided by $\rho$ is exactly the data needed to make the correspondence $(\bld n_1,\bld n_2)\mapsto \Omega^{n_2}(R_{n_1})$ functorial.
With this choice of ${\mathcal J}$ we define a \emph{${\mathcal J}$-space} to be a functor $X\colon{\mathcal J}\to{\mathcal S}$ and write ${\mathcal S}^{{\mathcal J}}$ for the category of ${\mathcal J}$-spaces. A map of ${\mathcal J}$-spaces $X\to Y$ is said to be a \emph{${\mathcal J}$-equivalence} if the induced map of homotopy colimits $X_{h{\mathcal J}}\to Y_{h{\mathcal J}}$ is a weak homotopy equivalence. We show that the ${\mathcal J}$-equivalences are the weak equivalences in a model structure on ${\mathcal S}^{{\mathcal J}}$, the \emph{projective ${\mathcal J}$-model structure}, with the property that there is a chain of Quillen equivalences
${\mathcal S}^{{\mathcal J}}\simeq {\mathcal S}/B{\mathcal J}$ relating it to the standard model structure on the category ${\mathcal S}/B{\mathcal J}$ of spaces over $B{\mathcal J}$. There also is a \emph{positive projective ${\mathcal J}$-model structure} on ${\mathcal S}^{{\mathcal J}}$ which lifts to a model structure on the category ${\mathcal C}{\mathcal S}^{{\mathcal J}}$ of commutative ${\mathcal J}$-space monoids (that is, commutative monoids in ${\mathcal S}^{{\mathcal J}}$). In the next theorem
$\mathbb E$ again denotes the monad associated to the Barratt-Eccles operad and ${\mathcal S}[\mathbb E]/B{\mathcal J}$ is the category of $\mathbb E$-algebras over the $\mathbb E$-algebra $B{\mathcal J}$.
\begin{theorem}\label{thm:intro-J-theorem}
There is a chain of Quillen equivalences ${\mathcal C}{\mathcal S}^{{\mathcal J}}\simeq {\mathcal S}[\mathbb E]/B{\mathcal J}$ relating the positive projective model structure on ${\mathcal C}{\mathcal S}^{{\mathcal J}}$ to the standard model structure on ${\mathcal S}[\mathbb E]/B{\mathcal J}$ lifted from ${\mathcal S}$.
\end{theorem}
By work of Barratt, Priddy, and Quillen, it is known that $B{\mathcal J}$ is equivalent to $Q(S^0)$, so the above theorem allows us to interpret ${\mathcal C}{\mathcal S}^{{\mathcal J}}$ as a model for the category of $E_{\infty}$ spaces over $Q(S^0)$. This fits well with the general point of view that in a spectral context the sphere spectrum $\mathbb S$ takes the role played by the ring of integers $\mathbb Z$ in the traditional algebraic context. Indeed, in algebra a graded monoid is logically the same as a monoid $A$ together with a monoid homomorphism
$A\to \mathbb Z$ to the underlying additive group $(\mathbb Z,+,0)$. In topology it is customary to think of $Q(S^0)$ as the ``additive group'' of
$\mathbb S$ and hence we can think of commutative ${\mathcal J}$-space monoids as representing graded commutative spaces.
The category of ${\mathcal J}$-spaces is related to the category of symmetric spectra by a pair of monoidal adjoint functors
${\mathbb S}^{{\mathcal J}}[-] \colon{\mathcal S}^{{\mathcal J}} \rightleftarrows {\mathrm{\textrm{Sp}^{\Sigma}}} \colon\Omega^{{\mathcal J}}$.
Given a symmetric ring spectrum $R$ we define the graded units $\GL^{\cJ}_1(R)$ as a suitable sub ${\mathcal J}$-space monoid of
$\Omega^{{\mathcal J}}(R)$. In general, we define for any positive fibrant
${\mathcal J}$\nobreakdash-space monoid $A$ a graded signed monoid $\pi_0(A)$ of ``components'' (see Definition \ref{def:graded-signed-monoid}). The definition of $\pi_0(A)$ is motivated by the next theorem where $\pi_*(R)^{\times}$ denotes the graded group of multiplicative units in the graded ring of homotopy groups $\pi_*(R)$.
\begin{theorem}
Let $R$ be a positive fibrant symmetric ring spectrum. Then there is a natural isomorphism of graded signed monoids $\pi_0(\Omega^{{\mathcal J}}(R))\simeq \pi_*(R)$ which restricts to an isomorphism of graded signed groups
$\pi_0(\GL^{\cJ}_1(R))\simeq \pi_*(R)^{\times}$.
\end{theorem}
Hence all units of the graded ring $\pi_*(R)$ are incorporated in the graded units $\GL^{\cJ}_1(R)$ while the corresponding notions of units using ${\mathcal I}$-space monoids or $E_{\infty}$ spaces only detect $\pi_0(R)^{\times}$. We illustrate the use of these concepts by applying them to the theory of topological logarithmic structures developed by Rognes \cite{Rognes-TLS}. The question of how to associate spectra with grouplike commutative ${\mathcal J}$-space monoids like $\GL^{\cJ}_1(R)$ and how to form group completions in this setting is studied by the first author in~\cite{Sagave_spectra-of-units}.
\subsection{Well-structured index categories}
In order to treat the theories of ${\mathcal I}$- and ${\mathcal J}$-spaces in a common framework and to express our results in the natural level of generality, we introduce the notion of a \emph{well-structured index category}. Such a category ${\mathcal K}$ is a small symmetric monoidal category equipped with a degree functor to the ordered set of natural numbers and satisfying a short list of axioms, see Definition~\ref{def:well-structured-index}. The axioms guarantee that the associated category of ${\mathcal K}$-spaces ${\mathcal S}^{{\mathcal K}}$ inherits a well-behaved \emph{projective ${\mathcal K}$-model structure} whose weak equivalences are the
${\mathcal K}$-equivalences, that is, the maps that induce weak homotopy equivalences of the associated homotopy colimits. Here ``well-behaved'' means that the projective ${\mathcal K}$-model structure is cofibrantly generated, proper, monoidal, and satisfies the monoid axiom.
Assuming that the full subcategory ${\mathcal K}_+$ of ${\mathcal K}$ whose objects have positive degree is homotopy cofinal, there also is a \emph{positive projective
${\mathcal K}$-model structure} on ${\mathcal S}^{{\mathcal K}}$. The latter model structure lifts to the category of commutative ${\mathcal K}$-space monoids ${\mathcal C}{\mathcal S}^{{\mathcal K}}$ provided that
for each pair of objects $\bld k$ and $\bld l$ in ${\mathcal K}_+$, the action of the symmetric group $\Sigma_n$ on the $n$-fold iterated monoidal product $\bld k^{\sqcup n}$ induces a free right action on the set of connected components of the comma category $(\bld k^{\sqcup n}\sqcup-\downarrow \bld l)$. We express this by saying that the discrete subcategory $O{\mathcal K}_+$ of identity morphisms with positive degree defines a \emph{very well-structured relative index category $({\mathcal K},O{\mathcal K}_+)$}.
The categories ${\mathcal I}$ and ${\mathcal J}$ are well-structured index categories and the next theorem generalizes
Theorem~\ref{thm:intro-I-theorem} and Theorem~\ref{thm:intro-J-theorem}. Here $\mathbb E$ again denotes the monad associated to the Barratt-Eccles operad. If the symmetric monoidal category ${\mathcal K}$ is permutative (that is, symmetric strict monoidal), then $B{\mathcal K}$ is an $\mathbb E$-algebra and we write ${\mathcal S}[\mathbb E]/B{\mathcal K}$ for the category of $\mathbb E$-algebras over $B{\mathcal K}$.
\begin{theorem}\label{thm:intro-K-theorem}
Let ${\mathcal K}$ be a well-structured index category.
\begin{enumerate}[(i)]
\item
There is a chain of Quillen equivalences ${\mathcal S}^{{\mathcal K}}\simeq {\mathcal S}/B{\mathcal K}$ relating the projective ${\mathcal K}$-model structure on ${\mathcal S}^{{\mathcal K}}$ to the standard model structure on ${\mathcal S}/B{\mathcal K}$.
\item
Suppose that the underlying symmetric monoidal category of ${\mathcal K}$ is permutative, that ${\mathcal K}_+$ is homotopy cofinal in ${\mathcal K}$, and that
$({\mathcal K},O{\mathcal K}_+)$ is very well-structured. Then there is a chain of Quillen equivalences
${\mathcal C}{\mathcal S}^{{\mathcal K}}\simeq {\mathcal S}[\mathbb E]/B{\mathcal K}$ relating the positive projective
${\mathcal K}$-model structure on ${\mathcal C}{\mathcal S}^{{\mathcal K}}$ to the standard model structure on
${\mathcal S}[\mathbb E]/B{\mathcal K}$.
\end{enumerate}
\end{theorem}
There are many examples of well-structured index categories and in each case one may view the above theorem as a kind of rectification principle. We give a further example related to algebraic $K$-theory. Let $R$ be an ordinary algebraic ring with invariant basis number (for instance any commutative ring with more than one element), let ${\mathcal F}_R$ be the isomorphism category of $R$-modules of the standard form $R^n$, and let ${\mathcal K}_R$ be Quillen's localization construction ${\mathcal F}_R^{-1}{\mathcal F}_R$ on this category. The classifying space $B{\mathcal K}_R$ represents the ``free'' algebraic $K$-theory functor which is equivalent to the ordinary algebraic $K$-theory functor in positive degrees. We refer the reader to Example~\ref{ex:K-R-example} for a full discussion and the verification that ${\mathcal K}_R$ defines a well-structured index category such that the conditions in
Theorem~\ref{thm:intro-K-theorem} are satisfied. Applied in this case the latter theorem provides a rectification of $K$-theoretical data:
Each homotopy class of maps $X\to B{\mathcal K}_R$ can be represented by a unique ${\mathcal K}_R$-space homotopy type, and each homotopy class of $E_{\infty}$ maps can be represented by a unique commutative ${\mathcal K}_R$-space monoid homotopy type.
For any serious work with diagram spaces it is important to understand
the homotopical properties of the monoidal structure. One may ensure
that a ${\mathcal K}$-space is homotopically well-behaved with respect to the
monoidal product by imposing suitable cofibrancy conditions and in
practice it often happens that there are several useful model
structures on the same diagram category ${\mathcal S}^{{\mathcal K}}$ providing such
notions of cofibrancy. This is analogous to the situation for
symmetric spectra where the stable projective model structure
\cite{HSS,MMSS} is accompanied by the stable flat (or $S$-) model
structure \cite{HSS, Shipley-convenient} and the corresponding
positive variants \cite{MMSS, Shipley-convenient}; see also
\cite{Schwede-SymSp}. In this paper we set up a general framework for
analyzing model structures on diagram categories by introducing the
notion of a \emph{well-structured relative index category} $({\mathcal K},{\mathcal A})$
given by a small symmetric monoidal category ${\mathcal K}$ together with a
subcategory of automorphisms ${\mathcal A}$ satisfying a suitable list of
axioms, see Definition~\ref{def:well-structured-relative-index}.
Letting ${\mathcal A}$ be the discrete category of identity morphisms in ${\mathcal K}$
we recover the notion of a well-structured index category and the
corresponding projective ${\mathcal K}$-model structure. Similarly, if we let
${\mathcal A}$ be the category of identity morphisms of positive degree we get
the positive projective ${\mathcal K}$-model structure. If in the case of ${\mathcal I}$
and ${\mathcal J}$ we let ${\mathcal A}$ be the subcategory of all (positive)
automorphisms, we get the analogues of the stable (positive) flat
model structure on symmetric spectra. This diversity might be
confusing at first sight, but there are useful features of each of
these model structures and no single one has all the desirable
properties simultaneously.
\subsection{Organization}
The paper is roughly divided into two parts: In Sections 2--4 we present the theory of ${\mathcal I}$- and ${\mathcal J}$-spaces in detail and we show how the graded
units $\GL^{\cJ}_1(R)$ can be used in connection with the theory of topological logarithmic structures. Many of the proofs in this first part of the paper are deferred to the remaining Sections 5--14 where we develop the homotopical properties in the general framework of diagram spaces indexed by a well-structured relative index category. The technical results on operad algebras needed for the paper are established in Appendix \ref{app:analysis-operad-algebras}. It is hoped that by first presenting the applications to ${\mathcal I}$- and ${\mathcal J}$-spaces, the reader will be motivated to go through the more general material in the second part of the paper. The specific organization of the material should be clear from the table of contents.
\tableofcontents
\subsection{Acknowledgments}
This project benefited from support through the YFF grant \emph{Brave
new rings} held by John Rognes and the \emph{Topology in Norway}
project, both funded by the Research Council of Norway, the SFB 478
at M\"unster and the HCM at Bonn. The authors like to thank Ruth
Joachimi, Thomas Kragh, Wolfgang L\"uck, John Rognes, Stefan Schwede,
Mirjam Solberg, and Rainer Vogt for helpful conversations related to
this project. Moreover, the suggestions made by an anonymous referee also helped to improve the manuscript.
\section{Preliminaries on diagram spaces}\label{sec:preliminaries}
We work simultaneously in a topological and a simplicial setting and write
${\mathcal S}$ for our category of spaces. Thus, unless stated otherwise, ${\mathcal S}$ denotes both the category of compactly generated weak Hausdorff topological spaces and the category of simplicial sets. The corresponding based categories are denoted by ${\mathcal S}_*$.
\begin{definition}
Given a small category ${\mathcal K}$, a \emph{${\mathcal K}$-space} is a functor
$X\colon{\mathcal K}\to{\mathcal S}$. We write ${\mathcal S}^{{\mathcal K}}$ for the category of ${\mathcal K}$-spaces
with morphisms the natural transformations.
\end{definition}
The next lemma recalls the basic formal properties of the category of
${\mathcal K}$-spaces.
\begin{lemma}\label{lem:K-spaces-S-category}
The category ${\mathcal S}^{{\mathcal K}}$ is bicomplete with limits and colimits constructed levelwise. Furthermore, ${\mathcal S}^{{\mathcal K}}$ is
enriched, tensored, and cotensored over
${\mathcal S}$. For a ${\mathcal K}$-space $X$ and a space $T$ (in ${\mathcal S}$), the
tensor $X \times T$ and cotensor $X^T$ are the ${\mathcal K}$-spaces defined by
\[
(X\times T)(\bld{k}) =X(\bld{k})\times T\quad\textrm{ and }\quad
X^T(\bld k)=\Map(T,X(\bld k)).
\]
The space of maps from $X$
to $Y$ is the end
\[ \Map(X,Y) = \int_{\bld{k}\in{\mathcal K}} \Map(X(\bld{k}),
Y(\bld{k})).
\eqno\qed
\]
\end{lemma}
Suppose now that $({\mathcal K},\sqcup,\bld 0)$ is a symmetric monoidal category with monoidal structure $\sqcup$ and unit $\bld 0$.
The left Kan extension along $\concat \colon {\mathcal K} \times {\mathcal K} \to {\mathcal K}$
defines a symmetric monoidal product $\boxtimes$ on ${\mathcal K}$-spaces in the usual way: Given a pair of ${\mathcal K}$-spaces $X$ and $Y$,
\[ (X \boxtimes Y)(\bld{n}) = \colim_{\bld{k}\concat \bld{l} \to
\bld{n}} X(\bld{k}) \times Y(\bld{l})\] with the colimit taken over
the comma category $(\concat \downarrow {\mathcal K})$. The monoidal unit is the levelwise discrete ${\mathcal K}$-space
$\bld 1_{{\mathcal K}}={\mathcal K}(\bld{0},-)$.
\begin{definition}
A \emph{(commutative) ${\mathcal K}$-space monoid} is a (commutative) monoid in
the symmetric monoidal category of ${\mathcal K}$-spaces $({\mathcal S}^{{\mathcal K}},\boxtimes,
\bld 1_{{\mathcal K}})$. We write
${\mathcal C}{\mathcal S}^{{\mathcal K}}$ for the category of commutative ${\mathcal K}$-space monoids.
\end{definition}
By the universal property of the left Kan extension, the data defining a monoid structure on a ${\mathcal K}$-space $A$ amounts to a map $*\to A(\bld 0)$ and a map of $({\mathcal K}\times{\mathcal K})$-spaces $A(\bld k)\times A(\bld l)\to A(\bld k\sqcup\bld l)$, subject to the usual associativity and unitality conditions. The commutativity condition amounts to the commutativity of the diagram of $({\mathcal K}\times{\mathcal K})$-spaces
\[
\xymatrix@-1pc{
A(\bld k)\times A(\bld l) \ar[r] \ar[d]& A(\bld k\sqcup \bld l)\ar[d]\\
A(\bld l)\times A(\bld k) \ar[r] & A(\bld l\sqcup \bld k)
}
\]
where the left hand side flips the factors and the right hand map is induced by the symmetry isomorphism of ${\mathcal K}$. An equivalent way of expressing this is to say that $A$ defines a (lax) symmetric monoidal functor from $({\mathcal K},\sqcup,\bld 0)$ to $({\mathcal S},\times,*)$.
The symmetric monoidal structure on ${\mathcal S}^{{\mathcal K}}$ is closed in the sense that there is an internal $\Hom$ functor
\[\Hom \colon \left({\mathcal S}^{{\mathcal K}}\right)^{{\mathrm{op}}} \times
{\mathcal S}^{{\mathcal K}} \to {\mathcal S}^{{\mathcal K}}
\]
and a natural isomorphism ${\mathcal S}^{{\mathcal K}}(X
\boxtimes Y, Z) \cong {\mathcal S}^{{\mathcal K}}(X, \Hom(Y,Z))$. This internal $\Hom$ can be
defined via the end construction
\[ \Hom(Y,Z)(\bld{n}) = \int_{\bld{k}\in{\mathcal K}}
\Map(Y(\bld{k}),Z(\bld{n}\concat\bld{k})).\]
\subsection{Free and semi-free
\texorpdfstring{${\mathcal K}$}{K}-spaces}\label{subs:free-semi-free}
Given an object $\bld k$ in ${\mathcal K}$, let us write ${\mathcal K}(\bld k)$ for the monoid of endomorphisms of $\bld k$ and ${\mathcal S}^{{\mathcal K}(\bld k)}$ for the category of spaces with left ${\mathcal K}(\bld k)$-action. The categories ${\mathcal S}$,
${\mathcal S}^{{\mathcal K}(\bld{k})}$, and ${\mathcal S}^{{\mathcal K}}$ are related by various
adjunctions that can be summarized as follows:
\begin{equation}\label{eq:two_adjunctions} \xymatrix{ {\mathcal S}
\ar@<.5ex>@/^/[drr]^(.3){F_{\bld{k}}^{{\mathcal K}}} \ar@<-0.5ex>[d]_{{\mathcal K}(\bld{k})\times -} \\
{\mathcal S}^{{\mathcal K}(\bld{k})} \ar@<0.5ex>[rr]^(.45){G_{\bld{k}}^{{\mathcal K}}} \ar@<-0.5ex>[u] & &
{\mathcal S}^{{\mathcal K}} \ar@<0.5ex>[ll]^(.55){\Ev_{\bld{k}}^{{\mathcal K}}}
\ar@<.5ex>@/_/[ull]^(.7){\Ev_{\bld{k}}^{{\mathcal K}}} \ar@<0.5ex>[r]^{\colim_{{\mathcal K}}} &
\ar@<0.5ex>[l]^{\const_{{\mathcal K}}}{\mathcal S} }
\end{equation}
Here $\const_{{\mathcal K}}$ takes a space to the corresponding constant ${\mathcal K}$-space,
$\colim_{{\mathcal K}}$ is its left adjoint, ${\mathcal K}(\bld{k})\times -$ is
the free ${\mathcal K}(\bld{k})$-space functor, the unlabeled forgetful functor
is its right adjoint, the two instances of $\Ev_{\bld{k}}^{{\mathcal K}}$ are
the evaluations of a ${\mathcal K}$-space at $\bld{k}$ considered as
a space or a ${\mathcal K}(\bld{k})$-space, and $F_{\bld{k}}^{{\mathcal K}}$ and
$G_{\bld{k}}^{{\mathcal K}}$ are the corresponding left adjoints. Explicitly,
for a space $K$ and a ${\mathcal K}(\bld{k})$-space $L$,
\begin{equation}\label{eq:free-functors}
F_{\bld{k}}^{{\mathcal K}}(K) = {\mathcal K}(\bld{k},-) \times K \quad\textrm{ and
}\quad G_{\bld{k}}^{{\mathcal K}}(L) = {\mathcal K}(\bld{k},-)
\times_{{\mathcal K}(\bld{k})} L
\end{equation}
where the expression for $G_{\bld k}^{{\mathcal K}}(L)$ indicates the coequalizer of the evident diagram.
\begin{lemma} \label{lem:product_of_free_Jspaces} There is a natural
isomorphism
\begin{align*}
&F_{\bld{k}}^{{\mathcal K}}(K) \boxtimes
F_{\bld{k'}}^{{\mathcal K}}(K') \cong F_{\bld{k}\concat\bld{k'}}^{{\mathcal K}}
(K\times K')\\
\intertext{for each pair of spaces $K$ and $K'$, and a natural isomorphism}
&G_{\bld{k}}^{{\mathcal K}}(L) \boxtimes G_{\bld{k'}}^{{\mathcal K}}(L')\cong
G_{\bld{k}\concat\bld{k'}}^{{\mathcal K}}
({\mathcal K}(\bld{k}\concat\bld{k'}) \times_{{\mathcal K}(\bld{k})\times{\mathcal K}(\bld{k'})}L\times
L')
\end{align*}
for each ${\mathcal K}(\bld k)$-space $L$ and each ${\mathcal K}(\bld k')$-space $L'$.\qed
\end{lemma}
\section{\texorpdfstring{${\mathcal I}$}{I}-spaces and symmetric spectra}\label{sec:I-space-section}
Let ${\mathcal I}$ be the category whose objects are the finite sets
$\bld{n}=\{1, \dots, n\}$ for $n\geq 0$ ($\bld 0$ is the empty set) and whose morphisms are the injections. The usual ordered concatenation of ordered sets $\concat$ makes this a symmetric monoidal category with unit
$\bld 0$. The symmetry isomorphism
$\chi_{m,n}\colon\bld{m}\concat\bld{n}\to \bld{n}\concat\bld{m}$ is the shuffle moving the first $m$ elements past the last $n$ elements.
\subsection{The category of \texorpdfstring{${\mathcal I}$}{I}-spaces}
Let ${\mathcal S}^{{\mathcal I}}$ be the category of ${\mathcal I}$-spaces, equipped with the symmetric monoidal structure $({\mathcal S}^{{\mathcal I}},\boxtimes, \bld 1_{{\mathcal I}})$ inherited from ${\mathcal I}$. The unit $\bld 1_{{\mathcal I}}={\mathcal I}(\bld 0,-)$ can be identified with the terminal
${\mathcal I}$-space $*$. By definition, an \emph{${\mathcal I}$-space monoid} is a monoid in
${\mathcal S}^{{\mathcal I}}$. We say that a map of ${\mathcal I}$-spaces $X\to Y$ is an
\begin{itemize}
\item
${\mathcal I}$-equivalence if the induced map of homotopy colimits $X_{h{\mathcal I}}\to
Y_{h{\mathcal I}}$ is a weak homotopy equivalence,
\item
${\mathcal I}$-fibration if it is a level fibration and the diagram
\begin{equation}\label{eq:I-fibration}
\xymatrix@-1pc{
X(\bld m) \ar[r] \ar[d]&X(\bld n)\ar[d]\\
Y(\bld m)\ar[r] & Y(\bld n)
}
\end{equation}
is homotopy cartesian for all morphisms $\bld m\to\bld n$ in ${\mathcal I}$,
\item
cofibration if it has the left lifting property with respect to maps of
${\mathcal I}$-spaces that are level acyclic fibrations.
\end{itemize}
These classes specify a model structure on ${\mathcal S}^{{\mathcal I}}$ as we show in Proposition \ref{prop:projective-I-model-structure} below. We shall refer to this as the \emph{projective ${\mathcal I}$-model structure}.
There also is a \emph{positive projective ${\mathcal I}$-model structure} on ${\mathcal S}^{{\mathcal I}}$: Let ${\mathcal I}_+$ be the full subcategory of ${\mathcal I}$ obtained by excluding the initial object $\bld 0$. We say that a map $X\to Y$ of ${\mathcal I}$-spaces is a
\begin{itemize}
\item
positive ${\mathcal I}$-fibration if it is a level fibration for the levels corresponding to objects in ${\mathcal I}_+$ and the diagrams \eqref{eq:I-fibration} are homotopy cartesian for all morphisms in ${\mathcal I}_+$,
\item
positive cofibration if it has the left lifting property with respect to maps of
${\mathcal I}$-spaces that are level acyclic fibrations for the levels corresponding to objects in ${\mathcal I}_+$.
\end{itemize}
A more explicit description of the (positive) cofibrations is given in
Proposition~\ref{prop:latching-characterization}.
\begin{proposition}\label{prop:projective-I-model-structure}
The ${\mathcal I}$-equivalences, the (positive) ${\mathcal I}$-fibrations, and the (positive) cofibrations specify a cofibrantly generated proper simplicial model structure on
${\mathcal S}^{{\mathcal I}}$. These model structures are monoidal and satisfy the monoid axiom.
\end{proposition}
\begin{proof}
The fact that these classes of maps specify a cofibrantly generated model structure is a consequence of Corollary~\ref{cor:I-J-well-structured} together with Proposition~\ref{prop:K-model-str}. These model structures are proper by Corollary~\ref{cor:proper}, simplicial by Proposition~\ref{prop:K-model-is-S-model-str}, monoidal by
Proposition~\ref{prop:K-pushout-product-axiom}, and satisfy the monoid axiom by Proposition~\ref{prop:monoid-axiom}.
\end{proof}
It follows from the definitions that the identity functor on ${\mathcal S}^{{\mathcal I}}$ is the left Quillen functor of a Quillen equivalence from the positive projective to the projective ${\mathcal I}$-model structure, see Proposition \ref{prop:well-structured-comparison}.
\begin{theorem}
The adjunction
$
\colim_{{\mathcal I}} \colon {\mathcal S}^{{\mathcal I}} \rightleftarrows
{\mathcal S} \thinspace \colon \! \! \const_{{\mathcal I}}
$
defines a Quillen equivalence between the (positive) projective ${\mathcal I}$-model structure on ${\mathcal S}^{{\mathcal I}}$ and the usual model structure on ${\mathcal S}$.
\end{theorem}
\begin{proof}
$B{\mathcal I}$ is contractible so this is a special case of Proposition~\ref{prop:colim-const-Q-adjunction}.
\end{proof}
\begin{remark}
A variant of the above Quillen equivalence is considered in \cite{Lind-diagram}
where also part of Proposition~\ref{prop:I-positive-lift} below is verified. One of the main objectives in \cite{Lind-diagram} is a comparison of the ${\mathcal I}$-space units of a symmetric ring spectrum with the corresponding construction in the $S$-module setting from~\cite{EKMM}.
\end{remark}
The next result is the main reason for introducing the positive projective model structure. We write ${\mathcal C}{\mathcal S}^{{\mathcal I}}$ for the category of commutative
${\mathcal I}$-space monoids.
\begin{proposition}\label{prop:I-positive-lift}
The positive projective ${\mathcal I}$-model structure on ${\mathcal S}^{{\mathcal I}}$ lifts to a cofibrantly generated proper simplicial model structure on ${\mathcal C}{\mathcal S}^{{\mathcal I}}$.
\end{proposition}
\begin{proof}
By Corollary \ref{cor:I-J-well-structured} this is a consequence of
Corollaries~\ref{cor:K-positive-projective-commutative} and \ref{cor:proper}.
\end{proof}
More generally, we show in Proposition \ref{prop:structured-lift-proposition} that if ${\mathcal D}$ is any operad in ${\mathcal S}$, then the positive projective model structure lifts to the category of algebras ${\mathcal S}^{{\mathcal I}}[\mathbb D]$ for the associated monad $\mathbb D$ (as usual defined by $\mathbb D(X)=\coprod_{n\geq 0}\mathcal D(n)\times_{\Sigma_n}X^{\boxtimes n}$). If the operad is $\Sigma$-free, then the projective model structure also lifts to ${\mathcal S}^{{\mathcal I}}[\mathbb D]$. Thus, for instance the projective model structure lifts to the category of (not necessarily commutative)
${\mathcal I}$-space monoids.
Recall that an \emph{$E_{\infty}$ operad} ${\mathcal D}$ is an operad which is $\Sigma$-free and whose spaces are contractible, see
Section~\ref{sec:structured-section} for details. As we recall in
Remark \ref{rem:model-str-D-spaces}, the assumption that ${\mathcal D}$ is
$\Sigma$-free ensures that the usual model structure on ${\mathcal S}$ lifts to a model structure on the category of algebras ${\mathcal S}[\mathbb D]$ for the associated monad $\mathbb D$ on ${\mathcal S}$.
\begin{theorem}
Let ${\mathcal D}$ be an $E_{\infty}$ operad and let $\mathbb D$ be the associated monad on ${\mathcal S}$. Then the positive projective ${\mathcal I}$-model structure on
${\mathcal C}{\mathcal S}^{{\mathcal I}}$ is related to the standard model structure on ${\mathcal S}[\mathbb D]$ by a chain of Quillen equivalences.
\end{theorem}
\begin{proof}
This is a special case of Theorem \ref{theorem:K-E-infinity-rectification}.
\end{proof}
Using the explicit Quillen equivalences in the above theorem we can rectify $E_{\infty}$ spaces to strictly commutative ${\mathcal I}$-space monoids in a precise sense.
\begin{corollary}
Let $X$ be an $E_{\infty}$ space for some $E_{\infty}$ operad. Then there exists a commutative ${\mathcal I}$-space monoid $A$ and a chain of
${\mathcal I}$-equivalences of $E_{\infty}$ ${\mathcal I}$-spaces $A\xleftarrow{\sim}
Y\xrightarrow{\sim} X$ relating $A$ to the constant ${\mathcal I}$-space $X$.
\end{corollary}
\begin{proof}
Suppose that $X$ is a $\mathbb D$-algebra in ${\mathcal S}$ for the monad
$\mathbb D$ associated to an $E_{\infty}$ operad ${\mathcal D}$. Then the corresponding constant ${\mathcal I}$-space is a $\mathbb D$-algebra in ${\mathcal S}^{{\mathcal I}}$ and we let $Y\to X$ be a cofibrant replacement in the positive projective model structure on ${\mathcal S}^{\mathcal I}[\mathbb D]$. Let $\pi\colon \mathbb D\to\mathbb C$ be the canonical projection onto the commutativity monad $\mathbb C$. By Proposition~\ref{prop:operad-change} this gives rise to a Quillen equivalence
$\pi_* \colon {\mathcal S}^{{\mathcal I}}[\mathbb D] \rightleftarrows
{\mathcal C}{\mathcal S}^{{\mathcal I}} \thinspace \colon \! \pi^*$ relating the respective positive projective ${\mathcal I}$-model structures. We let $A=\pi_*(Y)$ and observe that the cofibrancy assumption on $Y$ implies that the counit of the adjunction $Y\to\pi^*(A)$ is an ${\mathcal I}$-equivalence.
\end{proof}
In fact, by Lemma~\ref{lem:unit-lemma} we may even choose the chain of
${\mathcal I}$-equivalences in the corollary so that they are level equivalences in positive degrees.
\subsection{The flat model structure on \texorpdfstring{${\mathcal I}$}{I}-spaces}
\label{subsec:flat-I-model}
For the applications of the theory it is important to be able to decide whether a particular ${\mathcal I}$-space is homotopically well-behaved with respect to the
$\boxtimes$-product. We know from Proposition \ref{prop:projective-I-model-structure} that the cofibrant ${\mathcal I}$-spaces in the projective ${\mathcal I}$-model structure have this property, but it is inconvenient to restrict our attention to this class of cofibrant objects. In practice, such cofibrant objects rarely occur naturally and must almost always be manufactured using the small object argument. Also, a commutative ${\mathcal I}$-space monoid that is cofibrant in the lifted positive projective ${\mathcal I}$-model structure on ${\mathcal C}{\mathcal S}^{{\mathcal I}}$ will not in general have an underlying ${\mathcal I}$-space that is cofibrant in the projective ${\mathcal I}$-model structure on ${\mathcal S}^{{\mathcal I}}$. Thus, we need another argument to ensure that the cofibrant objects in ${\mathcal C}{\mathcal S}^{{\mathcal I}}$ are homotopically well-behaved with respect to the $\boxtimes$-product. This motivates introducing the flat
${\mathcal I}$-model structure which is the purpose of this section.
Given an object $\bld n$ in ${\mathcal I}$, we write $({\mathcal I}\downarrow \bld n)$ for the comma category of objects in ${\mathcal I}$ over $\bld n$ and $\partial({\mathcal I}\downarrow\bld n)$ for the full subcategory whose objects $\bld m\to\bld n$ are not isomorphisms. Composing with the forgetful functor $({\mathcal I}\downarrow\bld n)\to {\mathcal I}$, an ${\mathcal I}$-space $X$ gives rise to a diagram indexed by $\partial({\mathcal I}\downarrow\bld n)$. The \emph{$\bld n$th latching space} of $X$ is defined by $L_{\bld n}(X)=\colim_{\partial({\mathcal I}\downarrow\bld n)}X$.
\begin{definition}
A map of ${\mathcal I}$-spaces $X\to Y$ is a
\emph{flat cofibration} if the induced map $X(\bld n)\cup_{L_{\bld n}(X)}
L_{\bld n}(Y)\to Y(\bld n)$ is a cofibration in ${\mathcal S}$ for all $\bld n$. It is a
\emph{positive flat cofibration} if it is a flat cofibration and in addition
$X(\bld 0)\to Y(\bld 0)$ is an isomorphism.
\end{definition}
We say that a map of ${\mathcal I}$-spaces is a (positive) flat ${\mathcal I}$-fibration if it has the right lifting property with respect to the class of (positive) flat cofibrations that are ${\mathcal I}$-equivalences. A more explicit description of the flat $\mathcal I$-fibrations is given in Section~\ref{sec:K-model-structure}.
\begin{proposition}\label{prop:flat-I-model-structure}
The ${\mathcal I}$-equivalences, the (positive) flat ${\mathcal I}$-fibrations, and the (positive) flat cofibrations specify a cofibrantly generated proper simplicial model structure on
${\mathcal S}^{{\mathcal I}}$. These model structures are monoidal and satisfy the monoid axiom.
\end{proposition}
\begin{proof}
By Corollary~\ref{cor:flat-I-J-well-structured} and the remarks preceding it, the fact that these classes of maps specify a cofibrantly generated model structure is a consequence of Proposition~\ref{prop:K-model-str}.
The model structures are proper by Corollary~\ref{cor:proper}, simplicial by
Proposition~\ref{prop:K-model-is-S-model-str}, monoidal by
Proposition~\ref{prop:K-pushout-product-axiom}, and satisfy the monoid axiom by Proposition~\ref{prop:monoid-axiom}.
\end{proof}
We shall refer to this as the \emph{(positive) flat model structure} on
${\mathcal S}^{{\mathcal I}}$ and the cofibrant objects will be called \emph{flat ${\mathcal I}$-spaces}.
It is proved in Proposition~\ref{prop:well-structured-comparison} that the identity functor is the left Quillen functor in a Quillen equivalence from the (positive) projective model structure to the (positive) flat model structure on
${\mathcal S}^{{\mathcal I}}$. In particular, an ${\mathcal I}$-space which is cofibrant in the projective model structure is also flat. One of the convenient properties of a flat ${\mathcal I}$-space $X$ is that the endofunctor $X\boxtimes(-)$ on ${\mathcal S}^{{\mathcal I}}$ preserves ${\mathcal I}$-equivalences; this is proved in
Proposition~\ref{prop:boxtimes-flat-invariance}.
It is useful to reformulate the flat cofibration condition in terms of the well-known Reedy cofibrations of cubical diagrams.
For an object $\bld n$ in ${\mathcal I}$, let ${\mathcal P}(\bld n)$ denote the category with objects the subsets of $\bld n$ and morphisms the inclusions. Functors
$C\colon {\mathcal P}(\bld n)\to{\mathcal S}$ may be viewed as $\bld n$-cubical diagrams. The partial ordering of the objects in ${\mathcal P}(\bld n)$ gives rise to the usual Reedy model structure on the category of $\bld n$-cubical diagrams, see e.g.\ \cite[Ch.\ 15]{Hirschhorn-model}. We shall only make use of the cofibration part of this structure. Given an $\bld n$-cubical diagram $C$ and a subset $V$ of $\bld n$, the \emph{$V$th-latching space} is defined by
$L_V(C)=\colim_{U\varsubsetneq V}C(U)$. A map of $\bld n$-cubical diagrams $C\to D$ is said to be a cofibration if the induced map
$C(V)\cup_{L_V(C)}L_V(D)\to D(V)$ is a cofibration in ${\mathcal S}$ for all subsets
$V$ of $\bld n$. In particular, an $\bld n$-cube $C$ is cofibrant if and only if it is a cofibration cube in the sense of Goodwillie \cite{Goodwillie-calculus},
that is, the map $L_V(C)\to C(V)$ is a cofibration for each subset $V$.
Cubical diagrams arise from ${\mathcal I}$-spaces in the following way: The category
${\mathcal P}(\bld n)$ maps isomorphically onto the skeletal subcategory of
$({\mathcal I}\downarrow \bld n)$ given by the objects $\bld m\to\bld n$ that are order preserving. Composing with the forgetful functor to ${\mathcal I}$, an ${\mathcal I}$-space thus gives rise to an $\bld n$-cubical diagram for all $\bld n$.
It is clear from the definitions that a map of ${\mathcal I}$-spaces $X\to Y$ is a flat cofibration if and only if the induced map of $\bld n$-cubical diagrams is a cofibration for all $\bld n$.
\begin{proposition}\label{prop:simplicial-flatness}
In the simplicial setting an ${\mathcal I}$-space $X$ is flat if and only if each morphism $\bld m\to \bld n$ induces a cofibration $X(\bld m)\to X(\bld n)$ and for each diagram of the following form (with maps induced by the evident order preserving morphisms)
\begin{equation}\label{eq:flat-criterion}
\xymatrix@-1pc{
X(\bld m)\ar[r]\ar[d] & X(\bld m\sqcup\bld n)\ar[d]\\
X(\bld l\sqcup\bld m)\ar[r] & X(\bld l\sqcup\bld m\sqcup\bld n)
}
\end{equation}
the intersection of the images of $X(\bld l\sqcup\bld m)$ and
$X(\bld m\sqcup\bld n)$ in $X(\bld l\sqcup\bld m\sqcup\bld n)$ equals the image of $X(\bld m)$.
\end{proposition}
\begin{proof}
First notice that an $\bld n$-cubical diagram of simplicial sets is a cofibration cube if and only if (i) each inclusion $U\subseteq V$ induces a cofibration $C(U)\to C(V)$, and (ii) for each pair of subsets $U$ and $V$, the intersection of the images of $C(U)$ and $C(V)$ in $C(U\cup V)$ equals the image of $C(U\cap V)$. One can check this inductively using the following principle: Let $V$ be a finite set and $C$ a $V$-cubical diagram. Let $U$ be the subset obtained by removing a point from $V$. Then we may view $C$ as a map of $U$-cubical diagrams $C\colon D\to E$ and there is a pushout diagram
\[
\xymatrix@-1pc{
L_U(D) \ar[r]\ar[d] & D(U)\ar[d]\\
L_U(E) \ar[r] & L_V(C).
}
\]
With this description of a cofibration cube it is clear that the cubical diagrams associated to an ${\mathcal I}$-space are cofibration cubes precisely when the conditions in the lemma are satisfied.
\end{proof}
In the topological setting we cannot state the obvious analogue of the above flatness criterion since we lack a sufficiently general gluing principle for topological cofibrations. Instead we have the following weaker result which is proved by a similar argument.
\begin{proposition}
In the topological setting an ${\mathcal I}$-space $X$ is flat provided that each of the spaces $X(\bld n)$ is a CW-complex, each morphism $\bld m\to\bld n$ induces an isomorphism of $X(\bld m)$ onto a subcomplex of $X(\bld n)$, and for each diagram of the form \eqref{eq:flat-criterion} the intersection of the images of $X(\bld l\sqcup\bld m)$ and $X(\bld m\sqcup\bld n)$ in
$X(\bld l\sqcup\bld m\sqcup\bld n)$ equals the image of $X(\bld m)$.
\qed
\end{proposition}
\begin{example}
The ${\mathcal I}$-spaces $BO$ and $B\mathrm{GL}(R)$ in Examples~\ref{ex:BO-example} and \ref{ex:BGL(R)-example} are flat. The $\mathcal I$-space $X^{\bullet}$ in
Example~\ref{ex:X-bullet-example} is flat in the simplicial setting and is flat in the topological setting if we assume that $X$ is a based CW-complex. None of these ${\mathcal I}$-spaces are cofibrant in the projective model structure.
\end{example}
\begin{remark}
In the topological setting there also is a weaker $h$-cofibration notion of flatness which is characterized by a condition analogous to that in
Proposition~\ref{prop:simplicial-flatness}. This is the flatness criterion used in \cite{Blumberg-C-S_THH-Thom,Schlichtkrull-infinite}, but it is not the right notion in the present setting of cofibrantly generated model categories.
\end{remark}
The next result is one of the main reasons for considering the flat model structure.
\begin{proposition}\label{prop:I-underlying-flat}
\begin{enumerate}[(i)]
\item
The positive flat model structure on ${\mathcal S}^{{\mathcal I}}$ lifts to a cofibrantly generated proper simplicial model structure on ${\mathcal C}{\mathcal S}^{{\mathcal I}}$.
\item
Suppose that $A$ is a commutative ${\mathcal I}$-space monoid which is cofibrant in the lifted model structure in (i). Then the underlying ${\mathcal I}$-space of $A$ is flat.
\end{enumerate}
\end{proposition}
\begin{proof}
The statement in (i) that the positive flat model structure lifts to ${\mathcal C}{\mathcal S}^{{\mathcal I}}$ is a consequence of Corollary \ref{cor:flat-I-J-well-structured} and Proposition~\ref{prop:structured-lift-proposition}. Properness follows from Corollary~\ref{cor:proper} and the claim about the simplicial structure is verified in Proposition~\ref{prop:CSK-simplicial}. The statement in (ii) is a special case of Corollary~\ref{cor:comm-underlying-cofibrant}. \end{proof}
As remarked at the beginning of the section, the analogues result fails for the (positive) projective model structure.
The flat model structure on ${\mathcal I}$-spaces is analogous to the flat model structure on symmetric spectra established in \cite{HSS} and
\cite{Shipley-convenient} (we use the terminology introduced by Schwede \cite{Schwede-SymSp}; the flat model structure on ${\mathrm{\textrm{Sp}^{\Sigma}}}$ is what is called the $S$-model structure in \cite{HSS} and \cite{Shipley-convenient}). In Proposition \ref{prop:cScI-Spsym-adjunction} we make this analogy precise by establishing a Quillen adjunction relating the two model structures.
\subsection{Recollections on symmetric spectra}
When discussing symmetric spectra we shall frequently consider spheres indexed by finite sets and isomorphisms between them induced by bijections.
As explained in Section~\ref{sec:preliminaries} we simultaneously work in a topological and a simplicial setting and we write ${\mathcal S}_*$ for the category of based spaces.
Given a finite set $X$, let $S^X$ be the smash product $\bigwedge_{x\in X}S^1$ defined as the quotient space of the $X$-fold product
$\prod_{x\in X}S^1$ by the subspace where one of the components equals the base point. For a morphism $\alpha\colon\bld m\to\bld n$ in ${\mathcal I}$, we write $\bld n-\alpha$ for the complement of $\alpha(\bld m)$ in $\bld n$. There is a canonical extension of $\alpha$ to a bijection $\bld m\concat(\bld n-\alpha)\to\bld n$ (which is the inclusion of $\bld n-\alpha$) and this gives rise to the isomorphism
\begin{equation}\label{eq:reindexing-basic} S^{\bld{m}} \wedge
S^{\bld{n}-\alpha}\xrightarrow{\cong} S^{\bld{n}}.
\end{equation}
Restricting to morphisms $\alpha \in {\mathcal I}(\bld{n},\bld{n})=\Sigma_n$,
this defines the usual left $\Sigma_n$-action on $S^n$. For a pair of morphisms $\alpha\colon\bld{l}\to\bld{m}$ and $\beta\colon\bld{m}\to\bld{n}$
there is a canonical bijection
$(\bld{m}-\alpha)\concat(\bld{n}-\beta)\to\bld{n}-\beta\alpha$, obtained by applying $\beta$ to the elements in $\bld m-\alpha$,
and an associated isomorphism
\begin{equation}\label{eq:reindexing-composite} S^{\bld{m}-\alpha} \wedge
S^{\bld{n}-\beta} \xrightarrow{\cong} S^{\bld{n}-\beta\alpha}.
\end{equation}
Given morphisms $\alpha\colon\bld{m}\to\bld{n}$ and
$\alpha'\colon\bld{m'}\to\bld{n'}$, there is a canonical identification of
$(\bld{n}\concat\bld{n'})-(\alpha\concat\alpha')$ with
$(\bld{n}-\alpha) \concat (\bld{n'}-\alpha')$ and therefore an isomorphism
\begin{equation}\label{eq:reindexing-product}
S^{(\bld{n}\concat\bld{n'})-(\alpha\concat\alpha')} \cong
S^{\bld{n}-\alpha} \wedge S^{ \bld{n'}-\alpha'}.
\end{equation}
Recall from \cite{HSS} and \cite{MMSS} that a symmetric spectrum $E$ is a spectrum (in ${\mathcal S}_*$) with structure maps $E_m\wedge S^1\to
E_{m+1}$ such that the $m$th space $E_m$ has a left $\Sigma_m$-action and the iterated structure maps $E_m\wedge S^n\to E_{m+n}$ are
$\Sigma_m\times \Sigma_n$-equivariant. Given a morphism
$\alpha\colon\bld m\to\bld n$ in ${\mathcal I}$ there is an induced structure map
$\alpha_*\colon E_m\wedge S^{\bld n-\alpha}\to E_n$ defined as follows: Choose a bijection
$\beta\colon\bld l\to \bld n-\alpha$ for an object $\bld l$ in ${\mathcal I}$ and let
$\{\alpha,\beta\}\colon \bld m\concat\bld l\to\bld n$ be the resulting bijection. Then $\alpha_*$ is defined by
\[
\alpha_*\colon E_m\wedge S^{\bld n-\alpha}\xrightarrow{\mathbf{1}\wedge\beta^{-1}}
E_m\wedge S^{\bld l}\to E_{m+l}\xrightarrow{\{\alpha,\beta\}_*}E_n
\]
which is independent of the choice of $\beta$. With this convention, the subset inclusion $\iota_{\bld{m}}\colon\bld{m}\to\bld{m+1}$ induces the structure map $E_m \wedge S^1 \to E_{m+1}$ and the endomorphisms of
$\bld m$ induce the left $\Sigma_m$-action on $E_m$. See~\cite[\S 3.1]{Schlichtkrull-Thom_symm} and~\cite{Schwede-SymSp} for more details on this perspective on symmetric spectra.
The functor $\Ev_m \colon {\mathrm{\textrm{Sp}^{\Sigma}}} \to {\mathcal S}_*$ sending a symmetric
spectrum $E$ to $E_m$ has a left adjoint $F_m\colon {\mathcal S}_* \to
{\mathrm{\textrm{Sp}^{\Sigma}}}$. It can be defined explicitly as
\begin{equation}\label{eq:free_sym_sp} F_m(K)_n = \bigvee_{\alpha \in {\mathcal I}(\bld{m},\bld{n})} K \wedge
S^{\bld{n}-\alpha}.\end{equation} Here we use the notation $S^{\bld{n}-\alpha}$ to both keep track of the dimension of the sphere and the different copies of it. A morphism $\beta\colon\bld{n}\to\bld{p}$ in ${\mathcal I}$ induces
the structure map $\beta_*\colon F_m(K)_n\wedge S^{\bld{p}-\beta} \to
F_m(K)_p$. This maps the wedge summand indexed by
$\alpha\colon \bld{m}\to\bld{n}$ to the wedge summand indexed by $\beta\alpha$ via the isomorphism
\[ K \wedge S^{\bld{n}-\alpha} \wedge S^{\bld{p}-\beta} \to K \wedge
S^{\bld{p}-\beta\alpha},\] specified by $\beta$ as in
\eqref{eq:reindexing-composite}.
It is proved in \cite{HSS} that the smash product of symmetric spectra makes ${\mathrm{\textrm{Sp}^{\Sigma}}}$ a symmetric monoidal category with unit the sphere spectrum $\mathbb S$. Using the above notation, the smash product $E\wedge E'$ of a pair of symmetric spectra $E$ and $E'$ can be described explicitly in degree $n$ by
\[ (E\wedge E')_{n} =
\colim_{\alpha\colon\bld{k}\concat\bld{k'}\to\bld{n}}E_k\wedge
E'_{k'}\wedge S^{\bld{n}-\alpha}.\] The colimit is taken over the comma category
$(\concat \downarrow \bld{n})$, and a morphism
\[ (\gamma,\gamma')\colon(\bld{k},\bld{k'},
\bld{k}\concat{\bld{k'}}\xrightarrow{\alpha}
\bld{n})\to(\bld{l},\bld{l'},\bld{l}\concat\bld{l'}
\xrightarrow{\beta}\bld{n})\] in this category (where by definition $\alpha=
\beta(\gamma\concat\gamma')$)
induces the map
\[E_{k}\wedge E'_{k'} \wedge S^{\bld{n}-\alpha}
\to E_{k} \wedge S^{\bld{l}-\gamma} \wedge E_{k'} \wedge S^{\bld{l'}-\gamma'}
\wedge S^{\bld{n}-\beta}
\to E_{l}\wedge E_{l'} \wedge S^{\bld{n}-\beta}\]
in the colimit system, again
utilizing the isomorphisms \eqref{eq:reindexing-composite} and
\eqref{eq:reindexing-product}.
Coming back to free symmetric spectra, we can use the above to get an explicit description of the isomorphism
\begin{equation}\label{eq:smash_of_free_iso}
F_m(K) \wedge F_{m'}(K')\xrightarrow{\cong} F_{m+m'}(K \wedge K')
\end{equation}
of~\cite[Proposition 2.2.6(1)]{HSS}. In spectrum degree $n$ this is the map from the colimit over $(\concat \downarrow \bld{n})$ which for each object
$(\bld{k},\bld{k}',\alpha\colon\bld{k}\concat\bld{k}'\to\bld n)$ takes the wedge summand indexed by $\beta\colon\bld m\to\bld k$ and
$\beta'\colon\bld{m}'\to\bld{k}'$ to the wedge summand indexed by
$\gamma(\beta\concat\beta')\colon \bld{m}\concat\bld{m}'\to\bld n$ via the isomorphism
\[
K\wedge S^{\bld k-\beta}\wedge K'\wedge S^{\bld{k}'-\beta'}\wedge
S^{\bld n-\alpha}\to K\wedge K'\wedge S^{\bld n-\alpha(\beta\concat\beta')}
\]
induced by \eqref{eq:reindexing-composite} and \eqref{eq:reindexing-product}.
Under the isomorphism \eqref{eq:smash_of_free_iso}, the symmetry isomorphism of the smash product of $F_m(K)$ and $F_{m'}(K')$ corresponds to the map of free symmetric spectra
$F_{m+m'}(K\wedge K') \to F_{m'+m}(K' \wedge K)$ that maps the wedge summand indexed by $\alpha\colon\bld{m}\concat\bld{m}'\to\bld{n}$
to the wedge summand indexed by
$\alpha\chi_{m,m'}\colon\bld{m}'\concat\bld{m}\to\bld n$ via the isomorphism
\begin{equation}\label{eq:symmetry_iso_spsym}
K\wedge K' \wedge S^{\bld{n}-\alpha} \to K'\wedge K \wedge
S^{\bld{n} - \alpha \chi_{m,m'}}
\end{equation}
that flips the $K$ and $K'$ factors and is the identity on
$S^{\bld n-\alpha}=S^{\bld n-\alpha\chi_{m,m'}}$.
\subsection{\texorpdfstring{${\mathcal I}$}{I}-spaces and symmetric spectra}
An ordinary ring has an underlying monoid which in turn contains the group of units as displayed in the diagram of adjoint functors
\[
\big\{\textrm{(comm.) groups}\big\}
\rightleftarrows
\big\{\textrm{(comm.) monoids}\big\}
\rightleftarrows
\big\{\textrm{(comm.) rings}\big\}.
\]
We wish to model a topological version of these adjunctions using ${\mathcal I}$-space monoids and to use this to define the units of a symmetric ring spectrum.
By adjointness, maps of symmetric spectra $F_n(S^n)\to F_m(S^m)$ are in
one-to-one correspondence with maps $S^n\to F_m(S^m)_n$. For a morphism
$\alpha\colon\bld m\to\bld n$, let $\alpha^*\colon F_n(S^n)\to F_m(S^m)$ be the map which is adjoint to
\[
S^{\bld n}\xleftarrow{\cong}S^{\bld m}\wedge S^{\bld n-\alpha}\to
\bigvee_{\beta\colon\bld m\to\bld n}S^{\bld{m}}\wedge S^{\bld n-\beta}.
\]
The first map is the isomorphism \eqref{eq:reindexing-basic} induced by
$\alpha$ and the second map is the inclusion of $S^{\bld m}\wedge S^{\bld n-\alpha}$ as the wedge summand indexed by $\alpha$.
With the explicit descriptions of free symmetric spectra and smash
products given above, it is easy to verify the following lemma.
(It can also be deduced from Lemma \ref{lem:J-and-free-spsym} below.)
\begin{lemma}\label{lem:I-and-free-spsym}
The free symmetric spectra on spheres assemble to a strong symmetric
monoidal functor
$ F_{-}(S^{-}) \colon {\mathcal I}^{{\mathrm{op}}} \to {\mathrm{\textrm{Sp}^{\Sigma}}}, \ \bld{n} \mapsto
F_n(S^n)$. \qed
\end{lemma}
Let $X$ be an ${\mathcal I}$-space. As in the general situation considered in
Section~\ref{sec:diagram-spaces-symmetric-spectra} we write
$\mathbb S^{{\mathcal I}}[X]$ for the symmetric spectrum defined as the coend of the (${\mathcal I}^{{\mathrm{op}}}\times{\mathcal I}$)-diagram
$F_m(S^m)\wedge X(\bld n)_+$ where $(-)_+$ denotes a disjoint base point. Given a symmetric spectrum $E$ we write $\Omega^{{\mathcal I}}(E)$ for the ${\mathcal I}$-space defined by $\Map_{{\mathrm{\textrm{Sp}^{\Sigma}}}}(F_{-}(S^{-}),E)$.
Let ${\mathcal C}{\mathrm{\textrm{Sp}^{\Sigma}}}$ denote the category of commutative symmetric ring spectra. An application of Proposition \ref{prop:structured-adjunction-SK-Spsym} then
provides the two adjoint pairs of functors
\begin{equation}\label{eq:cScI-Spsym-adjunction-copied}
{\mathbb S}^{{\mathcal I}}[-] \colon
{\mathcal S}^{{\mathcal I}} \rightleftarrows {\mathrm{\textrm{Sp}^{\Sigma}}} \colon
\Omega^{{\mathcal I}}
\qquad \text{and} \qquad
{\mathbb S}^{{\mathcal I}}[-] \colon {\mathcal C}{\mathcal S}^{{\mathcal I}} \rightleftarrows {\mathcal C}{\mathrm{\textrm{Sp}^{\Sigma}}} \colon
\Omega^{{\mathcal I}},
\end{equation}
and more generally an adjunction relating the categories of
$\mathbb D$-algebras ${\mathcal S}^{{\mathcal I}}[\mathbb D]$ and ${\mathrm{\textrm{Sp}^{\Sigma}}}[\mathbb D]$ for any operad ${\mathcal D}$ with associated monad $\mathbb D$. Checking from the definitions, we find that
\[
{\mathbb S}^{{\mathcal I}}[X]_n=S^n\wedge X(\bld n)_+
\quad\textrm{ and } \quad \Omega^{{\mathcal I}}(E)(\bld{n})=
\Omega^{n}(E_n).
\]
Recall that we use the term \emph{flat model structure} for the model structure on ${\mathrm{\textrm{Sp}^{\Sigma}}}$ which is called the $S$-model structure in \cite{HSS} and \cite{Shipley-convenient}.
\begin{proposition}\label{prop:cScI-Spsym-adjunction}
\begin{enumerate}[(i)]
\item
The first adjunction in \eqref{eq:cScI-Spsym-adjunction-copied} is a Quillen adjunction with respect to the (positive) projective and (positive) flat
${\mathcal I}$-model structures on ${\mathcal S}^{{\mathcal I}}$ and the corresponding (positive) projective and (positive) flat stable model structures on ${\mathrm{\textrm{Sp}^{\Sigma}}}$.
\item
The second adjunction in \eqref{eq:cScI-Spsym-adjunction-copied} is a Quillen adjunction with respect to the positive projective and positive flat
${\mathcal I}$-model structures on ${\mathcal C}{\mathcal S}^{{\mathcal I}}$ and the corresponding positive projective and positive flat stable model structures on ${\mathcal C}{\mathrm{\textrm{Sp}^{\Sigma}}}$.
\end{enumerate}
\end{proposition}
\begin{proof}
This is a consequence of Proposition \ref{prop:K-spaces-Spsym-adjunction} and the descriptions of the respective model structures on ${\mathrm{\textrm{Sp}^{\Sigma}}}$ given in
\cite{HSS,MMSS,Shipley-convenient}; see also
\cite{Schwede-SymSp}. In the flat case the main point is that the diagonal
$\Sigma_n$-action on $F_n(S^n)$, acting both on ${\mathcal I}(\bld n,-)$ and $S^n$, is free away from the base point.
\end{proof}
The ${\mathcal I}$-space monoid of units $A^{\times}$ associated to an ${\mathcal I}$-space monoid $A$ is defined by letting $A^{\times}(\bld n)$ be the union of the components in $A(\bld n)$ that represent units in the monoid $\pi_0(A_{h\mathcal I})$.
It follows immediately from the definitions that if $A$ is (positive) fibrant, then $A^{\times}$ is also (positive) fibrant, and that if $A$ is commutative, then $A^{\times}$ is also commutative.
We say that an ${\mathcal I}$-space monoid $A$ is \emph{grouplike} if $A_{h{\mathcal I}}$ is a grouplike monoid in ${\mathcal S}$. Clearly $A$ is grouplike if and only if $A^{\times}=A$, which implies that the functor $A\mapsto A^{\times}$ from (commutative) ${\mathcal I}$-space monoids to (commutative) grouplike $\mathcal I$-space monoids is a right adjoint of the inclusion functor.
\begin{definition}
Let $R$ be a symmetric ring spectrum. The ${\mathcal I}$-space units of $R$ is the grouplike ${\mathcal I}$-space monoid $\GL^{\cI}_1(R)=\Omega^{{\mathcal I}}(R)^{\times}$.
\end{definition}
The functor $R\mapsto \GL^{\cI}_1(R)$ is defined for all symmetric ring spectra $R$ and provides a right adjoint of the functor that to a (commutative) grouplike ${\mathcal I}$-space monoid $A$ associates the (commutative) symmetric ring spectrum $\mathbb S^{{\mathcal I}}[A]$. However, one should keep in mind that
$\GL^{\cI}_1(R)$ only represents the ``correct'' homotopy type of the units when $R$ is positive fibrant (or at least semistable in the sense of \cite{Shipley-THH}).
\begin{proposition}
If $R$ is a (positive) fibrant symmetric ring spectrum, then the monoid homomorphism $\pi_0(\GL^{\cI}_1(R)_{h{\mathcal I}})\to \pi_0(\Omega^{{\mathcal I}}(R)_{h{\mathcal I}})$ realizes the inclusion of units $\pi_0(R)^{\times}\to\pi_0(R)$.\qed
\end{proposition}
\section{\texorpdfstring{${\mathcal J}$}{J}-spaces and symmetric spectra}\label{sec:J-space-section}
Here we introduce the category ${\mathcal J}$ and discuss the homotopy theory of ${\mathcal J}$-spaces and the relation to symmetric spectra. We end the section with a sample application to topological logarithmic structures.
\subsection{The category \texorpdfstring{${\mathcal J}$}{J}}\label{subs:cat-J}
First we give an explicit description of the category ${\mathcal J}$. After this, in Proposition \ref{prop:Grayson-Quillen-J} below, we exhibit ${\mathcal J}$ as Quillen's localization construction on the category of finite sets and bijections.
\begin{definition} The objects of the category ${\mathcal J}$ are pairs
$(\bld{n_1},\bld{n_2})$ of objects in ${\mathcal I}$ and a morphism
$(\bld{m_1},\bld{m_2}) \to (\bld{n_1},\bld{n_2})$ is a triple
$(\beta_1,\beta_2, \sigma)$ with $\beta_1\colon \bld {m_1}\to \bld{n_1}$ and
$\beta_2\colon \bld{m_2}\to\bld{n_2}$ morphisms in ${\mathcal I}$, and
$\sigma\colon\bld{n_1}-\beta_1 \to \bld{n_2} - \beta_2$ a bijection
identifying the complement of $\beta_1(\bld{m_1})$ in $\bld{n_1}$ with
the complement of $\beta_2(\bld{m_2})$ in $\bld{n_2}$.
Given composable morphisms
\[ (\bld{l_1},\bld{l_2}) \xrightarrow{(\alpha_1, \alpha_2,\rho)}(\bld{m_1},\bld{m_2})\xrightarrow{(\beta_1, \beta_2,\sigma)}(\bld{n_1},\bld{n_2}),\]
the first two entries of their composite $(\gamma_1,\gamma_2,\tau)$ are
$\gamma_1=\beta_1\alpha_1$ and $\gamma_2=\beta_2 \alpha_2$. It
remains to specify a bijection $\tau \colon \bld{n_1} - \beta_1\alpha_1
\to \bld{n_2}-\beta_2\alpha_2$. The set $\bld{n_i} -\beta_i\alpha_i$ is the
disjoint union of $\bld{n_i}-\beta_i$ and
$\beta_i(\bld{m_i}-\alpha_i)$ for $i=1,2$, and we define
\[\tau(s)=\begin{cases} \sigma(s) &\textrm{ if } s \in
\bld{n_1} - \beta_1 \quad\textrm{ and } \\ \beta_2(\rho(t))
&\textrm{ if } s= \beta_1(t) \in \beta_1(\bld{m_1} -
\alpha_1). \end{cases}\]
Slightly imprecisely, we refer to
$\tau$ as $\sigma \cup \beta_2 \rho \beta_1^{-1}$. To see that
${\mathcal J}$ is indeed a category we have to verify that composition is
associative. Consider the composable morphisms
\[
(\bld{k_1},\bld{k_2})\xrightarrow{(\alpha_1,\alpha_2,\rho)}(\bld{l_1},\bld{l_2})
\xrightarrow{(\beta_1,\beta_2,\sigma)}(\bld{m_1},\bld{m_2})
\xrightarrow{(\gamma_1,\gamma_2,\tau)}(\bld{n_1},\bld{n_2}).
\]
Associativity is clear for the injections in the first two entries. For the bijections one checks that
\[
\tau\cup\gamma_2\sigma\gamma_1^{-1}\cup \gamma_2\beta_2\rho(\gamma_1\beta_1)^{-1}=\tau\cup\gamma_2(\sigma\cup\beta_2\rho\beta_1^{-1})\gamma_1^{-1}.
\]
\end{definition}
Let $\concat \colon
{\mathcal J} \times {\mathcal J} \to {\mathcal J}$ be the functor defined on objects by
\[(\bld{m_1},\bld{m_2}) \concat (\bld{n_1},\bld{n_2}) =
(\bld{m_1}\concat\bld{n_1},\bld{m_2}\concat\bld{n_2}),\]
and on morphisms by
$(\alpha_1,\alpha_1,\rho)\concat(\beta_1,\beta_2,\sigma) =
(\alpha_1\concat\beta_1,\alpha_2\concat\beta_2,\rho\concat\sigma)$, where
$\rho\concat \sigma$ is the bijection induced by $\rho$ and $\sigma$.
Recall that a permutative category is a symmetric monoidal category with strict unit and strict associativity, see for example
\cite[Definition 3.1]{Elmendorf-M_infinite-loop}. The fact that $\mathcal I$ is permutative easily implies that the same holds for ${\mathcal J}$.
\begin{proposition} The data $({\mathcal J}, \concat, (\bld{0},\bld{0}))$ defines a permutative category with symmetry isomorphism
\[(\chi_{m_1,n_1},\chi_{m_2,n_2},\mathbf{1}_{\emptyset}) \colon
(\bld{m_1},\bld{m_2})\concat(\bld{n_1},\bld{n_2}) \to
(\bld{n_1},\bld{n_2})\concat(\bld{m_1},\bld{m_2}).
\eqno\qed
\]
\end{proposition}
It is easy to see that there is a strong symmetric monoidal diagonal
functor $\Delta\colon {\mathcal I} \to {\mathcal J}$ with
$\Delta(\bld{n})=(\bld{n},\bld{n})$ and $\Delta(\alpha\colon
\bld{m}\to\bld{n})=(\alpha,\alpha,\mathbf{1}_{\bld{n}-\alpha})$. Many
constructions in connection with ${\mathcal I}$ and ${\mathcal J}$ will be related
through $\Delta$.
\begin{proposition}\label{prop:Grayson-Quillen-J}
The category ${\mathcal J}$ is isomorphic to Quillen's localization construction $\Sigma^{-1}\Sigma$ on the category $\Sigma$ of finite sets and bijections.
\end{proposition}
\begin{proof}
Let $\Sigma \subset {\mathcal I}$ be the subcategory of finite sets and
bijections with the symmetric monoidal structure inherited from ${\mathcal I}$.
Quillen's localization construction~\cite[p.\ 219]{Grayson-higher} on $\Sigma$
is the category $\Sigma^{-1}\Sigma$ whose objects are pairs $(\bld{n_1},\bld{n_2})$ of objects in ${\mathcal I}$, and whose
morphisms from $(\bld{m_1},\bld{m_2})$ to $(\bld{n_1},\bld{n_2})$ are
isomorphism classes of tuples
\[ \big((\bld{m_1},\bld{m_2}), (\bld{n_1},\bld{n_2}), \bld{l},
(\bld{m_1}\concat\bld{l},\bld{m_2}\concat\bld{l})
\xrightarrow{(\alpha_1,\alpha_2)}(\bld{n_1},\bld{n_2})\big).\]
Here $\bld l$ is an object in $\Sigma$ and $(\alpha_1,\alpha_2)$ is a
morphism in $\Sigma\times\Sigma$.
An isomorphism of tuples is given by a morphism
$\sigma\colon\bld{l}\to\bld{l}$ in $\Sigma$ such that
\[\xymatrix@-1pc{
(\bld{m_1}\concat\bld{l},\bld{m_2}\concat\bld{l})
\ar[rr]^{(\mathbf{1}_{\bld{m_1}}\concat\sigma,
\mathbf{1}_{\bld{m_2}}\concat\sigma)}
\ar[dr]_{(\alpha_1,\alpha_2)} && (\bld{m_1}\concat\bld{l},
\bld{m_2}\concat\bld{l}) \ar[dl]^{(\alpha_1',\alpha_2')} \\
& (\bld{n_1},\bld{n_2}) }\] commutes. Notice, that whereas in
\cite{Grayson-higher} $\Sigma^{-1}\Sigma$ is defined using the monoidal
left action of $\Sigma$ on itself, we here use the right action instead. The
resulting categories are canonically isomorphic, but our conventions are
more convenient when defining the components of a ${\mathcal J}$-space
monoid (see Section \ref{subsec:J-components} below).
The desired isomorphism
$\Sigma^{-1}\Sigma \to {\mathcal J}$ is defined by sending a morphism
represented by $(\alpha_1,\alpha_2)$ as above to the morphism
\[ \big(\alpha_1|_{\bld{m_1}}, \alpha_2|_{\bld{m_2}},
(\bld{n_1}-\alpha_1({\bld{m_1}}))\xrightarrow{({\alpha_1}|_{\bld{l}})^{-1}}
\bld{l} \xrightarrow{\alpha_2|_{\bld{l}}}
(\bld{n_2}-\alpha_2(\bld{m_2}))\big).\]
This does not depend on the choice of representative
$(\alpha_1, \alpha_2)$.
\end{proof}
The arguments of~\cite[p.\ 224]{Grayson-higher} and the
Barratt-Priddy-Quillen Theorem therefore determine the homotopy type of the classifying space of ${\mathcal J}$.
\begin{corollary}\label{cor:BPQ-corollary}
The classifying space $B{\mathcal J}$ is homotopy equivalent to $Q(S^0)$.\qed
\end{corollary}
\begin{remark}
As pointed out to the authors, Kro~\cite{Kro-orthogonal}
considered the analogue of the category ${\mathcal J}$ for orthogonal spectra
in order to define a symmetric monoidal fibrant replacement functor
for orthogonal spectra. This application does not carry over to
symmetric spectra, cf.\ \cite[Remark 3.4]{Kro-orthogonal}.
The analogues to our applications of ${\mathcal J}$ in the
orthogonal context are not addressed in~\cite{Kro-orthogonal},
although it is potentially interesting to consider diagram spaces
indexed by the category Kro describes.
\end{remark}
\subsection{The category of ${\mathcal J}$-spaces}
Let ${\mathcal S}^{{\mathcal J}}$ be the category of ${\mathcal J}$-spaces, equipped with the symmetric monoidal structure $({\mathcal S}^{{\mathcal J}},\boxtimes, \bld 1_{{\mathcal J}})$ inherited from ${\mathcal J}$. Notice that, contrary to the situation for ${\mathcal I}$-spaces, the unit $\bld 1_{{\mathcal J}}={\mathcal J}((\bld{0},\bld{0}), -)$ is not isomorphic to the terminal ${\mathcal J}$-space $*$. By definition, a \emph{${\mathcal J}$-space monoid} is a monoid in ${\mathcal S}^{{\mathcal J}}$. We say that a map of
${\mathcal J}$-spaces $X\to Y$ is a
\begin{itemize}
\item
${\mathcal J}$-equivalence if the induced map of homotopy colimits $X_{h{\mathcal J}}\to
Y_{h{\mathcal J}}$ is a weak homotopy equivalence,
\item
${\mathcal J}$-fibration if it is a level fibration and the diagram
\begin{equation}\label{eq:J-fibration}
\xymatrix@-1pc{
X(\bld m_1,\bld m_2) \ar[r] \ar[d]&X(\bld n_1,\bld n_2)\ar[d]\\
Y(\bld m_1,\bld m_2)\ar[r] & Y(\bld n_1,\bld n_2)
}
\end{equation}
is homotopy cartesian for all morphisms $(\bld m_1,\bld m_2)\to
(\bld n_1,\bld n_2)$ in ${\mathcal J}$,
\item
cofibration if it has the left lifting property with respect to maps of
${\mathcal J}$-spaces that are level acyclic fibrations.
\end{itemize}
These classes specify a model structure on ${\mathcal S}^{{\mathcal J}}$ as we show in Proposition \ref{prop:projective-J-model-structure} below. We shall refer to this as the \emph{projective ${\mathcal J}$-model structure}. There also is a \emph{positive projective ${\mathcal J}$-model structure} on ${\mathcal S}^{{\mathcal J}}$ as we discuss next. Let ${\mathcal J}_+$ be the full subcategory of ${\mathcal J}$ with objects $(\bld n_1,\bld n_2)$ such that
$|\bld n_1|\geq 1$. We say that a map $X\to Y$ of ${\mathcal J}$-spaces is a
\begin{itemize}
\item
positive ${\mathcal J}$-fibration if it is a level fibration for the levels corresponding to objects in ${\mathcal J}_+$ and the diagrams \eqref{eq:J-fibration} are homotopy cartesian for all morphisms in ${\mathcal J}_+$,
\item
positive cofibration if it has the left lifting property with respect to maps of
${\mathcal J}$-spaces that are level acyclic fibrations for the levels corresponding to objects in ${\mathcal J}_+$.
\end{itemize}
A more explicit description of the (positive) cofibrations is given in
Proposition~\ref{prop:latching-characterization}.
\begin{proposition}\label{prop:projective-J-model-structure}
The ${\mathcal J}$-equivalences, the (positive) ${\mathcal J}$-fibrations, and the (positive) cofibrations specify a cofibrantly generated proper simplicial model structure on
${\mathcal S}^{{\mathcal J}}$. These model structures are monoidal and satisfy the monoid axiom.
\end{proposition}
\begin{proof}
The fact that these classes of maps specify a cofibrantly generated model structure is a consequence of Corollary~\ref{cor:I-J-well-structured} together with Proposition~\ref{prop:K-model-str}. These model structures are proper by Corollary~\ref{cor:proper}, simplicial by Proposition~\ref{prop:K-model-is-S-model-str}, monoidal by
Proposition~\ref{prop:K-pushout-product-axiom}, and satisfy the monoid axiom by Proposition~\ref{prop:monoid-axiom}.
\end{proof}
It is clear from the definitions that the identity functor on ${\mathcal S}^{{\mathcal J}}$ is the left Quillen functor of a Quillen equivalence from the positive projective to the projective ${\mathcal J}$-model structure, see Proposition \ref{prop:well-structured-comparison}.
For the next result we equip the category ${\mathcal S}/B{\mathcal J}$ of spaces over $B{\mathcal J}$ with the standard model structure inherited from the usual model structure on
${\mathcal S}$.
\begin{theorem}
There is a chain of Quillen equivalences relating ${\mathcal S}^{{\mathcal J}}$ with the projective ${\mathcal J}$-model structure to ${\mathcal S}/B{\mathcal J}$ with the standard model structure.
\end{theorem}
\begin{proof}
This is a special case of Theorem~\ref{thm:K-spaces-over-BK}.
\end{proof}
Any ${\mathcal J}$-space is naturally augmented over the terminal ${\mathcal J}$-space
$*$. On the level of homotopy categories the above adjunction takes a
${\mathcal J}$-space $X$ to the induced map $X_{h{\mathcal J}}\to *_{h{\mathcal J}}=B{\mathcal J}$. As
discussed in the introduction, this justifies interpreting
${\mathcal J}$-spaces as graded objects.
We write ${\mathcal C}{\mathcal S}^{{\mathcal J}}$ for the category of commutative ${\mathcal J}$-space monoids.
\begin{proposition}
The positive projective ${\mathcal J}$-model structure on ${\mathcal S}^{{\mathcal J}}$ lifts to a cofibrantly generated proper simplicial model structure on ${\mathcal C}{\mathcal S}^{{\mathcal J}}$.
\end{proposition}
\begin{proof}
By Corollary \ref{cor:I-J-well-structured} this is a special case of
Corollary~\ref{cor:K-positive-projective-commutative}.
\end{proof}
As for ${\mathcal I}$-spaces both the projective and the positive projective ${\mathcal J}$-model structures lift to the category of (not necessarily commutative) ${\mathcal J}$-space monoids; this is a consequence of
Proposition~\ref{prop:structured-lift-proposition}.
Let again ${\mathcal E}$ denote the Barratt-Eccles operad and $\mathbb E$ the associated monad on ${\mathcal S}^{{\mathcal J}}$. As we recall in
Lemma~\ref{lem:Barratt-Eccles-action}, the fact that
${\mathcal J}$ is permutative implies that $B{\mathcal J}$ is an $\mathbb E$-algebra, so
the standard model structure on the category ${\mathcal S}[\mathbb E]$ of
$\mathbb E$-algebras in ${\mathcal S}$ lifts to a model structure on the category ${\mathcal S}[\mathbb E]/B{\mathcal J}$ of $\mathbb E$-algebras over $B{\mathcal J}$. We shall refer to this as the standard model structure on ${\mathcal S}[\mathbb E]/B{\mathcal J}$.
\begin{theorem}
There is a chain of Quillen equivalences relating the positive projective
${\mathcal J}$-model structure on ${\mathcal C}{\mathcal S}^{{\mathcal J}}$ to the standard model structure on
${\mathcal S}[\mathbb E]/B{\mathcal J}$.
\end{theorem}
\begin{proof}
This is a special case of Theorem~\ref{thm:comm-graded-theorem}.
\end{proof}
For a commutative ${\mathcal J}$-space monoid $A$, the homotopy colimit $A_{h{\mathcal J}}$ is canonically an $\mathbb E$-algebra. On the level of homotopy categories the above adjunction takes $A$ to the induced map of $\mathbb E$-algebras
$A_{h{\mathcal J}}\to B{\mathcal J}$.
\begin{lemma}\label{lem:BJ-homotopy-cartesian}
If $X$ is a fibrant ${\mathcal J}$-space, then the commutative square
\[
\xymatrix@-1pc{
X(\bld n_1,\bld n_2) \ar[r]\ar[d] & X_{h{\mathcal J}}\ar[d] \\
\{(\bld n_1,\bld n_2)\} \ar[r] &B{\mathcal J}
}
\]
is homotopy cartesian for every object $(\bld n_1,\bld n_2)$ in ${\mathcal J}$.
If $X$ is positive fibrant, then the square is homotopy cartesian when
$|\bld n_1|\geq 1$.
\end{lemma}
\begin{proof}
We first prove the result in the simplicial setting. Since by definition a fibrant
${\mathcal J}$-space is homotopy constant, the first statement is an immediate consequence of \cite[IV Lemma 5.7]{Goerss_J-simplicial}. Using that ${\mathcal J}_+$ is homotopy cofinal in ${\mathcal J}$, a similar argument gives the second statement in the lemma. The topological versions of these statements can be reduced to the simplicial versions by applying the singular complex functor; see Remark \ref{rem:puppe-remark} for details.
\end{proof}
Let us write $\{\pm1\}$ for the group with two elements.
We know from Corollary~\ref{cor:BPQ-corollary} that $\pi_1(B{\mathcal J},*)$ is isomorphic to $\{\pm1\}$ for every choice of base point $*$. It will be convenient to have an explicit description of this isomorphism. If we view
$\{\pm1\}$ as a symmetric monoidal category with a single object, then the sign function defines a symmetric monoidal functor $\sgn\colon\Sigma\to \{\pm1\}$ and therefore a functor
\begin{equation}\label{eq:sgn}
\sgn\colon{\mathcal J}\cong\Sigma^{-1}\Sigma\to\{\pm1\}^{-1}\{\pm1\}\cong \{\pm1\}.
\end{equation}
In particular, an endomorphism $(\alpha_1,\alpha_2)$ of an object
$(\bld n_1,\bld n_2)$ in ${\mathcal J}$ is mapped to $\sgn(\alpha_2)\sgn(\alpha_1^{-1})$. Thus, if $|\bld n_1|\geq 2$ or $|\bld n_2|\geq 2$, then we can represent the non-trivial element of $\pi_1(B{\mathcal J},(\bld n_1,\bld n_2))$ by any endomorphism $(\alpha_1,\alpha_2)$ such that
$\sgn(\alpha_1)=-\sgn(\alpha_2)$.
\begin{corollary}\label{cor:sign-action}
If $X$ is a fibrant (respectively a positive fibrant) ${\mathcal J}$-space, then
$\pi_0(X(\bld n_1,\bld n_2))$ has a canonical action of $\pi_1(B{\mathcal J},(\bld n_1,\bld n_2))$ for all objects $(\bld n_1,\bld n_2)$ (respectively for all objects such that $|\bld n_1|\geq 1$). An element of $\pi_1(B{\mathcal J},(\bld n_1,\bld n_2))$ represented by an endomorphism $(\alpha_1,\alpha_2)$ acts as
$\pi_0(X(\alpha_1,\alpha_2))$.\qed
\end{corollary}
\subsection{Components and units of ${\mathcal J}$-space monoids}
\label{subsec:J-components}
Our first task is to decide what kind of object the components of a ${\mathcal J}$-space should be.
\begin{definition}\label{def:graded-signed-monoid}
A graded signed monoid $M$ is a collection of $\{\pm 1\}$-sets $M_t$ for
$t \in \mathbb Z$, together with a unit $e\in M_0$ and maps $\mu_{s,t}\colon M_s\times M_t\to M_{s+t}$ for all $s,t\in \mathbb Z$. The multiplication maps $\mu_{s,t}$ are assumed to be associative, unital, and
$(\{\pm1\}\!\times\!\{\pm1\})$-equivariant, where $\{\pm1\}\!\times\!\{\pm1\}$ acts on $M_{s+t}$ through the product. We say that $M$ is graded commutative if $\mu_{s,t}(a,b)=(-1)^{st}\mu_{t,s}(b,a)$ for all $a\in M_s$ and $b\in M_t$.
\end{definition}
This notion is placed in the general context of diagram categories as
follows. Let $\tilde {\mathcal J}$ be the product category $\{\pm1\}\times
\mathbb Z$ where we view $\mathbb Z$ as a discrete category with only
identity morphisms and $\{\pm1\}$ as a category with a single object.
We view $\{\pm1\}\times \mathbb Z$ as a product of monoidal categories
and define a symmetric monoidal structure on $\tilde {\mathcal J}$ by
specifying the isomorphisms $(-1)^{mn}\colon m+n\to n+m$. The
symmetric monoidal structure of $\tilde{\mathcal J}$ induces a symmetric
monoidal structure on the category of set valued $\tilde{\mathcal J}$-diagrams
and a monoid $M$ in this diagram category is the same thing as a
graded signed monoid as defined above. Moreover, $M$ is commutative as
a $\tilde{\mathcal J}$-diagram monoid if and only if it is graded commutative
as a graded signed monoid.
Next observe that there is a functor $\Sgn\colon {\mathcal J}\to\tilde{\mathcal J}$
which on objects takes $(\bld n_1,\bld n_2)$ to the integer $n_2-n_1$
and on morphisms is given by the functor $\sgn$ in \eqref{eq:sgn}.
This becomes a strong symmetric monoidal functor when we specify the
isomorphisms
\[
(-1)^{m_1(n_2-n_1)}\colon \Sgn(\bld m_1,\bld m_2)+\Sgn(\bld n_1,\bld n_2)\to
\Sgn(\bld m_1\sqcup\bld n_1,\bld m_2\sqcup\bld n_2).
\]
For this to work it is important that we have defined $\Sigma^{-1}\Sigma$ (and hence $\sgn$ in~\eqref{eq:sgn}) using the monoidal right action of $\Sigma$ on itself, cf.\ the proof of Proposition \ref{prop:Grayson-Quillen-J}. Our sign conventions are motivated by the comparison to homotopy groups of symmetric ring spectra in Proposition~\ref{prop:realizing-homotopy-groups} below.
The connected components of ${\mathcal J}$ are the full subcategories ${\mathcal J}_t$,
for $t\in \mathbb Z$, with objects $(\bld n_1,\bld n_2)$ such that
$n_2-n_1=t$. Let ${\mathcal N}_t$ be the subcategory of ${\mathcal J}_t$ with the same
objects and morphisms $(\iota_1,\iota_2,\chi)$, where $\iota_i$ is a subset inclusion of the form $\iota_i\colon \bld n_i\to\bld n_i \sqcup \bld1$ and $\chi$ is the unique bijection identifying the
complements (thus, $\mathcal N_t$ is isomorphic to the ordered set of
natural numbers). Given a ${\mathcal J}$-space $X$ we define $\pi_{0,t}(X)$ to
be the set
\[
\pi_{0,t}(X)=\colim_{{\mathcal N}_t}\{ \dots\to \pi_0(X(\bld n_1,\bld n_2))\to
\pi_0(X(\bld n_1\sqcup \bld 1,\bld n_2 \sqcup \bld 1))\to\dots \}
\]
and write $\pi_0(X)$ for the $\mathbb Z$-graded set
$\{\pi_{0,t}(X)\colon t\in \mathbb Z\}$. Recall from Corollary
\ref{cor:sign-action} that if $X$ is positive fibrant, then $\pi_0(X(\bld
n_1,\bld n_2))$ has a canonical $\{\pm 1\}$-action for $|\bld n_1|\geq 1$. This gives rise to
a $\{\pm 1\}$-action on $\pi_{0,t}(X)$ for each $t$ such that
$\pi_0(X)$ defines a $\tilde {\mathcal J}$-diagram.
Now suppose that $A$ is a (positive) fibrant ${\mathcal J}$-space monoid. Then
we claim that $\pi_0(A)$ has a uniquely determined structure as a
graded signed monoid such that the canonical map $A\to \pi_0(A)\circ
\Sgn$ is a map of ${\mathcal J}$-diagram monoids. Indeed, it easily follows
from the definitions that the maps
\[
\pi_0(A(\bld m_1,\bld m_2))\times\pi_0(A(\bld n_1,\bld n_2))
\xrightarrow{(-1)^{m_1(n_2-n_1)}\mu}
\pi_0(A(\bld m_1\sqcup \bld n_1,\bld m_2\sqcup \bld n_2))
\]
(where $\mu$ denotes the multiplication in $A$) give rise to the
required multiplication maps
$\pi_{0,s}(A)\times\pi_{0,t}(A)\to\pi_{0,s+t}(A)$. The unit is
represented by the image in $\pi_0(A(\bld 0,\bld 0))$ of the monoidal
unit $*\to A(\bld 0,\bld 0)$.
We summarize the properties of $\pi_0(A)$ in the next proposition.
\begin{proposition}
If $A$ is a (positive) fibrant ${\mathcal J}$-space monoid, then $\pi_0(A)$ inherits the structure of a graded signed monoid. If $A$ is commutative, then $\pi_0(A)$ is graded commutative. \qed
\end{proposition}
Since for a commutative ${\mathcal J}$-space monoid $A$ the homotopy colimit
$A_{h{\mathcal J}}$ is an $E_{\infty}$ space, it is clear that the monoid of components
$\pi_0(A_{h{\mathcal J}})$ is commutative. In general, this monoid has to be different from the underlying ungraded monoid of $\pi_0(A)$ because graded commutativity does not become commutativity when forgetting the grading.
By Lemma \ref{lem:BJ-homotopy-cartesian} and Corollary \ref{cor:sign-action}
we have the following description.
\begin{corollary}\label{cor:J-hocolim-components}
Let $A$ be a ${\mathcal J}$-space monoid and $A\to \bar A$ a (positive) fibrant replacement. Then the monoid $\pi_0(A_{h{\mathcal J}})$ is isomorphic to the quotient of the underlying ungraded monoid of $\pi_0(\bar A)$ by the action of $\{\pm1\}$. \qed
\end{corollary}
The ${\mathcal J}$-space monoid of units $A^{\times}$ associated to a ${\mathcal J}$-space monoid $A$ is defined by choosing a fibrant replacement $A\to \bar A$ and letting $A^{\times}(\bld n_1,\bld n_2)$ be the union of the components in $A(\bld n_1,\bld n_2)$ that represent units in the graded signed monoid
$\pi_0(\bar A)$. It is easy to see that $A^{\times}(\mathbf n_1, \mathbf n_2)$ is independent of the choice of fibrant replacement. In order to see that this definition of $A^{\times}$ actually produces a ${\mathcal J}$-space monoid it is convenient to give an equivalent description of
$A^{\times}(\mathbf n_1, \mathbf n_2)$. Notice first that
$\pi_0(A_{h{\mathcal J}})$ is naturally isomorphic to
$\colim_{(\mathbf n_1,\mathbf n_2)\in{\mathcal J}}\pi_0(A(\mathbf n_1,\mathbf n_2))$ (the analogous statement holds for any diagram space indexed by a small category). By Corollary~\ref{cor:J-hocolim-components} we can therefore characterize $A^{\times}(\mathbf n_1, \mathbf n_2)$ as the union of the components in $A(\mathbf n_1, \mathbf n_2)$ that represent units in
$\pi_0(A_{h{\mathcal J}})$. It is now clear that $A^{\times}$ has a unique ${\mathcal J}$-space monoid structure such that $A^{\times}\to A$ is a fibration of ${\mathcal J}$-space monoids. If $A$ is commutative, then so is $A^{\times}$.
\begin{lemma}\label{lem:unit-inclusion}
If $A$ is a (positive) fibrant ${\mathcal J}$-space monoid, then $A^{\times}\to A$ realizes the inclusion of units $\pi_0(A)^{\times}\to\pi_0(A)$.\qed
\end{lemma}
\begin{lemma}
Let $A$ be a ${\mathcal J}$-space monoid and let $A\to \bar A$ be a fibrant replacement. Then the following conditions are equivalent: (i) $\pi_0(A_{h{\mathcal J}})$ is a group, (ii) $\pi_0(\bar A)$ is a group, and (iii) $A^{\times}=A$.
\end{lemma}
\begin{proof}
By Corollary \ref{cor:J-hocolim-components} (i) implies (ii), by definition, (ii) implies (iii), and by the second description of $A^{\times}$ (iii) implies (i).
\end{proof}
A ${\mathcal J}$-space monoid $A$ is said to
be \emph{grouplike} if it satisfies the equivalent conditions in the above lemma.
\begin{proposition}\label{prop:J-grouplike-adjunction}
The functor $A\mapsto A^{\times}$ from ${\mathcal J}$-space monoids to grouplike
${\mathcal J}$-space monoids is a right adjoint of the inclusion functor.
\end{proposition}
\begin{proof}
Let $A\to \bar A$ be a fibrant replacement and notice that the induced map of units $A^{\times}\to \bar A^{\times}$ is the pullback along the inclusion $\bar A^{\times}\to \bar A$. This implies that $A^{\times}\to\bar A^{\times}$ is also a fibrant replacement and hence, by Lemma~\ref{lem:unit-inclusion}, that
$A^{\times}$ is indeed grouplike.
Consequently $(A^{\times})^{\times}=A^{\times}$ which gives the adjunction statement.
\end{proof}
\subsection{${\mathcal J}$-spaces and symmetric spectra}
An ordinary $\mathbb Z$-graded ring has an underlying graded signed monoid obtained by forgetting the additive structure except for the action of $\{\pm 1\}$ in each degree.
There are pairs of adjoint functors
\[
\left\{\begin{gathered}
\text{graded (comm.)}\\
\text{signed groups}
\end{gathered}\right\}
\rightleftarrows
\left\{\begin{gathered}
\text{graded (comm.)}\\
\text{signed monoids}
\end{gathered}\right\}
\rightleftarrows
\left\{\begin{gathered}
\text{graded (comm.)}\\
\text{rings}
\end{gathered}\right\}.
\]
and we wish to model a topological version of this using ${\mathcal J}$-space monoids.
The category ${\mathcal J}$ is designed to be an organizing device for maps
between all the free symmetric spectra on spheres
$F_{n_1}(S^{n_2})$, just as ${\mathcal I}$ encodes maps between the symmetric spectra $F_n(S^n)$ for different
$n$. By adjunction, maps of symmetric spectra
$F_{n_1}(S^{n_2})\to F_{m_1}(S^{m_2})$ are in
one-to-one correspondence with maps $S^{n_2} \to F_{m_1}(S^{m_2})_{n_1}$. For a morphism
$(\beta_1,\beta_2,\sigma)\colon(\bld{m_1},\bld{m_2})\to(\bld{n_1},\bld{n_2})$ in ${\mathcal J}$, let
\begin{equation}\label{eq:FS-induced-maps} (\beta_1,\beta_2,\sigma)^*
\colon F_{n_1}(S^{n_2})\to F_{m_1}(S^{m_2})
\end{equation} be the map that is adjoint to
\[ S^{\bld{n_2}} \xleftarrow{\cong} S^{\bld{m_2}} \wedge S^{\bld{n_2}-\beta_2}
\xleftarrow{\cong} S^{\bld{m_2}}\wedge S^{\bld{n_1}-\beta_1}
\hookrightarrow
\bigvee_{\beta\colon\bld{m_1}\to\bld{n_1}}S^{\bld{m_2}}\wedge
S^{\bld{n_1}-\beta}.\]
The first map is the isomorphism \eqref{eq:reindexing-basic} induced by
$\beta_2$, the second is the isomorphism induced by $\sigma$, and the last map is the inclusion of $S^{\bld m_2}\wedge S^{\bld n_1-\beta_1}$ as the wedge summand indexed by $\beta_1$.
\begin{lemma} \label{lem:J-and-free-spsym}
With the maps defined in \eqref{eq:FS-induced-maps},
\[ F_{-}(S^{-})\colon {\mathcal J}^{{\mathrm{op}}} \to {\mathrm{\textrm{Sp}^{\Sigma}}}, \quad
(\bld{n_1},\bld{n_2})\mapsto F_{n_1}(S^{n_2}) \] is a strong symmetric monoidal functor.
\end{lemma}
\begin{proof}
We first check that $F_{-}(S^{-})$ actually defines a functor. Consider a morphism $(\beta_1,\beta_2,\sigma)
\colon(\bld{m_1},\bld{m_2})\to(\bld{n_1},\bld{n_2})$ and the induced map
\[
(\beta_1,\beta_2,\sigma)^*\colon \bigvee_{\gamma\colon\bld{n_1}\to \bld p}
S^{\bld{n_2}}\wedge S^{\bld p-\gamma}\to\bigvee_{\delta\colon\bld{m_1}\to
\bld p}S^{\bld{m_2}}\wedge S^{\bld p-\delta}
\]
in spectrum degree $p$. This map takes the wedge summand $S^{\bld{n_2}}\wedge S^{\bld p-\gamma}$ indexed by
$\gamma\colon\bld{n_1}\to \bld{p}$ to the wedge summand $S^{\bld{m_2}}\wedge S^{\bld p-\gamma\beta_1}$ indexed by $\gamma\beta_1\colon\bld{m_1}\to \bld p$. On these wedge summands
the isomorphism
$S^{\bld{n_2}}\wedge S^{\bld p-\gamma}\to S^{\bld{m_2}}\wedge
S^{\bld p-\gamma\beta_1}$ is then induced by the chain of bijections
\[
\bld{n_2}\concat(\bld p-\gamma) \leftarrow \bld{m_2}\concat(\bld{n_2}-\beta_2)\concat (\bld p-\gamma)\leftarrow\bld{m_2}\concat (\bld{n_1}-\beta_1)\concat(\bld p-\gamma)\to\bld{m_2}\concat(\bld p-\gamma\beta_1).
\]
The first bijection is induced by $\beta_2$ (as in \eqref{eq:reindexing-basic}), the second by $\sigma$, and the third by $\gamma$ (as in \eqref{eq:reindexing-composite}). For a composable pair of morphisms
\[
(\bld{l_1},\bld{l_2})\xrightarrow{(\alpha_1,\alpha_2,\rho)}
(\bld{m_1},\bld{m_2})\xrightarrow{(\beta_1,\beta_2,\sigma)}(\bld{n_1},\bld{n_2})
\]
there is a commutative diagram of bijections
\[
\xymatrix{
& \bld{m_2}\concat(\bld{p}-\gamma\beta_1)\ar[dr]^{(\alpha_1,\alpha_2,\rho)^*}
& \\
\bld{n_2}\concat(\bld{p}-\gamma)\ar[ur]^{(\beta_1,\beta_1,\sigma)^*}
\ar[rr]^{(\beta_1\alpha_1,\beta_2\alpha_2,
\sigma\cup \beta_2\rho\beta_1^{-1})^*}& &
\bld{l_2}\concat(\bld{p}-\gamma\beta_1\alpha_1)
}
\]
which gives the functoriality of $F_{-}(S^{-})$. The required isomorphism making $F_{-}(S^{-})$ a strong symmetric monoidal functor is provided by
\eqref{eq:smash_of_free_iso}. Checking from the explicit descriptions of the relevant maps given above and in \eqref{eq:smash_of_free_iso}, one sees that this is indeed a natural transformation when both sides are viewed as functors on ${\mathcal J}\times {\mathcal J}$. Furthermore, it follows from the explicit
description in \eqref{eq:symmetry_iso_spsym} that the isomorphism
\eqref{eq:smash_of_free_iso} is compatible with the symmetry isomorphisms for ${\mathcal J}$ and ${\mathrm{\textrm{Sp}^{\Sigma}}}$.
\end{proof}
Let $X$ be a ${\mathcal J}$-space and let
$\mathbb S^{{\mathcal J}}[X]$ be the symmetric spectrum defined as the coend of the (${\mathcal J}^{{\mathrm{op}}}\times{\mathcal J}$)-diagram $F_{m_1}(S^{m_2})\wedge
X(\bld n_1,\bld n_2)_+$. Given a symmetric spectrum $E$ we write
$\Omega^{{\mathcal J}}(E)$ for the ${\mathcal J}$-space defined by
$\Map_{{\mathrm{\textrm{Sp}^{\Sigma}}}}(F_{-}(S^{-}),E)$.
An application of Proposition~\ref{prop:structured-adjunction-SK-Spsym} to the
strong symmetric monoidal functor $F_{-}(S^{-})$ provides the two adjoint
pairs of functors
\begin{equation}\label{eq:cScJ-Spsym-adjunction-copied}
{\mathbb S}^{{\mathcal J}}[-] \colon
{\mathcal S}^{{\mathcal J}} \rightleftarrows {\mathrm{\textrm{Sp}^{\Sigma}}} \colon
\Omega^{{\mathcal J}}
\qquad \text{and} \qquad
{\mathbb S}^{{\mathcal J}}[-] \colon {\mathcal C}{\mathcal S}^{{\mathcal J}} \rightleftarrows {\mathcal C}{\mathrm{\textrm{Sp}^{\Sigma}}} \colon
\Omega^{{\mathcal J}}.
\end{equation}
Checking from the definitions, we find that
\begin{equation}
{\mathbb S}^{{\mathcal J}}[X]_n=\bigvee_{k\geq 0} S^k\wedge_{\Sigma_k}
X(\bld n,\bld k)_+
\quad\textrm{ and } \quad \Omega^{{\mathcal J}}(E)(\bld{n_1},\bld{n_2})=\Omega^{n_2}(E_{n_1}).
\end{equation}
The next result can be deduced from the general criterion in
Proposition~\ref{prop:K-spaces-Spsym-adjunction} in the same way as the ${\mathcal I}$-space analogue in Proposition~\ref{prop:cScI-Spsym-adjunction}.
\begin{proposition}
\begin{enumerate}[(i)]
\item
The first adjunction in \eqref{eq:cScJ-Spsym-adjunction-copied} is a Quillen adjunction with respect to the (positive) projective ${\mathcal J}$-model structure on
${\mathcal S}^{{\mathcal J}}$ and the (positive) projective stable model structure on ${\mathrm{\textrm{Sp}^{\Sigma}}}$.
\item
The second adjunction in \eqref{eq:cScJ-Spsym-adjunction-copied} is a Quillen adjunction with respect to the positive projective ${\mathcal J}$-model structure on ${\mathcal C}{\mathcal S}^{{\mathcal J}}$ and the positive projective stable model structure on
${\mathcal C}{\mathrm{\textrm{Sp}^{\Sigma}}}$.
\qed
\end{enumerate}
\end{proposition}
Our definition of the signed monoid of components associated to a
positive fibrant ${\mathcal J}$-space monoid is partly motivated by the next result.
\begin{proposition}\label{prop:realizing-homotopy-groups}
Let $R$ be a (positive) fibrant symmetric ring spectrum. Then the
underlying graded signed monoid of the ring $\pi_*(R)$ is isomorphic
to $\pi_0(\Omega^{{\mathcal J}}(R))$.
\end{proposition}
\begin{proof}
For $p \in {\mathbb Z}$ the $p$th stable homotopy group of $R$ is
$\pi_p(R) = \colim_{u} \pi_{p+u}(R_u)$
where $p+u \geq 2$ and the colimit is taken over the
maps
\[ \pi_{p+u}(R_u) \to \pi_{p+u+1} (R_u \wedge S^1) \to \pi_{p+u+1}
(R_{u+1}). \] Setting $m_1 = u$ and $m_2 = p+u$, the chain of
isomorphisms
\[ \pi_{p+u}(R_u) \cong \pi_0(\Omega^{p+u}R_u) \cong
\pi_0(\Omega^{{\mathcal J}}(R)(\bld{m_1},\bld{m_2})) \] identifies the terms in
the colimit system defining $\pi_p(R)$ with the terms in the colimit system
defining $\pi_{0,p}(\Omega^{{\mathcal J}}(R))$. Since the isomorphisms are
compatible with the structure maps in the colimit system, we get an
isomorphism $\pi_{p}(R) \to \pi_{0,p}(\Omega^{{\mathcal J}}(R))$.
Because $R$ is positive fibrant, it is in particular semistable,
and~\cite[4.1 Theorem]{Schwede-homotopy_sym} implies that the action
of $\sigma \in\Sigma_u$ on $\pi_{p+u} (R_u)$ induced by the
$\Sigma_u$-action on $R_u$ coincides with the action of
$\sgn(\sigma)$ on the abelian group $\pi_{p+u}(R_u)$. This
implies that the $\{\pm1\}$-actions on $\pi_*(R)$ and
$\pi_0(\Omega^{{\mathcal J}}(R))$ coincide.
Let $[x]\in \pi_p(R)$ and $[y]\in\pi_q(R)$ be represented by maps
$x \colon S^{p+u} \to R_u$ and $y \colon S^{q+v} \to R_v$. It is shown
in~\cite[Proposition I.6.21]{Schwede-SymSp} that the product $[x][y] \in \pi_{p+q}(R)$ is
represented by $(-1)^{uq}[x\wedge y]$, where $x\wedge y$ is the
map $S^{p+u+q+v}\to R_{u+v}$ obtained by smashing the representatives
and using the multiplication in $R$.
Setting $m_1 = u, m_2 = p+u, n_1 = v,$ and $n_2 = q+v$, the above
isomorphism maps $(-1)^{uq}[x\wedge y]$ to $(-1)^{m_1(n_2-n_1)}\mu([x],[y])$.
This proves that the products coincide.
\end{proof}
\begin{definition}
Let $R$ be a symmetric ring spectrum.
The \emph{graded units} of $R$ is the grouplike ${\mathcal J}$-space monoid
$\GL^{\cJ}_1(R) = (\Omega^{{\mathcal J}}R)^{\times}$.
\end{definition}
The functor $R\mapsto \GL^{\cJ}_1(R)$ is defined for all symmetric ring spectra $R$ and provides a right adjoint of the functor that to a (commutative) grouplike ${\mathcal J}$-space monoid $A$ associates the (commutative) symmetric ring spectrum $\mathbb S^{{\mathcal J}}[A]$. However, as for the
${\mathcal I}$-space units, $\GL^{\cJ}_1(R)$ only represents the ``correct'' homotopy type when $R$ is positive fibrant.
In this case we have the following consequence of Lemma \ref{lem:unit-inclusion} and Proposition \ref{prop:realizing-homotopy-groups}.
\begin{proposition}
Let $R$ be a (positive) fibrant symmetric ring spectrum. Then the map of graded signed monoids $\pi_0(\GL^{\cJ}_1(R))\to \pi_0(\Omega^{{\mathcal J}}(R))$ realizes the inclusion of graded units $\pi_*(R)^{\times}\to \pi_*(R)$. \qed
\end{proposition}
\subsection{The flat model structure on ${\mathcal J}$-spaces}
\label{subsec:flat-J-model}
We here introduce the ${\mathcal J}$-space analogue of the flat model structure on
${\mathcal I}$-spaces. Given an object $(\bld n_1,\bld n_2)$ in ${\mathcal J}$, let
$\partial({\mathcal J}\!\downarrow(\bld n_1,\bld n_2))$ be the full subcategory of the comma category $({\mathcal J}\!\downarrow(\bld n_1,\bld n_2))$ obtained by excluding the objects $(\bld m_1,\bld m_2)\to(\bld n_1,\bld n_2)$ that are isomorphisms in ${\mathcal J}$. The $(\bld n_1,\bld n_2)$th latching space $L_{(\bld n_1,\bld n_2)}(X)$ of a ${\mathcal J}$-space $X$ is the colimit of the $\partial({\mathcal J}\!\downarrow(\bld n_1,\bld n_2))$-diagram obtained by composing $X$ with the forgetful functor from $\partial({\mathcal J}\downarrow(\bld n_1,\bld n_2))$ to ${\mathcal J}$. A map of ${\mathcal J}$-spaces $X\to Y$ is a \emph{flat cofibration} if the induced map
\[
X(\bld n_1,\bld n_2)\cup_{L_{(\bld n_1,\bld n_2)}(X)}
L_{(\bld n_1,\bld n_2)}(Y)\to Y(\bld n_1,\bld n_2)
\]
is a cofibration in ${\mathcal S}$ for all objects $(\bld n_1,\bld n_2)$. It is a
\emph{positive flat cofibration} if in addition $X(\bld 0,\bld n_2)\to
Y(\bld 0,\bld n_2)$ is an isomorphism for all objects $\bld n_2$ in ${\mathcal I}$.
We say that a map of ${\mathcal J}$-spaces is a (positive) flat ${\mathcal J}$-fibration if it has the right lifting property with respect to the class of (positive) flat cofibrations that are ${\mathcal J}$-equivalences. As in the case of ${\mathcal I}$-spaces it is a consequence of our general theory of diagram spaces that together with the
${\mathcal J}$-equivalences these classes of maps specify a cofibrantly generated simplicial model structure which is proper, monoidal, and satisfies the monoid axiom. We refer to this as the (positive) flat model structure and the cofibrant objects will be called flat ${\mathcal J}$-spaces. By
Proposition~\ref{prop:well-structured-comparison} the identity functor on
${\mathcal S}^{{\mathcal J}}$ is the left adjoint in a Quillen equivalence from the (positive) projective ${\mathcal J}$-model structure to the (positive) flat ${\mathcal J}$-model structure. The next proposition can be derived from our general theory of diagram spaces in the same way as the ${\mathcal I}$-space analogue in
Proposition~\ref{prop:I-underlying-flat}.
\begin{proposition}\label{prop:J-underlying-flat}
\begin{enumerate}[(i)]
\item
The positive flat model structure on ${\mathcal S}^{{\mathcal J}}$ lifts to a cofibrantly generated proper simplicial model structure on ${\mathcal C}{\mathcal S}^{{\mathcal J}}$.
\item
Suppose that $A$ is a commutative ${\mathcal J}$-space monoid which is cofibrant in the lifted model structure in (i). Then the underlying ${\mathcal J}$-space of $A$ is flat.
\qed
\end{enumerate}
\end{proposition}
By Proposition \ref{prop:boxtimes-flat-invariance} this has the following important implication: If a commutative ${\mathcal J}$-space monoid $A$ is cofibrant in the positive flat or projective ${\mathcal J}$-model structures on ${\mathcal C}{\mathcal S}^{{\mathcal J}}$, then the endofunctor
$A\boxtimes(-)$ preserves ${\mathcal J}$-equivalences.
\begin{remark}
The adjunctions in \eqref{eq:cScJ-Spsym-adjunction-copied} fail to be Quillen adjunctions with respect to the flat model structures. The reason is that the action of $\Sigma_{n_1}\times\Sigma_{n_2}$ on $F_{n_1}(S^{n_2})$ is not free so that condition (ii) in Proposition~\ref{prop:K-spaces-Spsym-adjunction} does not hold. However, there is a ``semi-flat'' model structure on ${\mathcal S}^{{\mathcal J}}$ which is compatible with the flat model structure on
${\mathrm{\textrm{Sp}^{\Sigma}}}$: In the notation of Proposition~\ref{prop:K-model-str}
this is defined by letting the subcategory of automorphisms ${\mathcal A}$ be the subgroups $\Sigma_{n_1}\times\{1_{n_2}\}$ of
$\Sigma_{n_1}\times\Sigma_{n_2}$.
\end{remark}
\subsection{An application to topological logarithmic structures}
Commutative ${\mathcal J}$-space monoids can be used to define a graded version
of John Rognes' notion of topological logarithmic structures
introduced in~\cite{Rognes-TLS}. We explain how this can be done for
the basic definitions by an almost verbatim translation of Rognes'
terminology to the context of ${\mathcal J}$-spaces. One advantage of graded
logarithmic structures on commutative symmetric ring spectra is that
they enable us to see the difference between a periodic ring spectrum
and its connective cover. For motivation and background, we refer the
reader to \cite{Rognes-TLS}. More results about the graded log
structures introduced here can be found
in~\cite{Sagave_log-on-k-theory}.
We start by introducing the graded analogue of a pre-log symmetric ring
spectrum~\cite[Definition 7.1]{Rognes-TLS}:
\begin{definition}
Let $A$ be a commutative symmetric ring spectrum. A \emph{graded
pre-log structure} on $A$ is a pair $(M,\alpha)$ consisting of a
commutative ${\mathcal J}$-space monoid $M$ and a map
$\alpha \colon M \to \Omega^{\cJ}(A)$ of commutative ${\mathcal J}$-space
monoids. A \emph{graded pre-log symmetric ring spectrum}
$(A,M,\alpha)$ is a commutative symmetric ring spectrum $A$ with a
graded pre-log structure $(M,\alpha)$. A map $(f,f^{\flat}) \colon
(A,M,\alpha) \to (B,N,\beta)$ of graded pre-log symmetric ring
spectra consists of a map $f\colon A \to B$ of commutative symmetric
ring spectra and a map $f^{\flat}\colon M \to N$ of commutative
${\mathcal J}$-space monoids such that $\Omega^{\cJ}(f) \alpha = \beta
f^{\flat}$.
\end{definition}
\begin{example}
Let $A$ be a commutative symmetric ring spectrum and let $x$ be a
point in $\Omega^{\cJ}(A)(\bld{n}_1,\bld{n}_2)$. By adjunction, $x$ gives
rise to a map
\[\alpha \colon M = {\mathbb C} F_{(\bld{n}_1,\bld{n}_2)}^{{\mathcal J}}(\pt) = \textstyle
\coprod_{i\geq 0}
\big(F_{(\bld{n}_1,\bld{n}_2)}^{{\mathcal J}}(\pt)\big)^{\boxtimes i}/\Sigma_i \to
\Omega^{\cJ}(A).\] from the free commutative ${\mathcal J}$-space monoid on a point
in degree $(\bld{n_1},\bld{n_2})$. We refer to $(M,\alpha)$ as the
free graded pre-log structure generated by $x$.
\end{example}
For a graded pre-log symmetric ring spectrum $(A,M,\alpha)$, the
commutative ${\mathcal J}$-space monoid $\alpha^{-1}(\GL^{\cJ}_1(A))$ is defined
by the pullback diagram
\[\xymatrix@-1pc{
\alpha^{-1}(\GL^{\cJ}_1(A))
\ar[d]_{\widetilde{\iota}}
\ar[r]^-{\widetilde{\alpha}}&
\GL^{\cJ}_1(A)\ar[d]_{\iota}\\
M \ar[r]^-{\alpha}& \Omega^{\cJ}(A).}\]
This enables us to state the analogue of~\cite[Definition 7.4]{Rognes-TLS}:
\begin{definition}
A graded pre-log structure $(M,\alpha)$ on $A$ is a \emph{graded
log structure} if the induced map $\widetilde{\alpha}\colon
\alpha^{-1}(\GL^{\cJ}_1(A)) \to \GL^{\cJ}_1(A)$ is a ${\mathcal J}$-equivalence. A
graded pre-log symmetric ring spectrum $(A,M,\alpha)$ is a {\em
graded log symmetric ring spectrum} if $(M,\alpha)$ is a graded
log structure.
\end{definition}
The basic example of a graded log structure on $A$ is the {\em
trivial graded log structure} $(\GL^{\cJ}_1(A),\iota)$. The {\em
logification} of a graded pre-log structure is the pushout $M^{a}$
of the diagram
\[ M \xleftarrow{\widetilde{\iota}} \alpha^{-1}(\GL^{\cJ}_1(A))
\xrightarrow{\widetilde{\alpha}}\GL^{\cJ}_1(A) \] together with the induced map
$\alpha^{a} \colon M^{a} \to \Omega^{\cJ}(A)$. It comes with a map
$(M,\alpha)\to(M^{a},\alpha^{a})$. As in~\cite[Lemma 7.7]{Rognes-TLS},
one can show that $(M^{a},\alpha^{a})$ is indeed a log structure.
\begin{lemma}\label{lem:criterion-trivial-log}
Let $(A,M,\alpha)$ be a graded pre-log symmetric ring spectrum. If
$\alpha$ factors as a composite
\[ M \xrightarrow{\zeta} \GL^{\cJ}_1(A) \xrightarrow{\iota} \Omega^{\cJ}(A)\]
with $\zeta$ a map of commutative ${\mathcal J}$-space monoids, then
$(M^{a},\alpha^{a})$ is isomorphic to the trivial log-structure.
\end{lemma}
\begin{proof}
If $\zeta$ exists, then $\alpha^{-1}(\GL^{\cJ}_1(A))\cong M$, hence
$M^a \cong \GL^{\cJ}_1(A)$. This uses that $\GL^{\cJ}_1(A) \to
\Omega^{\cJ}(A)$ is an inclusion of path components and hence a
monomorphism.
\end{proof}
\begin{example}
Let $KU$ be a positive fibrant model for the periodic complex
$K$-theory spectrum, and let $f\colon ku \to KU$ be a positive
fibration exhibiting the connective complex $K$-theory spectrum $ku$
as the connective cover of $KU$. The Bott class $\pi_2(ku)$ can be
represented by a point $x \in \Omega^{\cJ}(ku)(\bld{1},\bld{3})$. It
generates a graded pre-log structure
\[ \alpha \colon M = {\mathbb C} F_{(\bld{1},\bld{3})}^{{\mathcal J}}(\pt) \to
\Omega^{\cJ}(ku)\] which gives rise to a non-trivial log structure on
$ku$. We compose $\alpha$ with $\Omega^{\cJ}(f)$ in order to get the
induced (inverse image) graded pre-log structure $(f^*M,f^*\alpha)$
on $KU$. The map $f^* \alpha$ factors through the inclusion of the
units of $KU$ since the Bott element is invertible in
$\pi_*(KU)$. By Lemma \ref{lem:criterion-trivial-log}, the
logification of $(f^*M,f^*\alpha)$ is the trivial log structure.
In other words, the graded log structure generated by the Bott
element is trivial on $KU$ and non-trivial on $ku$. This is an
important feature of graded log structures which can only be
achieved using the graded units. In the ${\mathcal I}$-space case, the map
$\GL^{\cI}_1(ku) \to \GL^{\cI}_1(KU)$ is an equivalence, and the Bott class
being a unit is not detected by $\GL^{\cI}_1(KU)$. This issue is
discussed in~\cite[Remark 7.28]{Rognes-TLS}.
\end{example}
\begin{example}
As pointed out in~\cite[Remark 7.28]{Rognes-TLS}, there is another
source of interesting graded log structures. Let $E$ be a periodic
commutative symmetric ring spectrum and let $i\colon e \to E$ be its
connective cover. On $e$, one can form the \emph{direct image} graded
log structure $i_*(\GL^{\cJ}_1(E))$ of the trivial graded log
structure on $E$. It is defined to be the pullback of
\[ \Omega^{\cJ}(e) \to \Omega^{\cJ}(E) \leftarrow \GL^{\cJ}_1(E).\]
This is very much analogous to the situation in algebra, where a
discrete valuation ring $A$ inherits the log structure $A \setminus
\{0\}$ as the direct image of the trivial log structure on its
fraction field, compare for example~\cite[Remark
2.25]{Rognes-TLS}. Again, forming $i_*(\GL^{\cJ}_1(E))$ gives
something non-trivial in the graded case, while the same
construction in the ${\mathcal I}$-space case only leads to the trivial log
structure.
\end{example}
\section{Well-structured index categories}\label{sec:well-structured}
As we have seen in Sections \ref{sec:I-space-section} and \ref{sec:J-space-section} there are several useful model structures on the categories of ${\mathcal I}$- and ${\mathcal J}$-spaces. In order to set up and analyze these model structures in a common framework we here introduce the notion of a \emph{well-structured relative index category}. The main idea can be summarized as follows. Consider in general a small symmetric monoidal category ${\mathcal K}$ and the category of ${\mathcal K}$-spaces ${\mathcal S}^{{\mathcal K}}$. When defining a model structure
on ${\mathcal S}^{{\mathcal K}}$ we have a choice when specifying the requirements for a map of ${\mathcal K}$-spaces $X\to Y$ to be a fibration: For each object $\bld k$ in ${\mathcal K}$ the map $X(\bld k)\to Y(\bld k)$ is equivariant with respect to the action of the endomorphisms of $\bld k$ and we must specify the extend to which these maps are equivariant fibrations. This of course has a dual effect on the cofibrations; the stronger the condition for a map to be a fibration the weaker the condition is to be a cofibration. In practice we shall control this duality by specifying a subcategory ${\mathcal A}$ of automorphisms in
${\mathcal K}$ and require that the fibrations $X(\bld k)\to Y(\bld k)$ be equivariant with respect to the automorphisms in ${\mathcal A}$.
\subsection{Well-structured relative index categories}
Let $({\mathcal K},\sqcup,\bld 0)$ be a small symmetric monoidal category and let ${\mathcal A}$ be a subcategory of automorphisms. Given an object $\bld k$ in ${\mathcal A}$ we write ${\mathcal A}(\bld k)$ for the automorphism group ${\mathcal A}(\bld k,\bld k)$. We shall always assume that ${\mathcal A}$ be a \emph{normal subcategory}: for each isomorphism $\alpha\colon \bld k\to\bld l$ in ${\mathcal K}$ we require that $\bld k$ belongs to ${\mathcal A}$ if and only if $\bld l$ does, and that in this case conjugation by $\alpha$ gives an isomorphism ${\mathcal A}(\bld k)\to {\mathcal A}(\bld l)$ by mapping
$\gamma$ in ${\mathcal A}(\bld k)$ to $\alpha\gamma \alpha^{-1}$. We shall also require that
${\mathcal A}$ be \emph{multiplicative} in the sense that the monoidal structure
$\sqcup\colon{\mathcal K}\times {\mathcal K}\to {\mathcal K}$ restricts to a functor ${\mathcal A}\times {\mathcal A}\to {\mathcal A}$ (but we do not assume that ${\mathcal A}$ necessarily contains the unit $\bld 0$ for the monoidal structure).
Let ${\mathbb N}_0$ denote the ordered set of natural numbers $0\to1\to2\to\dots $, thought of as a symmetric monoidal category via the additive structure.
\begin{definition}\label{def:well-structured-relative-index}
A \emph{well-structured relative index category} is a triple consisting of a
symmetric monoidal category $({\mathcal K},\concat, \bld 0)$, a strong
symmetric monoidal functor $\lambda\colon {\mathcal K}\to {\mathbb N}_0$, and a normal and multiplicative subcategory of automorphisms ${\mathcal A}$ in ${\mathcal K}$. These data are required to satisfy the following conditions.
\begin{enumerate}
\item[(i)]
A morphism $\bld k\to \bld l$ in ${\mathcal K}$ is an isomorphism if and only if
$\lambda(\bld k)=\lambda(\bld l)$.
\item[(ii)]
For each object $\bld k$ in ${\mathcal A}$ and each object $\bld l$ in ${\mathcal K}$, each connected component of the category $(\bld k\concat -\downarrow \bld l)$ has a terminal object.
\item[(iii)]
For each object $\bld k$ in ${\mathcal A}$ and each object $\bld l$ in ${\mathcal K}$, the canonical right action of
${\mathcal A}(\bld k)$ on the category $(\bld k\concat -\downarrow \bld l)$ induces a free action on the set of connected components.
\item[(iv)]
Let ${\mathcal K}_{{\mathcal A}}$ be the full subcategory of ${\mathcal K}$ generated by the objects in
${\mathcal A}$. We require that the inclusion ${\mathcal K}_{{\mathcal A}}\to {\mathcal K}$ is homotopy cofinal.
\end{enumerate}
\end{definition}
Spelling the requirements out in more detail, the condition that $\lambda$ be strong symmetric monoidal means that $\lambda(\bld 0)=0$ and
$\lambda(\bld k_1\concat \bld k_2)=\lambda(\bld k_1)+\lambda(\bld k_2)$ for all pairs of objects $\bld k_1$ and $\bld k_2$. We think of $\lambda$ as a \emph{degree functor} on ${\mathcal K}$. The existence of such a degree functor ensures an explicit description of the cofibrations in the model structures we define on ${\mathcal S}^{{\mathcal K}}$, see Proposition \ref{prop:latching-characterization}.
In (ii) and (iii) the category $(\bld k\concat -\downarrow\bld l)$ is the comma category whose objects are pairs $(\bld n,\alpha)$ given by an object $\bld n$ in ${\mathcal K}$ and a morphism $\alpha\colon \bld k\concat \bld n\to\bld l$ in ${\mathcal K}$. A morphism $(\bld n,\alpha)\to (\bld n',\alpha')$ is specified by a morphism
$\gamma\colon \bld n\to \bld n'$ in ${\mathcal K}$ such that
$\alpha=\alpha'\circ(1_{\bld k}\concat \gamma)$. For the homotopy cofinality condition (iv) we recall that in general a subcategory ${\mathcal B}$ of a small category ${\mathcal C}$ is said to be homotopy cofinal if the comma categories $(c\downarrow {\mathcal B})$ have contractible classifying space for each object $c$ in ${\mathcal C}$.
This is equivalent to the condition that for any ${\mathcal C}$-diagram $X$ in ${\mathcal S}$, the canonical map $\hocolim_{{\mathcal B}}X\to\hocolim_{{\mathcal C}}X$ is a weak homotopy equivalence, see e.g.\ \cite[Theorem~19.6.13]{Hirschhorn-model}.
The degree functor $\lambda$ on ${\mathcal K}$ will usually be understood from the
context and we often use the notation $({\mathcal K},{\mathcal A})$ to indicate a well-structured relative index category. Notice that by condition (i) each endomorphism set
${\mathcal K}(\bld k,\bld k)$ is a group of automorphisms. We introduce the notation
${\mathcal K}(\bld k)$ for this group. The normality condition on
${\mathcal A}$ in particular implies that ${\mathcal A}(\bld k)$ is a normal subgroup of
${\mathcal K}(\bld k)$ for each object $\bld k$ in ${\mathcal A}$.
We shall later prove that a well-structured relative index category $({\mathcal K},{\mathcal A})$ gives rise to an ${\mathcal A}$-relative ${\mathcal K}$-model structure on ${\mathcal S}^{{\mathcal K}}$ and that this model structure is proper, monoidal, and lifts to the category of structured ${\mathcal K}$-spaces for any $\Sigma$-free operad. However, in order to lift this model structure to commutative ${\mathcal K}$-space monoids we need $({\mathcal K},{\mathcal A})$ to be \emph{very well-structured} in the following sense.
\begin{definition}\label{def:very-well-structured}
A well-structured relative index category $({\mathcal K},{\mathcal A})$ is \emph{very well-structured} if for each object $\bld k$ in ${\mathcal A}$, each object $\bld l$ in ${\mathcal K}$,
and each $n\geq 1$, the canonical right action of
$\Sigma_n\ltimes {\mathcal A}(\bld k)^{\times n}$ on the category
$(\bld k^{\sqcup n} \sqcup-\downarrow \bld l)$ induces a free action on the set of connected components.
\end{definition}
Here the group $\Sigma_n\ltimes {\mathcal A}(\bld k)^{\times n}$ is the semidirect product of $\Sigma_n$ acting from the right on ${\mathcal A}(\bld k)^{\times n}$ (also known as the wreath product $\Sigma_n\int {\mathcal A}(\bld k)$). The action on
$(\bld k^{\sqcup n}\sqcup -\downarrow \bld l)$ is via the homomorphism
$\Sigma_n\ltimes {\mathcal A}(\bld k)^{\times n}\to {\mathcal K}(\bld k^{\bld n})$ that
maps an element $(\sigma;f_1,\dots,f_n)$ with $\sigma\in \Sigma_n$ and
$f_i\in {\mathcal A}(\bld k)$ to the composition
$\sigma_*\circ(f_1\concat\dots\concat f_n)$ where $\sigma_*$ denotes the canonical automorphism of $\bld k^{\concat n}$ determined by the symmetric monoidal structure.
\begin{remark}
The stronger condition that $\Sigma_n\ltimes {\mathcal A}(\bld k)^{\times n}$ maps injectively into ${\mathcal A}(\bld k^{\sqcup n})$ is relevant for whether the forgetful functor from commutative ${\mathcal K}$-space monoids to ${\mathcal K}$-spaces preserves cofibrancy. We discuss this in Section~\ref{sec:structured-cofibrant}.
\end{remark}
Specializing to the case where ${\mathcal A}$ is the discrete subcategory of identity morphisms in ${\mathcal K}$ we get the notion of a \emph{well-structured index category}. Writing out the details of this we arrive at the following definition.
\begin{definition}\label{def:well-structured-index}
A \emph{well-structured index category} ${\mathcal K}$ is a small symmetric monoidal category, equipped with a strong symmetric monoidal functor
$\lambda\colon{\mathcal K}\to \mathbb N_0$, such that
\begin{itemize}
\item
a morphism $\bld k\to \bld l$ in ${\mathcal K}$ is an isomorphism if and only if
$\lambda(\bld k)=\lambda(\bld l)$, and
\item
for each pair of objects $\bld k$ and $\bld l$ in ${\mathcal K}$, each connected component of the category $(\bld k\sqcup-\downarrow\bld l)$ has a terminal object.
\end{itemize}
\end{definition}
Given a well-structured index category ${\mathcal K}$ we shall refer to the associated model structure on ${\mathcal S}^{{\mathcal K}}$ as the \emph{projective} model structure, see Definition \ref{def:K-projective-model}. It should be noted that the axioms for the monoidal unit $\bld 0$ in the symmetric monoidal category ${\mathcal K}$ imply that the homomorphism
$\Sigma_n\to {\mathcal K}(\bld 0^{\sqcup n})$ is trivial for all $n$. Thus, to obtain a very well-structured relative index category and hence a model structure on commutative ${\mathcal K}$-space monoids, we are forced to specify a subcategory of automorphisms ${\mathcal A}$ that does not contain the object $\bld 0$.
Recall the free functors $F_{\bld k}^{{\mathcal K}}(*)$ introduced in
Section~\ref{subs:free-semi-free}.
When analyzing the homotopical properties of the $\boxtimes$-product on
${\mathcal S}^{{\mathcal K}}$ it will be important to consider ${\mathcal K}$-spaces of the form
$F_{\bld k}^{{\mathcal K}}(*)\boxtimes X$ for an object $\bld k$ and a ${\mathcal K}$-space
$X$. The axioms for a well-structured relative index category ${\mathcal K}$ are partly motivated by the following lemmas.
\begin{lemma}\label{lem:k-Kan-extension}
The ${\mathcal K}$-space $F_{\bld k}^{{\mathcal K}}(*)\boxtimes X$ is isomorphic to the left Kan extension of $X$ along the functor $\bld k\concat-\colon {\mathcal K}\to {\mathcal K}$.
\end{lemma}
\begin{proof}
In fact, this result holds for any monoidal category ${\mathcal K}$. The statement in the lemma means that there is a natural isomorphism
\begin{equation}\label{eq:left-Kan}
(F_{\bld k}^{{\mathcal K}}(*)\boxtimes X)(\bld l)\cong
\operatornamewithlimits{colim}_{\bld k\concat \bld k'\to \bld l}
X(\bld k')
\end{equation}
(where the colimit on the right hand side is over the category $(\bld k\concat -\downarrow \bld l)$) and this is immediate from the universal properties of these constructions.
\end{proof}
Notice in particular, that for $X=*$ the isomorphism \eqref{eq:left-Kan} gives an identification of $(F_{\bld k}^{{\mathcal K}}(*)\boxtimes *)(\bld l)$ with the set of connected components of the category $(\bld k\concat -\downarrow\bld l)$.
\begin{lemma}\label{lem:free-right-action}
Let $({\mathcal K},{\mathcal A})$ be a well-structured relative index category. Then
the canonical right action of ${\mathcal K}(\bld k)$ on $F_{\bld k}^{{\mathcal K}}(*)\boxtimes X$ restricts to a levelwise free ${\mathcal A}(\bld k)$-action for all objects $\bld k$ in
${\mathcal A}$.
\end{lemma}
\begin{proof}
The projection $X\to *$ onto the terminal ${\mathcal K}$-space induces a map
of ${\mathcal K}$-spaces $F_{\bld k}^{{\mathcal K}}(*)\boxtimes X\to F_{\bld k}^{{\mathcal K}}(*)
\boxtimes *$, so it suffices to check that the ${\mathcal A}(\bld k)$-action on the target is levelwise free. By the observation before the lemma
$(F_{\bld k}^{{\mathcal K}}(*)\boxtimes *)(\bld l)$ can be identified with the set of connected components of the category
$(\bld k\concat -\downarrow \bld l)$, hence the result follows from condition
(iii) for a well-structured relative index category.
\end{proof}
The principal examples in this paper are the categories ${\mathcal I}$ and ${\mathcal J}$.
We define degree functors
$\lambda\colon {\mathcal I}\to\mathbb N_0 $ by $\lambda(\bld{k}) =|\bld k|$ and
$\lambda\colon{\mathcal J}\to\mathbb N_0$ by $\lambda(\bld{k}_1,\bld{k}_2)=
|\bld k_1|$. (Here $|-|$ indicates the cardinality of a finite set). There are other degree functors on ${\mathcal J}$ but the above choice is the one that will be relevant for our work.
\begin{proposition}\label{prop:I-J-well-structured}
Let ${\mathcal K}$ denote one of the categories ${\mathcal I}$ or ${\mathcal J}$, equipped with the above degree functor. Suppose that ${\mathcal A}$ is a normal and multiplicative subcategory of automorphisms in ${\mathcal K}$ such that the inclusion ${\mathcal K}_{{\mathcal A}}\to {\mathcal K}$ is homotopy cofinal. Then $({\mathcal K},{\mathcal A})$ is a well-structured relative index category and if all objects of ${\mathcal A}$ have positive degree, then $({\mathcal K},{\mathcal A})$ is very well-structured.
\end{proposition}
\begin{proof}
We explain the details for ${\mathcal J}$; the case of ${\mathcal I}$ is similar but easier. It
is clear that $\lambda$ is strong symmetric monoidal and that (i) holds.
For (ii) we first observe that the correspondence $(\alpha_1,\alpha_2,\rho)\mapsto (\alpha_1|_{\bld k_1},\alpha_2|_{\bld k_2})$ defines a bijection between the set of connected components of the category
$((\bld k_1,\bld k_2)\sqcup-\downarrow (\bld l_1,\bld l_2))$
and the set ${\mathcal I}(\bld k_1,\bld l_1)\times {\mathcal I}(\bld k_2,\bld l_2)$. An object
$(\alpha_1,\alpha_2,\rho)$ is terminal in its connected component if and only it is an isomorphism in ${\mathcal J}$ which implies that (ii) holds. Given an object
$(\bld k_1,\bld k_2)$ in ${\mathcal J}$, the automorphism group can be identified with the product $\Sigma_{|\bld k_1|}\times \Sigma_{|\bld k_2|}$ and it is clear from the above description that this acts freely on the set of connected components in $((\bld k_1,\bld k_2)\sqcup-\downarrow (\bld l_1,\bld l_2))$, hence (iii) holds.
Finally, (iv) holds by the assumption on ${\mathcal A}$.
Now suppose that the objects in ${\mathcal A}$ have positive degree. In order to check the condition in Definition \ref{def:very-well-structured} for being very
well-structured we may as well assume that the automorphism groups in ${\mathcal A}$ are the full automorphism groups in ${\mathcal J}$. Using the above identification in terms of symmetric groups, the homomorphism from the semidirect product in question to the full automorphism group in ${\mathcal J}$ is given by concatenation and block permutation in each factor,
\[
\Sigma_n\ltimes (\Sigma_{|\bld k_1|}^{\times n}\times
\Sigma_{|\bld k_2|}^{\times n})\to
\Sigma_{n|\bld k_1|}\times \Sigma_{n|\bld k_2|}.
\]
This is injective if $|\bld k_1|>0$, so the conclusion follows from the observation above that the right hand side acts freely on the relevant set of connected components.
\end{proof}
In general, for a well-structured index category ${\mathcal K}$, we write ${\mathcal K}_+$ for the full subcategory whose objects have positive degree and $O{\mathcal K}_+$ for the corresponding discrete subcategory of identity morphisms.
\begin{corollary}\label{cor:I-J-well-structured}
Let ${\mathcal K}$ denote one of the categories ${\mathcal I}$ or ${\mathcal J}$. Then ${\mathcal K}$ is a
well-structured index category and $({\mathcal K},O{\mathcal K}_+)$ is very well-structured.
\end{corollary}
\begin{proof}
It remains to check that the inclusion ${\mathcal K}_+\to {\mathcal K}$ is homotopy cofinal. Thus, given an object $\bld k$ in ${\mathcal K}$ we must show that the comma category $(\bld k\downarrow {\mathcal K}_+)$ has contractible classifying space. Choose a morphism $\alpha\colon\bld 0\to\bld l$ in ${\mathcal K}$ such that $\bld l$ has positive degree and consider the functor $(\bld k\downarrow{\mathcal K})\to
(\bld k\downarrow {\mathcal K}_+)$ defined by concatenation with $\alpha$ on objects.
We also have the functor $(\bld k\downarrow {\mathcal K}_+)\to
(\bld k\downarrow {\mathcal K})$ defined by the inclusion of ${\mathcal K}_+$ in ${\mathcal K}$ and it is easy to see that $\alpha$ gives rise to natural transformations between the two compositions of these functors and the respective identity functors. Hence it suffices to show that the classifying space of $(\bld k\downarrow{\mathcal K})$ is
contractible and this is clear since the identity on $\bld k$ is an initial object.
\end{proof}
The above corollary is the underlying reason why the corresponding (positive) projective ${\mathcal I}$- and ${\mathcal J}$-model structures on ${\mathcal S}^{{\mathcal I}}$ and ${\mathcal S}^{{\mathcal J}}$ have the pleasant properties stated in Propositions \ref{prop:projective-I-model-structure} and \ref{prop:projective-J-model-structure}.
As discussed in Sections~\ref{subsec:flat-I-model} and
\ref{subsec:flat-J-model} it is important for the applications that the projective model structures are accompanied by corresponding \emph{flat} model structures. These flat model structures arise by specifying ${\mathcal A}$ as the full automorphism subcategory in ${\mathcal I}$ and ${\mathcal J}$, respectively. We write $\Sigma$ for the full automorphism subcategory of ${\mathcal I}$ (that is, the category of finite sets and
isomorphisms) and $\Sigma\times \Sigma$ for the full automorphism subcategory of ${\mathcal J}$. The corresponding automorphism categories of objects of positive degree are then given by $\Sigma_+$ and
$\Sigma_+\times \Sigma$.
\begin{corollary}\label{cor:flat-I-J-well-structured}
With the above notation there are well-structured relative index categories
$({\mathcal I},\Sigma)$ and
$({\mathcal J},\Sigma\times\Sigma)$. Restricting to automorphisms of objects of positive degree we get very well-structured relative index categories
$({\mathcal I},\Sigma_+)$ and $({\mathcal J},\Sigma_+\times\Sigma)$.
\end{corollary}
\begin{proof}
The homotopy cofinality condition (v) follows from the proof of
Corollary~\ref{cor:I-J-well-structured}.
\end{proof}
We finally consider the $K$-theory example mentioned in the introduction.
\begin{example}\label{ex:K-R-example}
Let $R$ be a ring with invariant basis number and let ${\mathcal F}_R$ be the category with objects the free $R$-modules $R^n$ and morphisms the isomorphisms between such modules. This is a permutative category under direct sum and we write
${\mathcal K}_R$ for Quillen's localization construction ${\mathcal F}_R^{-1}{\mathcal F}_R$ on this category,
see \cite{Grayson-higher}. In order to give an explicit description of this category we first introduce the category ${\mathcal S}{\mathcal F}_R$ of \emph{free split injections} (in Grayson's
notation \cite{Grayson-higher} this is the category
$\langle {\mathcal F}_R,{\mathcal F}_R\rangle$). The objects of ${\mathcal S}{\mathcal F}_R$ are the
$R$-modules $R^n$ and a morphism $(f,p)\colon R^m\to R^n$ is a pair of
$R$-linear maps $f\colon R^m\to R^n$ and $p\colon R^n\to R^m$ such that
$p\circ f$ is the identity on $R^m$ and the cokernel $R^n/\image(f)$ is free.
The category ${\mathcal K}_R$ has objects all pairs of free $R$-modules
$(R^{n_1},R^{n_2})$ and a morphism
\[
((f_1,p_1),(f_2,p_2),\rho)\colon
(R^{m_1},R^{m_2})\to
(R^{n_1},R^{n_2})
\]
is a triple given by morphisms $(f_i,p_i)\colon R^{m_i}\to R^{n_i}$ in
${\mathcal S}{\mathcal F}_R$ for $i=1,2$, and an isomorphism
$\rho\colon R^{n_1}/\image(f_1)\xrightarrow{\sim} R^{n_2}/\image(f_2)$ between the corresponding cokernels. Composition of morphisms is defined in the natural way.
We define a degree functor $\lambda\colon {\mathcal K}_R\to \mathbb N_0$ by
$\lambda(R^{n_1},R^{n_2})=n_1$ and claim that this makes ${\mathcal K}_R$ a well-structured index category. Since we assume that $R$ has invariant basis number it is clear that a morphism in ${\mathcal K}_R$ is an isomorphism if and only if the domain and codomain have the same degree (and hence are equal). For the second condition we observe that there is a bijective correspondence
\[
\left\{\text{components of
$((R^{k_1},R^{k_2})\oplus -\downarrow (R^{l_1},R^{l_2}))$}\right\}\simeq
{\mathcal S}{\mathcal F}_R(R^{k_1},R^{l_1})\times {\mathcal S}{\mathcal F}_R(R^{k_2},R^{l_2}).
\]
This takes an object defined by the data $(R^{n_i},(f_i,p_i)\colon R^{k_i}\oplus R^{n_i}\to R^{l_i})$ for $i=1,2$, and
$\rho\colon R^{l_1}/\image(f_1)\cong R^{l_2}/\image(f_2)$ to the pair of morphisms $((\bar f_1,\bar p_1), (\bar f_2,\bar p_2))$, where $\bar f_i$ is the restriction of $f_i$ to $R^{k_i}$ and $\bar p_i$ is the composition of $p_i$ with the projection of $R^{k_i}\oplus R^{n_i}$ onto $R^{k_i}$. Using this it is easy to see that each connected component has a terminal object, hence ${\mathcal K}_R$ is indeed a well-structured index category.
\end{example}
\section{Model structures on \texorpdfstring{${\mathcal K}$}{K}-spaces}
In this section we introduce the various model structures on diagram spaces associated to well-structured relative index categories. We assume that the reader is familiar with the basic theory of cofibrantly generated model categories as presented in \cite[Chapter 11]{Hirschhorn-model} and
\cite[Section 2.1]{Hovey-model}.
\subsection{Model structures on \texorpdfstring{$G$}{G}-spaces}
\label{subs:G-space-model}
We review some well-known facts about equivariant homotopy theory.
Recall that the category of spaces
${\mathcal S}$ is cofibrantly generated with generating cofibrations $I$ the set of maps
$\partial \Delta^n\to \Delta^n$ for $n\geq 0$ and generating acyclic cofibrations $J$ the set of maps $\Lambda^n_i\to \Delta ^n$ for $n>0$ and $0\leq i\leq n$, see
\cite[Section 11.1]{Hirschhorn-model} for details. Here the notation indicates that the generating (acyclic) cofibrations in the topological setting are obtained from those in the simplicial setting by geometric realization.
Consider a discrete group $G$ and write $\mathcal S^G$ for the category
of left $G$-spaces. It admits several model structures that are of
interest for us. Fix a normal subgroup $A$ in $G$. We say that a map of $G$-spaces $X\to Y$ is an \emph{$A$-relative weak equivalence} (or an
\emph{$A$-relative fibration}) if the induced map of fixed points $X^H\to Y^H$ is a weak equivalence (or fibration) for all subgroups $H\subseteq A$. We say that a map of $G$-spaces is an $A$-relative cofibration if it has the left lifting property with respect to maps that are $A$-relative weak equivalences and $A$-relative fibrations. Let $I_{(G,A)}$ be the set of maps in ${\mathcal S}^G$ of the form $G/H\times i$ for $H\subseteq A$ and $i\in I$, and let $J_{(G,A)}$ be the set of maps of the form $G/H\times j$ for $H\subseteq A$ and $j\in J$.
The following well-known result is an easy consequence of the recognition principle for cofibrantly generated model categories~\cite[Theorem 2.1.19]{Hovey-model}.
\begin{proposition}
Let $A$ be a normal subgroup of $G$.
The $A$-relative weak equivalences, fibrations, and cofibrations specify a cofibrantly generated model structure on ${\mathcal S}^G$ with generating cofibrations $I_{(G,A)}$ and generating acyclic cofibrations $J_{(G,A)}$. \qed
\end{proposition}
We shall refer to this as the \emph{$A$-relative model structure} on ${\mathcal S}^G$.
When $A$ is the trivial group this is also known as the \emph{coarse} (or weak or naive) model structure and when $A=G$ this is sometimes called the \emph{fine} (or strong or genuine) model structure.
As observed by Shipley \cite{Shipley-convenient} (in the case $A=G$) it is possible to combine the weak equivalences in the coarse model structure with the cofibrations in the $A$-relative model structure to get a cofibrantly generated \emph{mixed model structure} on ${\mathcal S}^G$. We recall the details of this construction.
Let $EH$ denote the one-sided bar construction $B(H,H,*)$ of a subgroup
$H$ in $A$, and let $\pi_H\colon G\times_HEH\to G/H$ be the projection. We write $M(\pi_H)$ for the mapping cylinder of $\pi_H$ and consider the
standard factorization
\begin{equation}\label{eq:pi_H-factorization}
G\times_HEH\xrightarrow{j_H}M(\pi_H)\xrightarrow{r_H}G/H
\end{equation}
where $j_H$ is an $A$-relative cofibration and $r_H$ is a $G$-equivariant homotopy equivalence. Let $J'_{(G,A)}$ be the set of morphisms in
$\mathcal S^G$ of the form $j_H\Box i$, where $j_H$ is as above
(for $H\subseteq A$), $i\in I$, and $j_H \Box i$ is the pushout-product of
$j_H$ and $i$ (see the remarks preceding
Proposition~\ref{prop:level-S-model}).
We write $J^{\textit{mix}}_{(G,A)} =J_{(G,A)}\cup J'_{(G,A)}$. The significance of the set $J^{\textit{mix}}_{(G,A)}$ is explained in the next lemma which is implicit in the proof of \cite[Proposition 1.3]{Shipley-convenient}. Recall that given a group $H$ and an $H$-space $X$, the homotopy fixed points $X^{hH}$ is the space of equivariant maps $\Map_H(EH, X)$ (which is the same thing as the homotopy limit of $X$ viewed as a diagram over the one-object category $H$).
\begin{lemma}\label{flat-injective-lemma}
A map $X\to Y$ in $\mathcal S^G$ is $J^{\textit{mix}}_{(G,A)}$-injective if
and only if the induced maps $X^H\to Y^H$ are fibrations and the
diagrams
\[
\xymatrix@-1pc{
X^H\ar[r]\ar[d] & X^{hH}\ar[d]\\
Y^H\ar[r] & Y^{hH}
}
\]
are homotopy cartesian for all subgroups $H$ in $A$.\qed
\end{lemma}
A map of $G$-spaces $X\to Y$ is said to be an \emph{$A$-relative mixed fibration} if it satisfies the equivalent conditions of Lemma
\ref{flat-injective-lemma}. Arguing as in the proof of \cite[Proposition 1.3]{Shipley-convenient} we get the \emph{$A$-relative mixed model structure on ${\mathcal S}^G$}:
\begin{proposition}
The coarse weak equivalences, the $A$-relative mixed fibrations, and the $A$-relative cofibrations specify a cofibrantly generated model structure on
${\mathcal S}^G$ with generating cofibrations $I_{(G,A)}$ and generating acyclic cofibrations $J^{\textit{mix}}_{(G,A)}$. \qed
\end{proposition}
\begin{remark}
For $A=G$ and ${\mathcal S}$ the category of simplicial sets, it is easy to check that the $A$-relative (that is, the \emph{fine}) cofibrations are the maps in
${\mathcal S}^{G}$ whose underlying maps in ${\mathcal S}$ are cofibrations.
\end{remark}
\subsection{The ${\mathcal A}$-relative level model structure on
\texorpdfstring{${\mathcal K}$}{K}-spaces}\label{subs:level-model}
\label{subs:level-model-structures}
Consider now a well-structured relative index category $({\mathcal K},{\mathcal A})$ in the sense of Definition \ref{def:well-structured-relative-index}. Recall that for an object $\bld k$ in ${\mathcal K}$ we write ${\mathcal K}(\bld k)$ and ${\mathcal A}(\bld k)$ for the automorphism groups of $\bld k$ in ${\mathcal K}$ and ${\mathcal A}$, respectively.
We say that a map
$X \to Y$ of ${\mathcal K}$-spaces is an \emph{${\mathcal A}$-relative level equivalence} if \mbox{$X(\bld{k})\to Y(\bld{k})$} is a weak equivalence of spaces for every object $\bld{k}$ in ${\mathcal A}$. (This will not lead to confusion with the notion of an $A$-relative weak equivalence of $G$-spaces introduced in Section
\ref{subs:G-space-model}.)
A map of ${\mathcal K}$-spaces $X \to Y$ is an \emph{${\mathcal A}$-relative level fibration}
if for all objects $\bld{k}$ in ${\mathcal A}$ and all subgroups
$H$ in ${\mathcal A}(\bld{k})$, the map $X(\bld{k})^H\to Y(\bld{k})^H$
is a fibration and the diagram
\begin{equation}\label{eq:flat-level-fibration-square}
\xymatrix@-1pc{
X(\bld{k})^H\ar[r] \ar[d] & X(\bld{k})^{hH}\ar[d]\\
Y(\bld{k})^H\ar[r] & Y(\bld{k})^{hH} }
\end{equation}
is homotopy cartesian. Finally, a map of ${\mathcal K}$-spaces is an ${\mathcal A}$-relative cofibration if it has the left lifting property with respect to maps that are
${\mathcal A}$-relative level equivalences and ${\mathcal A}$-relative level fibrations.
Recall the functors $G_{\bld{k}}^{{\mathcal K}}$ from Section~\ref{subs:free-semi-free} and let
\[ \begin{split} I^{\textit{level}}_{({\mathcal K},{\mathcal A})} &= \{
G^{{\mathcal K}}_{\bld{k}}(i) | \bld{k}\in O({\mathcal A}) \textrm{ and } i\in
I_{({\mathcal K}(\bld k),{\mathcal A}(\bld k))}\}\quad \textrm{and}\\
J^{\textit{level}}_{({\mathcal K},{\mathcal A})} &= \{ G^{{\mathcal K}}_{\bld{k}}(j) |
\bld{k}\in O({\mathcal A})\textrm{ and }
j\in J^{\textit{mix}}_{({\mathcal K}(\bld k),{\mathcal A}(\bld k))} \}. \end{split}\]
Here we write $O({\mathcal A})$ for the set of objects in ${\mathcal A}$.
\begin{proposition}\label{prop:level-A-relative}
The ${\mathcal A}$-relative level equivalences, level fibrations, and cofibrations specify a cofibrantly generated model structure on ${\mathcal S}^{{\mathcal K}}$ with generating cofibrations $I^{\textit{level}}_{({\mathcal K},{\mathcal A})}$ and generating acyclic cofibrations
$J^{\textit{level}}_{({\mathcal K},{\mathcal A})}$.
\end{proposition}
We shall refer to this as the \emph{${\mathcal A}$-relative level model structure} on
$\mathcal S^{{\mathcal K}}$.
\begin{proof}
We use the recognition criterion for cofibrantly generated model
categories as stated in \cite[Theorem 2.1.19]{Hovey-model}. The
smallness requirements are satisfied because the generating (acyclic)
cofibrations are levelwise cofibrations in ${\mathcal S}$ and ${\mathcal S}$ is small relative to
the cofibrations.
It follows from the definition that a map $X\to Y$ in ${\mathcal S}^{{\mathcal K}}$ is
$ I^{\textit{level}}_{({\mathcal K},{\mathcal A})}$-injective if and only if for all
$\bld{k}$ in ${\mathcal A}$ and all subgroups
$H\subseteq {\mathcal A}(\bld{k})$ the induced map $X(\bld{k})^H\to
Y(\bld{k})^H$ is an acyclic fibration. Similarly, Lemma
\ref{flat-injective-lemma} implies that $X\to Y$ is
$J^{\textit{level}}_{({\mathcal K},{\mathcal A})}$-injective if and only if it is an
${\mathcal A}$-relative level fibration. From these explicit descriptions it
is clear that the $ I^{\textit{level}}_{({\mathcal K},{\mathcal A})}$-injective maps are
the $J^{\textit{level}}_{({\mathcal K},{\mathcal A})}$-injective maps that are
${\mathcal A}$-relative level equivalences. Furthermore, this has as a formal
consequence that the class $J^{\textit{level}}_{({\mathcal K},{\mathcal A})}\cof$ is
contained in $I^{\textit{level}}_{({\mathcal K},{\mathcal A})}\cof$.
It remains to show that the maps in
$J^{\textit{level}}_{({\mathcal K},{\mathcal A})}\cell$ are ${\mathcal A}$-relative level
equivalences. We first show that the maps in
$J^{\textit{level}}_{({\mathcal K},{\mathcal A})}$ are level equivalences and level cofibrations at all levels, not only those corresponding to objects in ${\mathcal A}$. Indeed, for a map of the form $G^{{\mathcal K}}_{\bld{k}}(j)$ for $\bld k$ in ${\mathcal A}$ and
$j\in J_{({\mathcal K}(\bld k),{\mathcal A}(\bld k))}$ this easily follows from the explicit description of the functor $G^{{\mathcal K}}_{\bld{k}}$ in \eqref{eq:free-functors}.
Consider then a map of the form $G^{{\mathcal K}}_{\bld{k}}(j_H\Box i)$ for $\bld k$ in
${\mathcal A}$ and $j_H\Box i$ in $J'_{({\mathcal K}(\bld k),{\mathcal A}(\bld k))}$. Since
$G_{\bld k}^{{\mathcal K}}$ preserves colimits and tensors (over ${\mathcal S}$) we can identify this map with the pushout-product $G_{\bld k}^{{\mathcal K}}(j_H)\Box i$.
Using that $H$ acts freely from the right on the morphism sets ${\mathcal K}(\bld k, -)$ (by Lemma~\ref{lem:free-right-action}, letting $X=F_{\bld 0}^{{\mathcal K}}(*)$) we again conclude from \eqref{eq:free-functors} that
$G^{{\mathcal K}}_{\bld{k}}(\pi_H)$ is a level equivalence. Identifying
$G^{{\mathcal K}}_{\bld{k}}(M(\pi_H))$ with the mapping cylinder of
$G_{\bld k}^{{\mathcal K}}(\pi_H)$ we see that $G^{{\mathcal K}}_{\bld{k}}(j_H)$ is both a level equivalence and a level cofibration. By the pushout-product axiom for
${\mathcal S}$ we finally conclude that $G_{\bld k}^{{\mathcal K}}(j_H)\Box i$ is a level equivalence and a level cofibration. By definition, a map in
$J^{\textit{level}}_{({\mathcal K},{\mathcal A})}\cell$ is the transfinite composition
of a sequence of maps each of which is the pushout of a map in
$J^{\textit{level}}_{({\mathcal K},{\mathcal A})}$. At each level such a map is therefore the
transfinite composition of a sequence of acyclic cofibrations, hence a weak
equivalence.
\end{proof}
As promised, there is a more explicit description of the ${\mathcal A}$-relative cofibrations. Let $\partial\left({\mathcal K} \!\downarrow \!\bld{k}\right)$ be the full
subcategory of $({\mathcal K} \!\downarrow \!\bld{k})$ whose objects are the
non-isomor\-phisms. For a ${\mathcal K}$-space $X$, the $\bld{k}$-th latching
space $L_{\bld k}(X)$ is the colimit of the
$\partial\left({\mathcal K} \!\downarrow \!\bld{k}\right)$-diagram
$(\bld l\to\bld k)\mapsto X(\bld l)$, and the $\bld k$th latching map is the
canonical ${\mathcal K}(\bld k)$-equivariant map
\[ L_{\bld{k}}(X) = \operatornamewithlimits{colim}_{(\bld{l}\to\bld{k}) \in \partial\left({\mathcal K}
\downarrow \bld{k}\right)} X(\bld{l}) \to
\operatornamewithlimits{colim}_{(\bld{l}\to\bld{k}) \in ({\mathcal K} \downarrow \bld{k})}X(\bld l)
\xrightarrow{\cong} X(\bld{k}).\]
For a map of ${\mathcal K}$-spaces $X \to Y$, the $\bld{k}$-th latching map
is the ${\mathcal K}(\bld k)$-equivariant map
\begin{equation}\label{eq:latching-map}
L_{\bld{k}}(Y) \cup_{L_{\bld{k}}(X)}X(\bld{k}) \to Y(\bld{k}).
\end{equation}
Recall the notion of an ${\mathcal A}(\bld k)$-relative cofibration in ${\mathcal S}^{{\mathcal K}(\bld k)}$ from Section \ref{subs:G-space-model}.
\begin{proposition}\label{prop:latching-characterization}
A map of ${\mathcal K}$-spaces $f\colon X\to Y$ is an ${\mathcal A}$-relative
cofibration if and only if the latching map \eqref{eq:latching-map}
is an ${\mathcal A}(\bld k)$-relative cofibration for all $\bld k$ in ${\mathcal A}$ and an isomorphism for all $\bld k$ not in ${\mathcal A}$.
\end{proposition}
\begin{proof}
Choosing representatives for the isomorphism classes of objects $\bld{k}$
in ${\mathcal K}$ with common value $\lambda(\bld{k})=n$, one shows by induction
on $n$ that maps satisfying the stated condition have the left lifting property
with respect to the maps that are ${\mathcal A}$-relative level equivalences and
${\mathcal A}$-relative level fibrations. Hence maps satisfying the condition are
${\mathcal A}$-relative cofibrations.
For the other direction, one first shows that the generating cofibrations
satisfy the condition in the proposition. This uses the normality condition on ${\mathcal A}$. Since the condition is preserved under
cobase change, transfinite composition, and retracts, this implies that it holds for all ${\mathcal A}$-relative cofibrations.
\end{proof}
\begin{remark}\label{rem:simpl-top-K-model} One may also compare the
simplicial and the topological version of these model structures:
The geometric realization functor $|- |$ and the singular complex functor
$\Sing$ induce an adjunction between ${\mathcal K}$-diagrams in simplicial sets and (compactly generated weak Hausdorff) topological spaces. It is easy to
check that this defines a Quillen equivalence with respect to each of the
model structures we consider. Using the $(|-|,\Sing)$-adjunction we can also turn any ${\mathcal S}$-model structure in the topological setting into a simplicial model structure.
\end{remark}
Recall from \cite[Definition 4.2.18]{Hovey-model} that an ${\mathcal S}$-model category ${\mathcal M}$ is a category which is enriched, tensored, and cotensored over
${\mathcal S}$, and equipped with a model structure such that if $f\colon X\to Y$ is a cofibration in ${\mathcal M}$ and $g\colon S\to T$ a cofibration in ${\mathcal S}$, then the pushout-product
\[
f\Box g\colon Y\times S\cup_{X\times S}X\times T\to Y\times T
\]
is a cofibration in ${\mathcal M}$ which is acyclic if either $f$ or $g$ is.
\begin{proposition}\label{prop:level-S-model}
The ${\mathcal A}$-relative level model structure on
${\mathcal S}^{{\mathcal K}}$ is an ${\mathcal S}$-model structure.
\end{proposition}
\begin{proof}
We already know from Lemma \ref{lem:K-spaces-S-category} that ${\mathcal S}^{{\mathcal K}}$ is enriched, tensored, and cotensored over ${\mathcal S}$. For the statements concerning the pushout-product, it suffices by
\cite[Corollary 4.2.5]{Hovey-model} to consider the generating (acyclic) cofibrations for the respective model structures. Let $\bld k$ be an object in ${\mathcal A}$, let $f$ be a map in ${\mathcal S}^{{\mathcal K}(\bld k)}$, and let $g$ be a map in ${\mathcal S}$. Then $G_{\bld k}^{{\mathcal K}}(f)\Box g$ can be identified with
$G_{\bld k}^{{\mathcal K}}(f\Box g)$ where $f\Box g$ is the pushout-product in
${\mathcal S}^{{\mathcal K}(\bld k)}$. If $f$ is a generating ${\mathcal A}(\bld k)$-relative cofibration and $g$ is a generating cofibration in ${\mathcal S}$, then $f\Box g$ is an ${\mathcal A}(\bld k)$-relative cofibration which in turn implies that
$G_{\bld k}^{{\mathcal K}}(f\Box g)$ is an ${\mathcal A}$-relative cofibration. Now suppose that either $f$ or $g$ is a generating acyclic cofibration. Applying the pushout-product axiom for ${\mathcal S}$ levelwise we see that $G_{\bld k}^{{\mathcal K}}(f)\Box g$ is then an ${\mathcal A}$-relative level equivalence.
\end{proof}
\subsection{The ${\mathcal A}$-relative \texorpdfstring{${\mathcal K}$}{K}-model structures}
\label{sec:K-model-structure}
Let ${\mathcal C}$ be a small category and let $X$ be a ${\mathcal C}$-diagram
of spaces. We either write $X_{h{\mathcal C}}$ or $\hocolim_{{\mathcal C}}X$ for the homotopy colimit of $X$ over ${\mathcal C}$, defined in the usual way as the realization of the simplicial replacement of the diagram,
\[
X_{h{\mathcal C}}=\hocolim_{{\mathcal C}}X = \bigg| [s] \mapsto \displaystyle\coprod_{\bld{k_0} \leftarrow \dots
\leftarrow \bld{k_s}} X(\bld{k_s})\bigg|,
\]
see e.g.\ \cite{Bousfield_K-homotopy,Hirschhorn-model} for details.
(In the simplicial setting $|\ |$ indicates the diagonal simplicial set.) For homotopy colimits in a general model category, one has to
assume that the diagram $X$ is object-wise cofibrant for this construction
to capture the correct homotopy type. We do not need this assumption here
because we either work in simplicial sets (where it is automatically satisfied) or in (compactly generated weak Hausdorff) topological spaces, where it follows from \cite[Appendix A]{Dugger_I-hypercovers} that the assumption can be dropped.
We will freely use many standard properties of homotopy colimits as
for example developed in~\cite[Chapter 18]{Hirschhorn-model}.
Moreover, we will use the following result which is reproduced here for easy reference.
\begin{lemma}\label{lem:puppe-lemma}\cite[Proposition
4.4]{Rezk_SS-simplicial} Let ${\mathcal C}$ be a small category, let $X\to Y$ be
a map of ${\mathcal C}$-diagrams in ${\mathcal S}$, and let $\alpha \colon \bld{k}\to\bld{l}$ be
a morphism in ${\mathcal C}$. Consider the two squares
\[\xymatrix@-1pc{X(\bld{k}) \ar[r] \ar[d]_{X(\alpha)}&
Y(\bld{k})\ar[d]^{Y(\alpha)}\\
X(\bld{l})\ar[r] & Y(\bld{l})}
\qquad\qquad\xymatrix@-1pc{X(\bld{k}) \ar[r] \ar[d]& Y(\bld{k})\ar[d]\\
X_{h{\mathcal C}} \ar[r] &Y_{h{\mathcal C}}.}\] If the left hand square is homotopy
cartesian for every $\alpha$, then the right hand square is homotopy
cartesian for every object $\bld{k}$.\qed
\end{lemma}
\begin{remark}\label{rem:puppe-remark}
In \cite{Rezk_SS-simplicial} the above lemma is only stated for simplicial sets, but the analogous result for (compactly generated weak Hausdorff) topological spaces is an immediate consequence. Indeed, recall that a square diagram of topological spaces is homotopy cartesian if and only if applying the singular complex functor $\Sing$ gives a homotopy cartesian diagram of simplicial sets. Conversely, a square diagram of simplicial sets is homotopy cartesian if and only if the geometric realization is homotopy cartesian. Thus, given a map $X\to Y$ of ${\mathcal C}$-diagrams of topological spaces such that the left hand squares are homotopy cartesian, the lemma implies that the diagram
\[
\xymatrix@-1pc{
\Sing X(\bld k) \ar[r] \ar[d]& \Sing Y(\bld k)\ar[d]\\
(\Sing X)_{h{\mathcal C}} \ar[r] & (\Sing Y)_{hC}
}
\]
is homotopy cartesian. This in turn implies that the geometric realization is homotopy cartesian and the natural transformation $|\Sing X|\to X$ defines a natural weak equivalence between this realization and the right hand square in the lemma.
\end{remark}
The following definition is central to the rest of the paper.
\begin{definition}
A map $X \to Y$ of ${\mathcal K}$-spaces is a \emph{${\mathcal K}$-equivalence} if the induced map $X_{h{\mathcal K}} \to Y_{h{\mathcal K}}$ is a weak equivalence of spaces.
\end{definition}
Now let $({\mathcal K},{\mathcal A})$ be a well-structured relative index category. We proceed to construct an ${\mathcal A}$-relative model structure on ${\mathcal S}^{{\mathcal K}}$ with the
${\mathcal K}$-equivalences as the weak equivalences. The cofibrations for this model structure are the ${\mathcal A}$-relative cofibrations as characterized in Proposition~\ref{prop:latching-characterization}. Recall that ${\mathcal K}_{{\mathcal A}}$ denotes the full subcategory of ${\mathcal K}$ generated by the objects of ${\mathcal A}$.
A map of ${\mathcal K}$-spaces $X\to Y$ is an \emph{${\mathcal A}$-relative
${\mathcal K}$-fibration} if it is an ${\mathcal A}$-relative level
fibration with the additional property that every morphism
$\alpha \colon \bld{k}\to\bld{l}$ in ${\mathcal K}_{{\mathcal A}}$
induces a homotopy cartesian square
\[
\xymatrix@-1pc{
X(\bld{k})\ar[r] \ar[d] & X(\bld{l})\ar[d]\\
Y(\bld{k})\ar[r] & Y(\bld{l}). }
\]
Let $\alpha\colon \bld k\to\bld l$ be a morphism in ${\mathcal K}_{{\mathcal A}}$ and consider the induced map of ${\mathcal K}$-spaces $\alpha^*\colon F_{\bld{l}}^{{\mathcal K}}(*)\to
F_{\bld{k}}^{{\mathcal K}}(*)$ defined by precomposition with $\alpha$.
\begin{lemma}\label{lem:alpha^*-equivalence}
The map $\alpha^*\colon F_{\bld{l}}^{{\mathcal K}}(*)\to F_{\bld{k}}^{{\mathcal K}}(*)$ is a
${\mathcal K}$-equivalence.
\end{lemma}
\begin{proof}
By definition of the homotopy colimits, $F_{\bld{k}}^{{\mathcal K}}(*)_{h{\mathcal K}}$ and $F_{\bld{l}}^{{\mathcal K}}(*)_{h{\mathcal K}}$ can be identified with the classifying spaces of the categories $(\bld k\downarrow {\mathcal K})$ and $(\bld l\downarrow {\mathcal K})$. These categories each has an initial object and the map induced by $\alpha^*$ is therefore trivially a weak equivalence.
\end{proof}
We now use the tensor with the interval
in ${\mathcal S}$ to factor $\alpha^*$ through the mapping cylinder $M(\alpha^*)$ in the usual way,
\begin{equation}\label{eq:alpha-star-fact}
\alpha^*\colon F_{\bld{l}}^{{\mathcal K}}(*)\xrightarrow{j_{\alpha}}M(\alpha^*)\xrightarrow{r_{\alpha}} F_{\bld{k}}^{{\mathcal K}}(*).
\end{equation}
Arguing as in the case of symmetric spectra \cite[Lemma 3.4.10]{HSS}
we see that $j_{\alpha}$ is an ${\mathcal A}$-relative cofibration and $r_{\alpha}$ is a homotopy equivalence. Let $J_{({\mathcal K},{\mathcal A})}'$ be the set of morphisms of the form
$j_{\alpha}\Box i$ where $j_{\alpha}$ is as above (for $\alpha$ in ${\mathcal K}_{{\mathcal A}}$), $i$ is a generating cofibration in ${\mathcal S}$, and
$\Box$ denotes the pushout-product map associated to the tensor with an object of ${\mathcal S}$.
We define $I_{({\mathcal K},{\mathcal A})}=I^{\textit{level}}_{({\mathcal K},{\mathcal A})}$
and $J_{({\mathcal K},{\mathcal A})}=J^{\textit{level}}_{({\mathcal K},{\mathcal A})}\cup J'_{({\mathcal K},{\mathcal A})}$.
\begin{proposition}\label{prop:K-model-str}
The ${\mathcal K}$-equivalences together with the ${\mathcal A}$-relative ${\mathcal K}$-fibrations and
the ${\mathcal A}$-relative cofibrations specify a cofibrantly
generated model structure on ${\mathcal S}^{{\mathcal K}}$ with generating
cofibrations $I_{({\mathcal K},{\mathcal A})}$ and generating acyclic
cofibrations $J_{({\mathcal K},{\mathcal A})}$.
\end{proposition}
We shall refer to this as the \emph{${\mathcal A}$-relative ${\mathcal K}$-model structure} on
${\mathcal K}$-spaces.
\begin{proof}
We again use the criterion of~\cite[Theorem 2.1.19]{Hovey-model}.
One can apply Proposition \ref{prop:h-cofibration-properties} (viii) below to
see that the smallness requirements are satisfied.
As in the ${\mathcal A}$-relative level model structure, the
$I_{({\mathcal K},{\mathcal A})}$-injective maps are the maps $X\to Y$
such that for all objects $\bld{k}$ in ${\mathcal A}$ and all
subgroups $H\subseteq {\mathcal A}(\bld{k})$, the induced map
$X(\bld{k})^H\to Y(\bld{k})^H$ is an acyclic fibration. Moreover,
arguing as in the proof of \cite[Lemma 3.4.12]{HSS}, we see
that $X\to Y$ is $J_{({\mathcal K},{\mathcal A})}$-injective if and
only if it is an ${\mathcal A}$-relative ${\mathcal K}$-fibration.
Thus, a map $X \to Y$ which is $I_{({\mathcal K},{\mathcal A})}$-injective is clearly both
$J_{({\mathcal K},{\mathcal A})}$-injective and a ${\mathcal K}$-equivalence.
Suppose then that $f\colon X\to Y$ is $J_{({\mathcal K},{\mathcal A})}$-injective
and a ${\mathcal K}$-equivalence. Then $f$ is an ${\mathcal A}$-relative level
fibration and it follows from the homotopy cofinality condition (iv) for a well-structured relative index category that the induced map
$X_{h{\mathcal K}_{{\mathcal A}}}\to Y_{h{\mathcal K}_{{\mathcal A}}}$ is a weak equivalence. Therefore
Lemma \ref{lem:puppe-lemma} implies that $X\to Y$ is also an
${\mathcal A}$-relative level equivalence, hence $I_{({\mathcal K},{\mathcal A})}$-injective.
The last thing to be checked is that the maps in the class
$J_{({\mathcal K},{\mathcal A})}\cell$ also belong to the class
$I_{({\mathcal K},{\mathcal A})}\cof$ and are ${\mathcal K}$-equivalences. Here
the first part follows formally from the above discussion. For the
second part we first observe that the maps in $J_{({\mathcal K},{\mathcal A})}$ are
${\mathcal K}$-equivalences by Lemma \ref{lem:alpha^*-equivalence}.
We next observe that the functor $\hocolim_{{\mathcal K}}$ takes the
class $I_{({\mathcal K},{\mathcal A})}\cof$ to cofibrations in ${\mathcal S}$. Indeed, since
$\hocolim_{{\mathcal K}}$ preserves
colimits it suffices to check that it takes the elements in
$I_{({\mathcal K},{\mathcal A})}$ to cofibrations in ${\mathcal S}$ and
this is easy to check directly. By definition, a map in
$J_{({\mathcal K},{\mathcal A})}\cell$ is the transfinite composition
of a sequence of maps each of which is a pushout of a map in
$J_{({\mathcal K},{\mathcal A})}$. The induced map
$X_{h{\mathcal K}}\to Y_{h{\mathcal K}}$ is therefore the
transfinite composition of a sequence of
maps each of which is a pushout of an acyclic cofibration; again
because $\hocolim_{{\mathcal K}}$ preserves colimits. The induced map itself
is therefore also a weak equivalence as had to be shown.
\end{proof}
\begin{remark}
Dugger studied hocolim model structures on ${\mathcal C}$-diagrams in a model
category ${\mathcal M}$ for a contractible category ${\mathcal C}$ in~\cite[Theorem
5.2]{Dugger_replacing-simplicial}. These coincide with the model
structures of the previous proposition if ${\mathcal K}$ is contractible and
${\mathcal A} = O{\mathcal K}$.
\end{remark}
For future reference we spell out the condition for a ${\mathcal K}$-space to be fibrant in the model structure of Proposition~\ref{prop:K-model-str}.
\begin{proposition}\label{prop:A-relative-fibrant-K-space}
A ${\mathcal K}$-space $X$ is fibrant in the ${\mathcal A}$-relative ${\mathcal K}$-model
structure if and only if
\begin{enumerate}[(i)]
\item for each $\bld{k}$ in ${\mathcal A}$ and
each subgroup $H\subseteq {\mathcal A}(\bld{k})$ the space
$X(\bld{k})^{H}$ is fibrant and the map $X(\bld{k})^{H}\to
X(\bld{k})^{hH}$ is a weak equivalence, and
\item for each morphism
$\alpha\colon\bld{k}\to\bld{l}$ in ${\mathcal K}_{{\mathcal A}}$ the induced map
$X(\bld{k})\to X(\bld{l})$ is a weak equivalence.\qed
\end{enumerate}
\end{proposition}
Here the fibrancy condition on the spaces $X(\bld k)^H$ is of course automatically satisfied in the topological setting.
\begin{proposition}\label{prop:K-model-is-S-model-str}
The ${\mathcal A}$-relative ${\mathcal K}$-model structure on ${\mathcal S}^{{\mathcal K}}$ is an ${\mathcal S}$-model structure.
\end{proposition}
\begin{proof}
Let $f\colon X\to Y$ be an ${\mathcal A}$-relative cofibration in ${\mathcal S}^{{\mathcal K}}$ and let
$g\colon S\to T$ be a cofibration in ${\mathcal S}$. Then the pushout-product
$f\Box g$ is an ${\mathcal A}$-relative cofibration by
Proposition~\ref{prop:level-S-model} and we must show it to be a
${\mathcal K}$-equivalence if either $f$ is a ${\mathcal K}$-equivalence or $g$ is a weak equivalence. Using that the homotopy colimit functor preserves colimits and tensors we can identify $(f\Box g)_{h{\mathcal K}}$ with the pushout-product $f_{h{\mathcal K}}\Box g$ in ${\mathcal S}$. Since the homotopy colimit functor also preserves cofibrations (see the proof of Proposition
\ref{prop:K-model-str}) the result now follows from the pushout-product axiom for ${\mathcal S}$.
\end{proof}
As discussed in Remark~\ref{rem:simpl-top-K-model}, we can use the
$(|-|,\Sing)$-adjunction to make the ${\mathcal A}$-relative ${\mathcal K}$-model structure a simplicial model structure also in the topological version of the theory.
In the next proposition we compare the relative ${\mathcal K}$-model structures associated to different subcategories of automorphisms. Recall the normality and multiplicative conditions on our subcategories of automorphisms stated before Definition \ref{def:well-structured-relative-index}.
\begin{proposition}\label{prop:well-structured-comparison}
Let $({\mathcal K},{\mathcal B})$ be a well-structured relative index category. Suppose that
${\mathcal A}$ is a normal and multiplicative subcategory of automorphisms contained in ${\mathcal B}$ and that the inclusion ${\mathcal K}_{{\mathcal A}}\to {\mathcal K}_{{\mathcal B}}$ is homotopy cofinal. Then $({\mathcal K},{\mathcal A})$ is a well-structured relative index category and the
${\mathcal A}$- and ${\mathcal B}$-relative ${\mathcal K}$-model structures on ${\mathcal S}^{{\mathcal K}}$ are Quillen equivalent.
\end{proposition}
\begin{proof}
It is immediate from the definitions that the identity functor on ${\mathcal S}^{{\mathcal K}}$ is the left Quillen functor of a Quillen equivalence from the ${\mathcal A}$-relative
${\mathcal K}$-model structure to the ${\mathcal B}$-relative ${\mathcal K}$-model structure.
\end{proof}
Now let us specialize to the case of a well-structured index category ${\mathcal K}$ as specified in Definition \ref{def:well-structured-index}.
\begin{definition}\label{def:K-projective-model}
Let ${\mathcal K}$ be a well-structured index category.
\begin{enumerate}[(i)]
\item
The \emph{projective ${\mathcal K}$-model structure} on ${\mathcal S}^{{\mathcal K}}$ is obtained from Proposition \ref{prop:K-model-str} by letting ${\mathcal A}$ be the category of identity morphisms $O{\mathcal K}$ in ${\mathcal K}$.
\item
Suppose that the full subcategory ${\mathcal K}_+$ of objects with positive degree is homotopy cofinal in ${\mathcal K}$. The \emph{positive projective ${\mathcal K}$-model structure} on ${\mathcal S}^{{\mathcal K}}$ is then obtained from Proposition \ref{prop:K-model-str} by letting ${\mathcal A}$ be the category of identity morphisms $O{\mathcal K}_+$ in ${\mathcal K}_+$.
\end{enumerate}
\end{definition}
We next describe some features of the projective
${\mathcal K}$-model structure that is not shared by the ${\mathcal A}$-relative ${\mathcal K}$-model structures in general.
\begin{lemma}\label{lem:projective-hocolim}
Let $X$ be a ${\mathcal K}$-space which is cofibrant in the projective ${\mathcal K}$-model structure on ${\mathcal S}^{{\mathcal K}}$. Then the canonical map $\hocolim_{{\mathcal K}}X\to
\colim_{{\mathcal K}}X$ is a weak equivalence.
\end{lemma}
\begin{proof}
This is proved in the simplicial setting in \cite[Proposition
18.9.4]{Hirschhorn-model} and a similar argument applies for
${\mathcal K}$-diagrams of topological spaces.
\end{proof}
\begin{proposition}\label{prop:colim-const-Q-adjunction}
Let ${\mathcal K}$ be a well-structured index category and give ${\mathcal S}^{{\mathcal K}}$ the projective ${\mathcal K}$-model structure. Then the adjunction
$\colim_{{\mathcal K}} \colon {\mathcal S}^{{\mathcal K}} \rightleftarrows {\mathcal S}
\thinspace \colon\! \!\const_{{\mathcal K}}$ is a Quillen adjunction. It is a Quillen equivalence if and only if $B{\mathcal K}$ is contractible.
\end{proposition}
\begin{proof}
The adjunction is a Quillen adjunction because $\const_{{\mathcal K}}$ preserves fibrations and acyclic fibrations. Suppose that $B{\mathcal K}$ is contractible. To show that the adjunction is a Quillen equivalence we must check that for a cofibrant ${\mathcal K}$-space $X$ and a fibrant space $Y$, a map $\colim_{{\mathcal K}}X\to Y$ is a weak equivalence if and only if its adjoint $X\to \const_{{\mathcal K}}Y$ is a ${\mathcal K}$-equivalence. These maps fit into a commutative diagram
\[
\xymatrix@-1pc{
X_{h{\mathcal K}} \ar[r]\ar[d] & (\const_{{\mathcal K}}Y)_{h{\mathcal K}} \ar[r]^-{\sim} & B{\mathcal K}\times Y\ar[d]\\
\colim_{{\mathcal K}}X \ar[rr] && Y
}
\]
where the vertical maps are induced by the canonical map from the homotopy colimit to
the colimit. This gives the result since the vertical maps are weak equivalences by Lemma~\ref{lem:projective-hocolim} and the assumption on $B{\mathcal K}$. Next assume that the adjunction is a Quillen equivalence. Letting
$X=F^{{\mathcal K}}_{\bld 0}(*)$ and $Y=*$ in the above diagram then shows that $B{\mathcal K}$ is contractible.
\end{proof}
\section{The class of \texorpdfstring{$h$}{h}-cofibrations} \label{subs:h-cofibrations}
In this section $({\mathcal K},{\mathcal A})$ again denotes a well-structured relative index category. For the homotopical analysis of the monoidal structure of
$\mathcal S^{{\mathcal K}}$ it will be convenient to have available a weaker
notion of cofibrations than the ${\mathcal A}$-relative cofibrations. Specifically, in the topological setting these will be the classical Hurewicz cofibrations, while in the simplicial setting these will be the levelwise injections.
We shall use the term \emph{$h$-cofibration} for a map in one of these classes.
Below we state a number of results about the $h$-cofibrations which we
verify in the topological and simplicial settings in Sections
\ref{topological-h-cofibration} and \ref{simplicial-h-cofibration},
respectively. Given a map $f \colon X \to Y$ of
${\mathcal K}$-spaces, we write $f^{\Box n}\colon Q^n_{n-1}(f) \to Y^{\boxtimes
n}$ for the $n$-fold iterated pushout product map of $f$. We will
study this construction in more detail in
Section~\ref{iterated-pushout-section} where we
show that $f^{\Box n}$ is a $\Sigma_n$-equivariant map of ${\mathcal K}$-spaces
with $\Sigma_n$-action.
\begin{proposition}\label{prop:h-cofibration-properties}
\begin{enumerate}[(i)]
\item Every ${\mathcal A}$-relative cofibration is an $h$-cofibration.
\item The $h$-cofibrations are preserved under cobase change,
transfinite composition, and retracts.
\item The gluing lemma for $h$-cofibrations and ${\mathcal A}$-relative level
equivalences holds.
\item The gluing lemma for $h$-cofibrations and
${\mathcal K}$-equivalences holds.
\item Let $\lambda$ be an ordinal. If $\{X_{\alpha}\colon
\alpha<\lambda\}$ is a $\lambda$-sequence of $h$-cofibrations, then
the canonical map $\hocolim_{\alpha<\lambda}X_{\alpha}\to
\colim_{\alpha<\lambda}X_{\alpha}$ is a level equivalence.
\item For every ${\mathcal K}$-space $X$, the functor $X\boxtimes(-)$ sends
${\mathcal A}$-relative cofibrations to $h$-cofibrations.
\item Let $f \colon X \to Y$ be a generating cofibration for the ${\mathcal A}$-relative
${\mathcal K}$-model structure and let $Z$ be a ${\mathcal K}$-space with
a right $\Sigma_n$-action. Then $ Z \boxtimes_{\Sigma_n} f^{\Box n}$ is an
$h$-cofibration.
\item Every ${\mathcal K}$-space is small relative to the
$h$-cofibrations.
\end{enumerate}
\end{proposition}
The gluing lemmas in (iii) and (iv) are the statements that given a map of diagrams
\begin{equation}\label{h-cofibrant-diagram}
\xymatrix@-1pc{
Y\ar[d] & X\ar[l] \ar[r]^{i} \ar[d]& Z\ar[d]\\
Y' & X'\ar[l] \ar[r]^{\imath'} & Z'
}
\end{equation}
in which $i$ and $\imath'$ are $h$-cofibrations and the vertical maps are
${\mathcal A}$-relative level equivalences (respectively ${\mathcal K}$-equivalences),
then the map of pushouts $Y\cup_{X}Z\to Y'\cup_{X'}Z'$ is also
an ${\mathcal A}$-relative level equivalence (respectively a ${\mathcal K}$-equivalence).
In (viii) the term \emph{small} has its usual set theoretical meaning, see e.g.,
\cite[Section 10.4]{Hirschhorn-model}.
\begin{remark}
Hill, Hopkins and Ravenel~\cite[Definition
B.15]{Hill-H-R_Kervaire-invariant} define a map $f$ in a model
category to be \emph{flat} if cobase change along $f$ preserves weak
equivalences. The previous proposition shows that the
$h$-cofibrations in ${\mathcal S}^{{\mathcal K}}$ are flat in this sense. In
particular, the flat cofibrations introduced for ${\mathcal I}$- and
${\mathcal J}$-spaces in Sections~\ref{sec:I-space-section}
and~\ref{sec:J-space-section} also satisfy this more general
flatness condition.
\end{remark}
We note some immediate consequences of these results which we state explicitly for easy reference.
\begin{corollary} \label{cor:pushout-K-or-level-h} The cobase change
of a map which is both an ${\mathcal A}$-relative level equivalence (or ${\mathcal K}$-equivalence) and an $h$-cofibration is also an ${\mathcal A}$-relative level equivalence (or ${\mathcal K}$-equivalence) and an $h$-cofibration.\qed
\end{corollary}
\begin{corollary} \label{cor:transfinite-K-or-level-equivalence} If
$\{X_{\alpha}\colon \alpha<\lambda\}$ is a $\lambda$-sequence in
$\mathcal S^{{\mathcal K}}$ such that each of the maps $X_{\alpha}\to
X_{\alpha+1}$ is an ${\mathcal A}$-relative level equivalence (or ${\mathcal K}$-equivalence)
and an $h$-cofibration, then the transfinite composition $X_0\to \colim_{\alpha<\lambda}X_{\alpha}$ is also an ${\mathcal A}$-relative level equivalence
(or ${\mathcal K}$-equivalence) and an $h$-cofibration.\qed
\end{corollary}
\subsection{Topological \texorpdfstring{$h$}{h}-cofibrations}\label{topological-h-cofibration}
In this paragraph we write $\mathrm{Top}= {\mathcal S}$ for our category of spaces to
emphasize that we work in the context of (compactly generated weak
Hausdorff) topological spaces. For the proof of Proposition
\ref{prop:h-cofibration-properties} it is convenient to have
available a Str\o m type model structure on $\mathrm{Top}^{{\mathcal K}}$.
We review the relevant definitions and results. Consider in general a
small category ${\mathcal C}$ and write $\mathrm{Top}^{{\mathcal C}}$ for the category of
${\mathcal C}$-spaces. This category is tensored and cotensored over $\mathrm{Top}$ with
the tensor $X\times K$ and cotensor $X^K$ of a ${\mathcal C}$-space $X$ and a
space $K$ defined by the obvious object-wise constructions. Homotopies
are defined in the usual way using the tensor with the unit interval
$I$.
A morphism $f\colon X\to Y$ in $\mathrm{Top}^{{\mathcal C}}$ is a \emph{Hurewicz
cofibration} (\emph{$h$-cofibration}) if it has the left
lifting property with respect to the map $p_0\colon Z^I\to Z$ (induced
by the inclusion $\{0\}\to I$) for any $\mathcal C$-space $Z$. Writing $M(f)$ for
the mapping cylinder $Y\cup_{X\times\{0\}}X\times I$, this is
equivalent to the condition that the canonical map $M(f)\to Y\times I$
admits a retraction. Similarly, we say that a map in $\mathrm{Top}^{{\mathcal C}}$ is a
\emph{Hurewicz fibration} (\emph{$h$-fibration}) if it has the
right lifting property with respect to the map $i_0\colon A\to A\times
I$ (again induced by $\{0\}\to I$) for any space $A$. When ${\mathcal C}$ is
the terminal category these notions reduce to the usual Hurewicz
cofibrations and fibrations in $\mathrm{Top}$. We refer the reader to
\cite{Barthel-R_functorial-factorizations, Cole_many-homotopy, Schwanzl-V_cofibrations} for a general
discussion of homotopy theory in topological categories. The
following is an extension of Str\o m's model structure~\cite{Strom_homotopy} to arbitrary diagram spaces
in $\mathrm{Top}$.
\begin{proposition}
The classes of homotopy equivalences, $h$-cofibrations, and
$h$-fibrations specify a model structure on $\mathrm{Top}^{{\mathcal C}}$.
\end{proposition}
\begin{proof}
One may either reuse Str\o m's proof in the setting of ${\mathcal C}$-spaces,
or one may apply the general criterion for the existence of such
model structures formulated by Cole~\cite{Cole_many-homotopy} and
corrected by
Barthel-Riehl~\cite{Barthel-R_functorial-factorizations}. In the
latter approach there are two conditions that must be checked.
The first is the assumption that $\mathrm{Top}^{{\mathcal C}}$ is locally bounded which
is needed to apply~\cite[Corollary
5.18]{Barthel-R_functorial-factorizations}. By a similar argument as
in~\cite[Example 5.14]{Barthel-R_functorial-factorizations},
$\mathrm{Top}^{{\mathcal C}}$ inherits this property from $\mathrm{Top}$. The second
condition is that the $h$-cofibrations are exactly the maps that
have the left lifting property with respect to $h$-fibrations that
are homotopy equivalences (i.e., in the formulation of Cole that the
class of $h$-cofibrations equals the class of strong
$h$-cofibrations). For this we use the characterization of
Schw\"anzl-Vogt~\cite[Proposition 3.5(6)] {Schwanzl-V_cofibrations}
which shows that a map $f\colon X\to Y$ has the left lifting
property with respect to $h$-fibrations that are homotopy
equivalences if and only if the induced map $M(f)\to Y\times I$ has
the left lifting property with respect to all $h$-fibrations. It
follows from the proof of \cite[Theorem 4]{Strom_note} that the
latter condition is satisfied if $f$ is an $h$-cofibration (Str\o{}m
formulates the result for the category of all topological spaces but
the argument works the same for the category $\mathrm{Top}^{{\mathcal C}}$).
\end{proof}
\begin{lemma}\label{lem:functors-preserving-h-cof}
If ${\mathcal C}$ and ${\mathcal B}$ are small categories and
$F\colon \mathrm{Top}^{{\mathcal C}}\to\mathrm{Top}^{{\mathcal B}}$ is a functor that preserves colimits
and tensors, then $F$ also preserves $h$-cofibrations
\end{lemma}
\begin{proof}
This easily follows from the mapping
cylinder retract characterization.
\end{proof}
In particular, this applies to the functor
$\hocolim_{{\mathcal C}}\colon \mathrm{Top}^{{\mathcal C}}\to \mathrm{Top}$ and the evaluation
functor $\Ev_{\bld{k}}\colon \mathrm{Top}^{{\mathcal C}}\to \mathrm{Top}$ for an object $\bld{k}$ in
${\mathcal C}$.
\begin{proof}[Proof of Proposition \ref{prop:h-cofibration-properties} in
the topological setting]
For (i) it suffices to show that an ${\mathcal A}$-relative cofibration has the left lifting property with respect to the map $p_0\colon Z^I \to Z$ for any
$\mathcal K$-space $Z$. We know that the ${\mathcal A}$-relative
cofibrations have the left lifting property with respect to the acyclic ${\mathcal A}$-relative $\mathcal K$-fibrations, that is, the maps $X\to Y$ such that $X(\mathbf k)^H\to Y(\mathbf k)^H$ is an acyclic fibration for all objects
$\mathbf k$ in ${\mathcal A}$ and all subgroups $H\subseteq{\mathcal A}(\mathbf k)$. The map
$p_0^H(\mathbf k)$ can be identified with the
evaluation map $(Z(\mathbf k)^H)^I\to Z(\mathbf k)^H$ and the result follows
since this is even a homotopy equivalence and a Hurewicz fibration.
Part (ii) is immediate since the $h$-cofibrations are the cofibrations in a
model structure. For the gluing lemma (iii), we are given a
commutative diagram as in (\ref{h-cofibrant-diagram}) in which the vertical
maps are ${\mathcal A}$-relative level equivalences. Since the $h$-cofibrations in
$\mathrm{Top}^{\mathcal K}$ are object-wise $h$-cofibrations, the claim follows from the gluing lemma for $h$-cofibrations and weak equivalences in $\mathrm{Top}$, see e.g.\ \cite[Appendix, Proposition 4.8(b)]{Boardman-V_homotopy-invariant}.
The gluing lemma for the ${\mathcal K}$-equivalences in (iv) follows for the same reason because $\hocolim$ preserves colimits and $h$-cofibrations.
To show (v), we use that the Str\o m model structure on $\mathrm{Top}^{{\mathcal K}}$
induces a Reedy model structure on the category of $\lambda$-diagrams
in $\mathrm{Top}^{{\mathcal K}}$ in which a morphism $X\to Y$ is a weak equivalence or
fibration if and only if $X_{\alpha}\to Y_{\alpha}$ is respectively a
homotopy equivalence or $h$-fibration for all $\alpha<\lambda$. In
this model structure a $\lambda$-sequence is cofibrant if and only if
each map $X_{\alpha}\to X_{\alpha+1}$ is an $h$-cofibration, hence it
follows from general results about Reedy model structures
\cite[Theorem 19.9.1]{Hirschhorn-model} that the canonical map from
the homotopy colimit to the colimit is a homotopy equivalence.
For (vi), we notice that the functor $X\boxtimes(-)$ preserves colimits because
it is a left adjoint. Since it also preserves tensors,
the claim follows from Lemma \ref{lem:functors-preserving-h-cof}
and part~(i).
In (vii) we may assume that $f$ has the form $G_{\bld{k}}^{{\mathcal K}}({\mathcal K}(\bld{k})/H
\times i)$ for an object $\bld k$ in ${\mathcal A}$, a subgroup $H\subseteq {\mathcal A}(\bld k)$,
and $i$ a generating cofibration in $\mathrm{Top}$. Using
Lemma~\ref{lem:product_of_free_Jspaces} we have the identification
\[
f^{\Box n} \cong G_{\bld{k}^{\concat n}}^{{\mathcal K}}({\mathcal K}(\bld{k}^{\concat n})/
H^{\times n}) \times i^{\Box n}
\]
where $i^{\Box n}$ is the iterated pushout-product map in $\mathrm{Top}$. The category
$\mathrm{Top}^{\mathcal K\times\Sigma_n}$ is tensored over the category
$\mathrm{Top}^{\Sigma_n}$
of $\Sigma_n$-spaces and we may view the above map as the tensor of the identity
on $G_{\bld{k}^{\concat n}}^{{\mathcal K}}({\mathcal K}(\bld{k}^{\concat n})/H^{\times n})$ with $i^{\Box n}$.
Since the latter is an $h$-cofibration in $\mathrm{Top}^{\Sigma_n}$ (it is the realization of a
$\Sigma_n$-equivariant map of simplicial sets) and every object in
$\mathrm{Top}^{\mathcal K\times \Sigma_n}$ is $h$-cofibrant, we conclude from the
pushout-product axiom for $h$-cofibrations \cite[Corollary 2.9]{Schwanzl-V_cofibrations}
that $f^{\Box n}$ is an $h$-cofibration in $\mathrm{Top}^{\mathcal K\times \Sigma_n}$.
Applying Lemma \ref{lem:functors-preserving-h-cof} to the functor
$Z\boxtimes_{\Sigma_n}(-)$ from $({\mathcal K} \times \Sigma_n)$-spaces to
${\mathcal K}$-spaces then proves (vii).
The smallness requirement (viii) follows because $h$-cofibrations in
$\mathrm{Top}^{{\mathcal K}}$ are
object-wise $h$-cofibrations and any space is small relative to the $h$-cofibrations
in $\mathrm{Top}$ by \cite[Lemma 2.4.1]{Hovey-model}. This uses that in the category
of compactly generated weak Hausdorff spaces, a Hurewicz
cofibration is a closed inclusion.
\end{proof}
Notice, that the argument actually proves a stronger version of (vi): the functor
$X\boxtimes (-)$ preserves $h$-cofibrations in general. Similarly, one can show that the map $Z\boxtimes_{\Sigma_n} f^{\Box n}$ in (vii) is an $h$-cofibration provided only that $f$ is an $h$-cofibration.
\subsection{Simplicial \texorpdfstring{$h$}{h}-cofibrations}\label{simplicial-h-cofibration}
The situation in the simplicial case is easier than in the topological
case. Here, we say that a map of $\mathcal K$-spaces is an $h$-cofibration
if it is a levelwise cofibration of simplicial sets. We recall that the functor $\hocolim_{{\mathcal K}}$ preserves colimits and sends $h$-cofibrations to cofibrations in ${\mathcal S}$ \cite[Theorem
18.5.1]{Hirschhorn-model}.
\begin{proof}[Proof of Proposition \ref{prop:h-cofibration-properties} in
the simplicial setting]
The generating cofibrations are levelwise cofibrations, so (i)
follows and (ii) is clear. For (iii), (iv), and (v), the same arguments as in the
topological case apply. For (vi), we first consider a generating ${\mathcal A}$-relative
cofibration of the form $G_{\bld k}^{{\mathcal K}}({\mathcal K}(\bld k)/H\times i)$ with
$i$ a generating cofibration in ${\mathcal S}$. Then $X\boxtimes f$ can be identified with the map $X\boxtimes G_{\bld k}^{{\mathcal K}}({\mathcal K}(\bld k)/H)\times i$ which is clearly a level cofibration. This implies that $X\boxtimes(-)$ maps the generating ${\mathcal A}$-relative cofibrations and hence all
${\mathcal A}$-relative cofibrations to level cofibrations. An analogous argument using the description of $f^{\Box n}$ derived in the topological setting shows that $Z\boxtimes f^{\Box n}$ is a level cofibration. Together with the fact that an injective map of $\Sigma_n$-sets induces an injective map on
$\Sigma_n$-orbits this imply (vii). Finally, (viii) follows as in the topological case from the fact that any simplicial set
is small relative to the cofibrations. (In fact, the category of $\mathcal K$-spaces is locally presentable in the simplicial setting, so all objects are small relative to the whole category.)
\end{proof}
\section{Monoidal properties of the model structures}\label{sec:monoidal-properties}
In this section we verify the pushout-product axiom and the monoid
axiom for the ${\mathcal A}$-relative level and ${\mathcal K}$-model structures on ${\mathcal S}^{{\mathcal K}}$ for a well-structured relative index category $({\mathcal K},{\mathcal A})$.
We begin by stating an easy lemma which will also be useful in
the analysis of structured diagram spaces in the next section.
The lemma concerns the following situation: We
have a ${\mathcal K}$-equivalence $X\to Y$ which is $G$-equivariant with
respect to an action of a discrete group $G$ and we would like to
conclude that the induced map of $G$-orbit spaces is again a
${\mathcal K}$-equivalence.
\begin{lemma}\label{lem:orbit-lemma}
Let $G$ be a discrete group and consider a commutative diagram of
$G$-objects in ${\mathcal S}^{{\mathcal K}}$,
\[
\xymatrix@-1pc{
X\ar[rr]\ar[rd]_p & &Y\ar[dl]^q\\
& E &
}
\]
where we assume that the $G$-action on $E(\bld{k})$ is free for all objects
$\bld{k}$ in ${\mathcal K}$. In the topological case we make the additional
assumption that either (i) $E$ is the geometric realization of a ${\mathcal K}$-diagram in $G$-simplicial sets, or (ii) that $G$ is finite and $E$ is object-wise Hausdorff (and not just weak Hausdorff). Then if $X\to Y$ is a ${\mathcal K}$-equivalence so is the induced map of orbit ${\mathcal K}$-spaces $X/G\to Y/G$. The analogous statement holds for the
${\mathcal A}$-relative level equivalences.
\end{lemma}
When applying this to a given map $X\to Y$ we usually only specify the map $q$ in the lemma. It will then be understood that $p$ is defined as the composition.
\begin{proof}
Passing to $G$-orbits commutes with homotopy colimit, so it suffices to check that if $X_{h{\mathcal K}}\to Y_{h{\mathcal K}}$ is a weak equivalence, then
$X_{h{\mathcal K}}/G\to Y_{h{\mathcal K}}/G$ is also a weak equivalence. We shall see that
the stated conditions on $E$ imply that the projections $X_{h{\mathcal K}}\to X_{h{\mathcal K}}/G$ and $Y_{h{\mathcal K}}\to Y_{h{\mathcal K}}/G$ are covering maps. The conclusion in the lemma will then follow from the exact sequence of homotopy groups associated with a covering map.
Let us first consider the simplicial case. The condition that $E$ be object-wise $G$-free implies that $E_{h{\mathcal K}}$ is $G$-free, hence $X_{h{\mathcal K}}$ and
$Y_{h{\mathcal K}}$ are also $G$-free. The conclusion in the simplicial setting therefore follows from the general fact that the projection onto the orbit space of a free group action is a covering map.
Next consider the topological case and suppose that the assumption (i) in the lemma holds. Then, by the above discussion, $E_{h{\mathcal K}}\to E_{h{\mathcal K}}/G$ is the geometric realization of a simplicial covering map, hence a topological covering map \cite[III, Satz 7.6]{Lamotke-semi}. This is equivalent to the condition that $G$ acts properly discontinuously on $E_{h{\mathcal K}}$: For every point $e$ in $E_{h{\mathcal K}}$ there is an open neighborhood $U_e$ such that $gU_e \cap U_e \neq \emptyset$ implies $g=1$. Pulling such a neighborhood back via $p_{h{\mathcal K}}$ or $q_{h{\mathcal K}}$ shows that $G$ also acts properly discontinuously on $X_{h{\mathcal K}}$ and
$Y_{h{\mathcal K}}$ which gives the result. Now suppose instead that the assumption (ii) in the lemma holds. The condition that $E$ be object-wise Hausdorff implies that $E_{h{\mathcal K}}$ is also Hausdorff, see e.g.~\cite[Lemma 11.3]{May_geometry}. Since we assume $G$ to be finite this ensures that $G$ acts properly discontinuously on $E_{h{\mathcal K}}$ and the argument now proceeds as above.
For the ${\mathcal A}$-relative level equivalences, we apply the previous argument before taking homotopy colimits.
\end{proof}
The verification of the pushout-product axiom and the monoid axiom is based on the next proposition which is useful in its own right.
\begin{proposition}\label{prop:boxtimes-flat-invariance}
If the ${\mathcal K}$-space $X$ is ${\mathcal A}$-relative cofibrant, then the
functor $X\boxtimes(-)$ preserves ${\mathcal A}$-relative level equivalences and
${\mathcal K}$-equivalences.
\end{proposition}
\begin{proof}
We start with the ${\mathcal K}$-equivalences and first consider the case
where $X$ has the form $G_{\bld k}^{{\mathcal K}}({\mathcal K}(\bld k)/H\times L)$ for an object $\bld k$ in ${\mathcal A}$, a subgroup $H\subseteq {\mathcal A}(\bld k)$, and a space
$L$ with trivial ${\mathcal K}(\bld k)$-action. Then $X\boxtimes Y$ is isomorphic to $(F^{{\mathcal K}}_{\bld k}(L)\boxtimes Y)/H$ for any ${\mathcal K}$-space $Y$ (where $H$ acts trivially on $Y$). We know from
Lemma~\ref{lem:k-Kan-extension} that
$(F^{{\mathcal K}}_{\bld{k}}(*)\boxtimes Y)(\bld{m})$ is naturally isomorphic to
$\colim_{\bld{k}\concat\bld{l}\to\bld{m}}Y(\bld{l})$; the colimit calculated over
the category $(\bld{k}\sqcup -\downarrow \bld{m})$.
The colimit decomposes as a coproduct indexed by the components of
this category and the same holds for the analogous homotopy colimit.
Each of the connected components of the category in question has a
terminal object by condition (ii) for a well-structured relative index category.
Hence the Kan extension and the homotopy Kan extension of $Y$ along the
functor $\mathbf k\sqcup-$ are equivalent, i.e., the canonical map
\[
\operatornamewithlimits{hocolim}_{\bld{k}\concat\bld{l}\to\bld{m}}Y(\bld{l})
\to \operatornamewithlimits{colim}_{\bld{k}\concat\bld{l}\to\bld{m}}Y(\bld{l})
\]
is a weak equivalence for each $\bld{m}$. Given a map $Y\to Z$ we therefore have a commutative diagram
\[
\xymatrix@-1pc{
(F_{\bld{k}}^{{\mathcal K}}(L)\boxtimes Y)_{h{\mathcal K}} \ar[rr] & &
(F_{\bld{k}}^{{\mathcal K}}(L)\boxtimes Z)_{h{\mathcal K}}\\
L\times \displaystyle{\operatornamewithlimits{hocolim}_{\bld{m}\in {\mathcal K}}}\
\displaystyle{\operatornamewithlimits{hocolim}_{\bld{k}\concat\bld{l}\to\bld{m}}}
Y(\bld l) \ar[rr]\ar[u]_{\sim}\ar[d]^{\sim} & &
L\times \displaystyle{\operatornamewithlimits{hocolim}_{\bld{m}\in {\mathcal K}}}\
\displaystyle{\operatornamewithlimits{hocolim}_{\bld{k}\concat\bld{l}\to\bld{m}}}
Z(\mathbf l) \ar[u]_{\sim}\ar[d]^{\sim}\\
L\times Y_{h{\mathcal K}} \ar[rr] & &L\times Z_{h{\mathcal K}}
}
\]
in which the vertical maps are weak equivalences as indicated. (The
vertical maps in the bottom part of the diagram are the canonical weak
equivalences associated to the homotopy colimit of a homotopy Kan
extension, see e.g.~\cite[Lemma 1.4]{Schlichtkrull-infinite}.) Thus,
if $Y\to Z$ is a ${\mathcal K}$-equivalence, then the map on the top of the
diagram is a weak equivalence. Consider then the canonical map
$F_{\bld{k}}^{{\mathcal K}}(L)\boxtimes Z\to F_{\bld{k}}^{{\mathcal K}}(*)\boxtimes * $
induced by the projections onto the terminal objects $*$ in ${\mathcal S}$ and
${\mathcal S}^{{\mathcal K}}$. It follows from Lemma \ref{lem:free-right-action} that the
$H$-action on the target of this map is object-wise
free, hence Lemma \ref{lem:orbit-lemma} ensures that the top map in
the above diagram induces a weak equivalence of
$H$-orbit spaces. This is exactly the statement of
the proposition when $X$ has this special form.
For the next step we assume that $X_0 \boxtimes(-)$ preserves
${\mathcal K}$-equivalences and that $X_1$ arises from $X_0$
as the pushout obtained by attaching an ${\mathcal A}$-relative generating cofibration. Then parts (iv) and (vi) of
Proposition~\ref{prop:h-cofibration-properties} and the preceding argument show that $X_1 \boxtimes(-)$ also preserves ${\mathcal K}$-equivalences.
In general, any ${\mathcal A}$-relative cofibrant $\mathcal K$-space
$X$ is a retract of a cell complex constructed from the
generating ${\mathcal A}$-relative cofibrations, hence we may assume that $X$ is itself such a cell complex. This means that there is an ordinal $\lambda$ and a $\lambda$-sequence
$\{X_\alpha\colon \alpha<\lambda\}$ such that $X_0=\emptyset$, $X=\colim_{\alpha<\lambda}X_{\alpha}$, and each of the maps
$X_{\alpha}\to X_{\alpha+1}$ is the pushout of a generating ${\mathcal A}$-relative
cofibration. By an inductive argument based on the above we conclude that each of the functors $X_{\alpha}\boxtimes (-)$ preserves
${\mathcal K}$-equivalences and the conclusion now follows from parts (v) and (vi) of Proposition \ref{prop:h-cofibration-properties}
and the homotopy invariance of homotopy colimits.
The statement for the ${\mathcal A}$-relative level equivalences follows from an analogous induction argument where again the basic case is a consequence of Lemma \ref{lem:free-right-action}.
\end{proof}
\subsection{The pushout-product axiom} Recall that given maps
$f_1\colon X_1\to Y_1$ and $f_2\colon X_2\to Y_2$ in ${\mathcal S}^{{\mathcal K}}$,
the pushout-product is the induced map
\[
f_1\Box f_2\colon Y_1\boxtimes X_2\cup_{X_1\boxtimes X_2}X_1\boxtimes
Y_2\to Y_1\boxtimes Y_2.
\]
We say that a model structure on ${\mathcal S}^{{\mathcal K}}$ satisfies the pushout-product axiom if given two cofibrations $f_1$ and $f_2$, the pushout-product $f_1 \Box f_2$ is a cofibration that is in addition acyclic if $f_1$ or $f_2$ is. By definition, see \cite{Schwede_S-algebras}, a monoidal model category is a closed symmetric monoidal category equipped with a model structure that satisfies the pushout-product axiom.
\begin{proposition}[The pushout-product
axiom]\label{prop:K-pushout-product-axiom}
The ${\mathcal A}$-relative level and ${\mathcal K}$-model structures on
${\mathcal S}^{{\mathcal K}}$ satisfy the pushout-product axiom.
\end{proposition}
\begin{proof}
It suffices by~\cite[Lemma 3.5]{Schwede_S-algebras} to consider the generating (acyclic) cofibrations for the model structures. We start with the
${\mathcal A}$-relative ${\mathcal K}$-model structure. For $s=1,2$, assume we
are given generating cofibrations
$f_s =G_{\bld{k_s}}^{{\mathcal K}}({\mathcal K}(\bld{k_s})/H_s \times i_s)$ with $\bld k_s$ in
${\mathcal A}$, $H_s\subseteq {\mathcal A}(\bld k_s)$, and $i_s$ a generating cofibration in
${\mathcal S}$. Then Lemma \ref{lem:product_of_free_Jspaces} implies that $f_1
\Box f_2$ is isomorphic to
\[
G_{\bld{k_1}\concat\bld{k_2}}^{{\mathcal K}}({\mathcal K}(\bld{k_1}\concat\bld{k_2})/(H_1
\times H_2) \times i_1 \Box i_2), \] where $i_1\Box i_2$ is the pushout product in
$\mathcal S$. The latter is a cofibration because the pushout-product axiom in ${\mathcal S}$ holds and the condition that ${\mathcal A}$ be a multiplicative subcategory of automorphisms therefore implies that $f_1\Box f_2$ is an ${\mathcal A}$-relative
cofibration. Suppose then that $f_1\colon X_1\to Y_1$ is a generating acyclic
${\mathcal A}$-relative cofibration and that $f_2\colon X_2\to Y_2$ is a generating
${\mathcal A}$-relative cofibration. Consider the diagram
\[
\xymatrix@-1pc{
Y_1\boxtimes X_2 \ar[d]_{\id} & X_1\boxtimes X_2 \ar[l] \ar[r]\ar[d] & X_1\boxtimes Y_2 \ar[d]\\
Y_1\boxtimes X_2 & Y_1\boxtimes X_2 \ar[l]_{\id}\ar[r]& Y_1\boxtimes Y_2
}
\]
and notice that $f_1\Box f_2$ may be identified with the induced map
of pushouts. Since $X_2$ and $Y_2$ are ${\mathcal A}$-relative cofibrant by definition, it follows from Proposition \ref{prop:boxtimes-flat-invariance} that the vertical maps are ${\mathcal K}$-equivalences. The map of pushouts is therefore also a
${\mathcal K}$-equivalence by Proposition \ref{prop:h-cofibration-properties}(iv).
The cofibrations for the ${\mathcal A}$-relative level model structure
are the same as for the ${\mathcal A}$-relative ${\mathcal K}$-model structure and the
second part of the argument follows from the same argument used above for the ${\mathcal K}$-equivalences.
\end{proof}
\subsection{The monoid axiom}
Recall from ~\cite{Schwede_S-algebras} that a
monoidal model category ${\mathcal C}$ satisfies the monoid axiom if the
following holds: If ${\mathcal M}$ denotes the class of maps of the form
$X\boxtimes (Y \to Z)$ with $X$ an arbitrary object and $Y \to Z$
an acyclic cofibration, then any map obtained from ${\mathcal M}$
by cobase change and transfinite composition is a weak equivalence
in ${\mathcal C}$.
\begin{proposition}[The monoid axiom]\label{prop:monoid-axiom}
The ${\mathcal A}$-relative level and ${\mathcal K}$-model structures on
${\mathcal S}^{{\mathcal K}}$ satisfy the monoid axiom.
\end{proposition}
\begin{proof}
For each of the model structures it suffices by
\cite[Lemma 3.5]{Schwede_S-algebras} to consider the subclass of ${\mathcal M}$ given
by the maps $X\boxtimes (Y\to Z)$ where $Y\to Z$ is a generating acyclic
cofibration. We start with the ${\mathcal A}$-relative ${\mathcal K}$-model structure. Let
$f\colon Y\to Z$ be a generating acyclic ${\mathcal A}$-relative cofibration
and let $X$ be an arbitrary ${\mathcal K}$-space. We choose a cofibrant replacement
$X' \to X$ and consider the diagram
\[
\xymatrix@-1pc{
X'\boxtimes Y \ar[r]\ar[d] & X'\boxtimes Z\ar[d]\\
X\boxtimes Y \ar[r] & X\boxtimes Z. }
\]
Since $Y$ and $Z$ are ${\mathcal A}$-relative cofibrant it follows
from Proposition \ref{prop:boxtimes-flat-invariance} that the
vertical and the upper horizontal maps are
${\mathcal K}$-equivalences, hence the same holds for $X \boxtimes f$. Proposition
\ref{prop:h-cofibration-properties}(vi) implies that $X \boxtimes f$ is also
an $h$-cofibration and since the property of being both a ${\mathcal K}$-equivalence and an $h$-cofibration in preserved under pushouts and transfinite composition by Corollaries \ref{cor:pushout-K-or-level-h}
and \ref{cor:transfinite-K-or-level-equivalence}, we get the result for the
${\mathcal A}$-relative ${\mathcal K}$-model structure. An analogous argument gives the
${\mathcal A}$-relative level version.
\end{proof}
\section{Structured diagram spaces}\label{sec:structured-section}
By an operad $\mathcal D$ in ${\mathcal S}$ we understand a sequence of spaces
$\mathcal D(n)$ in ${\mathcal S}$ for $n\geq 0$ such that $\mathcal D(0)=*$,
there is a unit $\{1\}\to\mathcal D(1)$, each of the spaces $\mathcal D(n)$
comes equipped with an action of $\Sigma_n$, and there are structure
maps
\[
\gamma\colon \mathcal D(k)\times \mathcal
D(j_1)\times\dots\times\mathcal D(j_k)\to \mathcal D(j_1+\dots+j_k)
\]
satisfying the defining relations listed in~\cite[Definition
1.1]{May_geometry}.
\begin{definition}\label{def:Sigma-free}
An operad $\mathcal D$ in ${\mathcal S}$ is $\Sigma$-free if the
$\Sigma_n$-action on $\mathcal D(n)$ is free for all $n$. In the
topological setting ${\mathcal S}=\mathrm{Top}$ we make the additional
assumption that the spaces of a $\Sigma$-free operad be Hausdorff
(not just weak Hausdorff).
\end{definition}
With this terminology, an $E_{\infty}$ operad in the sense of
\cite{May_geometry} is the same thing as a $\Sigma$-free operad
$\mathcal D$ (in the topological setting) such that each of the spaces
$\mathcal D(n)$ is contractible. An operad $\mathcal D$ in ${\mathcal S}$ gives
rise to a monad $\mathbb D$ on ${\mathcal S}^{{\mathcal K}}$ in the usual way by letting
\begin{equation}\label{eq:monoad-from-operad}
\mathbb D(X)=\coprod_{n\geq 0}\mathcal D(n)\times_{\Sigma_n}X^{\boxtimes n}.
\end{equation}
Here $X^{\boxtimes 0}$ denotes the unit $\bld1_{{\mathcal K}}$
for the monoidal structure on ${\mathcal S}^{{\mathcal K}}$. We write ${\mathcal S}^{{\mathcal K}}[\mathbb
D]$ for the category of $\mathbb D$-algebras in ${\mathcal S}^{{\mathcal K}}$. By a {\em
structured ${\mathcal K}$-space} we understand a ${\mathcal K}$-space equipped with
such an algebra structure.
\begin{lemma}\label{structured-cocomplete}
The category ${\mathcal S}^{{\mathcal K}}[\mathbb D]$ is complete and cocomplete and
the forgetful functor from ${\mathcal S}^{{\mathcal K}}[\mathbb D]$ to ${\mathcal S}^{{\mathcal K}}$
preserves limits and filtered colimits.
\end{lemma}
\begin{proof}
By~\cite[Proposition 4.3.6]{Borceux_Handbook2}, the category
${\mathcal S}^{{\mathcal K}}[\mathbb D]$ is complete and cocomplete provided that the functor
$\mathbb D \colon {\mathcal S}^{{\mathcal K}} \to {\mathcal S}^{{\mathcal K}}$ preserves filtered
colimits. By the definition of $\mathbb D$
in~\eqref{eq:monoad-from-operad}, this reduces to showing that
$(-)^{\boxtimes n}$ preserves filtered colimits. Since ${\mathcal S}^{{\mathcal K}}$
is closed monoidal, the iterated $\boxtimes$-product preserves
colimits as a functor $({\mathcal S}^{{\mathcal K}})^{\times n} \to
{\mathcal S}^{{\mathcal K}}$. Combining this with the fact that the diagonal functor
associated with a filtered category is final in the sense
of~\cite[Section IX.3]{MacLane_working} shows the claim.
The forgetful functor preserves limits by~\cite[Proposition
4.3.1]{Borceux_Handbook2}, and~\cite[Proposition~4.3.2]{Borceux_Handbook2} shows that it also preserves
filtered colimits since $\mathbb D \colon {\mathcal S}^{{\mathcal K}} \to {\mathcal S}^{{\mathcal K}}$
does.
\end{proof}
Given a well-structured relative index category $({\mathcal K},{\mathcal A})$, we say that the
${\mathcal A}$-relative level or the ${\mathcal A}$-relative ${\mathcal K}$-model structure on ${\mathcal S}^{{\mathcal K}}$ lifts to ${\mathcal S}^{{\mathcal K}}[\mathbb D]$ if there exists a model
structure on ${\mathcal S}^{{\mathcal K}}[\mathbb D]$ whose weak equivalences and
fibrations are the weak equivalences and fibrations of the underlying
${\mathcal K}$-spaces. Since our model structures on ${\mathcal S}^{{\mathcal K}}$ are
cofibrantly generated, there is a well-known strategy for constructing
such a lift: Let us generically write $I_{{\mathcal K}}$ for the generating
cofibrations and $J_{{\mathcal K}}$ for the generating acyclic cofibrations for the
given model structure on ${\mathcal S}^{{\mathcal K}}$. Then we let $\mathbb D(I_{{\mathcal K}})$ and
$\mathbb D(J_{{\mathcal J}})$ be the corresponding sets of morphisms in
${\mathcal S}^{{\mathcal K}}[\mathbb D]$, obtained by applying the functor $\mathbb D$,
and we may ask if these sets satisfy the conditions in the recognition
principle for cofibrantly generated model categories~\cite[Theorem
2.1.19]{Hovey-model}. This will not always be the case as for instance not every ${\mathcal A}$-relative ${\mathcal K}$-model structure lifts to the category of
commutative ${\mathcal K}$-space monoids ${\mathcal C}{\mathcal S}^{{\mathcal K}}$ (which is the same as
${\mathcal S}^{{\mathcal K}}[\mathbb C]$ where $\mathbb
C$ is the monad associated to the commutativity operad $\mathcal C$
with $\mathcal C(n)=*$ for all $n$). Roughly speaking, for this
strategy to be successful, the $\Sigma_n$-action on $\mathcal
D(n)\times X^{\boxtimes n}$ must be sufficiently free and this can be
obtained by imposing either a freeness assumption on the operad spaces
$\mathcal D(n)$ or on the iterated $\boxtimes$-products
$X^{\boxtimes n}$. This is reflected in the following lifting result where we use the notion of a very well-structured relative index category introduced in Definition~\ref{def:very-well-structured}.
\begin{proposition}\label{prop:structured-lift-proposition}
The ${\mathcal A}$-relative level and ${\mathcal K}$-model structures on
${\mathcal S}^{{\mathcal K}}$ lift to model structures on ${\mathcal S}^{{\mathcal K}}[\mathbb D]$ provided that either (i) $\mathcal D$ is $\Sigma$-free, or (ii) $({\mathcal K},{\mathcal A})$ is very
well-structured. Under one of these assumptions, the lifted model structure on ${\mathcal S}^{{\mathcal K}}[\mathbb D]$ is cofibrantly generated with generating (acyclic) cofibrations obtained by applying $\mathbb D$ to the generating (acyclic) cofibrations for the given model structure on ${\mathcal S}^{{\mathcal K}}$.
\end{proposition}
We shall also use the terms \emph{${\mathcal A}$-relative level} and \emph{${\mathcal A}$-relative ${\mathcal K}$-model structure} for the lifted model structures on
${\mathcal S}^{{\mathcal K}}[\mathbb D]$.
For the proof of Proposition~\ref{prop:structured-lift-proposition}
we need some results on pushouts in ${\mathcal S}^{{\mathcal K}}[\mathbb D]$. Suppose
we are given a $\mathbb D$-algebra $A$, a map of ${\mathcal K}$-spaces
$f\colon X\to Y$, and a map of ${\mathcal K}$-spaces $X\to A$. Consider the
associated pushout diagram in ${\mathcal S}^{{\mathcal K}}[\mathbb D]$:
\begin{equation}\label{D-structured-pushout}
\xymatrix@-1pc{
\mathbb D(X)\ar[r]^{\mathbb D(f)}\ar[d] &\mathbb D(Y)\ar[d]\\
A\ar[r]^{\bar f} &B
}
\end{equation}
\begin{lemma}\label{structured-h-cofibration-pushout}
Let $\mathcal D$ be any operad in ${\mathcal S}$ and suppose that the map $f$
in \eqref{D-structured-pushout} is an ${\mathcal A}$-relative cofibration. Then
$\bar f$ is an $h$-cofibration.
\end{lemma}
\begin{lemma}\label{structured-acyclic-pushout}
Suppose that the map $f$ in (\ref{D-structured-pushout}) is a
generating acyclic cofibration for the ${\mathcal A}$-relative level model structure (respectively the ${\mathcal A}$-relative ${\mathcal K}$-model structure). Then $\bar f$ is an
${\mathcal A}$-relative level equivalence (respectively a ${\mathcal K}$-equivalence) of the underlying ${\mathcal K}$-spaces provided that either (i) $\mathcal D$ is
$\Sigma$-free, or (ii) $({\mathcal K},{\mathcal A})$ is very well-structured.
\end{lemma}
The proofs of these lemmas are based on a careful analysis of this
kind of pushout diagrams and will be given in Section
\ref{structured-pushout-proofs}.
\begin{proof}[Proof of Proposition \ref{prop:structured-lift-proposition}]
We use the criterion for lifting model structures to categories of
algebras formulated in~\cite[Lemma 2.3]{Schwede_S-algebras} (which
in turn is an easy consequence of the general recognition principle
for cofibrantly generated model categories \cite[Theorem
2.1.19]{Hovey-model}). Thus, we must check the following two conditions:
\begin{enumerate}[(i)]
\item If $K$ is the set of generating cofibrations or generating
acyclic cofibrations for the given model structure on
${\mathcal S}^{{\mathcal K}}$, then the domains of the maps in $\mathbb D(K)$ are
small relative to $\mathbb D(K)\cell$.
\item If $J_{{\mathcal K}}$ is the set of generating acyclic cofibrations for the given
model structure on ${\mathcal S}^{{\mathcal K}}$, then the maps in
$\mathbb D(J_{{\mathcal K}})\cell$ are
weak equivalences in the same model structure.
\end{enumerate}
The first condition follows from the fact that if $X$ is any
${\mathcal K}$-space, then $\mathbb D(X)$ is small relative to $\mathbb
D(K)\cell$. Indeed, by Lemma \ref{structured-h-cofibration-pushout}, a
relative $\mathbb D(K)$-cell complex in ${\mathcal S}^{{\mathcal K}}[\mathbb D]$ is the
transfinite composition of a sequence whose underlying maps in
${\mathcal S}^{{\mathcal K}}$ are $h$-cofibrations. Since we know from Lemma
\ref{structured-cocomplete} that the forgetful functor from
${\mathcal S}^{{\mathcal K}}[\mathbb D]$ to ${\mathcal S}^{{\mathcal K}}$ preserves filtered colimits, the
above smallness claim follows from the adjointness property and the
fact that by Proposition \ref{prop:h-cofibration-properties}(viii) all
${\mathcal K}$-spaces are small relative to the $h$-cofibrations.
For the second condition we use that by Lemma
\ref{structured-acyclic-pushout} a relative $\mathbb D(J_{{\mathcal K}})$-cell
complex is the transfinite composition of a sequence of maps each of
which is a weak equivalence in the given model structure on ${\mathcal S}^{{\mathcal K}}$. Since these maps are also $h$-cofibrations by
Lemma~\ref{structured-h-cofibration-pushout}, it follows from
Corollary~\ref{cor:transfinite-K-or-level-equivalence} that the transfinite
composition is a weak equivalence.
\end{proof}
\begin{remark}\label{rem:model-str-D-spaces}
Let $\mathcal K$ be the terminal category $*$, viewed as a well-structured index category. For a $\Sigma$-free operad $\mathcal D$, the model structure in Proposition \ref{prop:structured-lift-proposition} is the usual model structure on ${\mathcal S}[\mathbb D]$, see for example \cite[Section~4]{Berger_M-axiomatic}.
\end{remark}
\begin{remark}
Recall from Remark \ref{rem:simpl-top-K-model} the adjoint functors $|-|$ and $\Sing$ relating the simplicial and topological versions of ${\mathcal K}$-spaces. For a monad $\mathbb D$ associated to an operad ${\mathcal D}$ in simplicial sets, write $|\mathbb D|$ for the monad associated to the topological operad $|{\mathcal D}|$ obtained by geometric realization. The fact that $|-|$ is strong monoidal and $\Sing$ is (lax) monoidal easily implies that the above model structures on $\mathbb D$-algebras in the simplicial setting are Quillen equivalent to the corresponding model structures on $|\mathbb D|$-algebras in the topological setting whenever they exist. (See the remarks at the beginning of Section \ref{subs:structured-comma} for a general discussion of such adjunctions). This implies in particular that $|-|$ and $\Sing$ gives rise to Quillen equivalences between the relevant model structures on (commutative) ${\mathcal K}$-space monoids in the simplicial and topological settings.
\end{remark}
By definition of the lifted model structures we have the following structured version of the Quillen equivalence in Proposition \ref{prop:well-structured-comparison}.
\begin{proposition}\label{prop:operad-well-structured-comparison}
Let $({\mathcal K},{\mathcal B})$ be a well-structured relative index category. Suppose that
${\mathcal A}$ is a normal and multiplicative subcategory of automorphisms contained in ${\mathcal B}$ and that the inclusion ${\mathcal K}_{{\mathcal A}}\to {\mathcal K}_{{\mathcal B}}$ is homotopy cofinal. Then $({\mathcal K},{\mathcal A})$ is a well-structured relative index category and the following hold:
\begin{enumerate}[(i)]
\item
If ${\mathcal D}$ is a $\Sigma$-free operad, then the ${\mathcal A}$- and ${\mathcal B}$-relative ${\mathcal K}$-model structures on ${\mathcal S}^{{\mathcal K}}[\mathbb D]$ are Quillen equivalent.
\item
If ${\mathcal D}$ is any operad and $({\mathcal K},{\mathcal B})$ is very well-structured,
then $({\mathcal K},{\mathcal A})$ is also very well-structured and the ${\mathcal A}$- and ${\mathcal B}$-relative ${\mathcal K}$-model structures on ${\mathcal S}^{{\mathcal K}}[\mathbb D]$ are Quillen equivalent. \qed
\end{enumerate}
\end{proposition}
It follows from general results for operads acting on objects in suitable symmetric monoidal ${\mathcal S}$-model categories that the model structures on ${\mathcal S}^{{\mathcal K}}[\mathbb D]$ considered above can be viewed as ${\mathcal S}$-model categories in a canonical way. A detailed account of how this works in a topological setting can be found in~\cite[VII.2]{EKMM}. In order to simplify the discussion we shall only consider the case of greatest interest to us: the simplicial model structure on the category of commutative
${\mathcal K}$-space monoids ${\mathcal C}{\mathcal S}^{{\mathcal K}}$. In this case the cotensor is defined on
the underlying ${\mathcal K}$-spaces whereas the tensor and simplicial mapping spaces are defined by
\[ A \otimes K = | [n] \mapsto A^{\boxtimes K_n} | \qquad \text{and}
\qquad \Map(A,B) = \big\{[n]\mapsto {\mathcal C}{\mathcal S}^{{\mathcal K}}(A\otimes
\Delta^{n},B)\big\}
\]
for $A$ and $B$ in ${\mathcal C}{\mathcal S}^{{\mathcal K}}$ and $K$ a simplicial set.
\begin{proposition}\label{prop:CSK-simplicial}
Let $({\mathcal K},{\mathcal A})$ be a very
well-structured index category. Then the ${\mathcal A}$-relative ${\mathcal K}$-model structure on ${\mathcal C}{\mathcal S}^{{\mathcal K}}$
is simplicial.
\end{proposition}
\begin{proof}
The condition for being a simplicial model category can be expressed in terms of the
cotensor structure (see e.g.~\cite[Lemma 4.2.2]{Hovey-model}). Using this, the result follows from the fact that the ${\mathcal A}$-relative ${\mathcal K}$-model structure on ${\mathcal S}^{{\mathcal K}}$ is simplicial by Proposition \ref{prop:K-model-is-S-model-str} and the remarks following that proposition.
\end{proof}
Now we specialize to a well-structured index category ${\mathcal K}$ with homotopy cofinal inclusion ${\mathcal K}_+\to{\mathcal K}$ and the problem of lifting the positive projective ${\mathcal K}$-model structure in Definition \ref{def:K-projective-model} to ${\mathcal C}{\mathcal S}^{{\mathcal K}}$. By definition, $({\mathcal K},O{\mathcal K}_+)$ is very well-structured if $\Sigma_n$ acts freely on the set of connected components of $(\bld k^{\sqcup n}\sqcup-
\downarrow \bld l)$ for each pair of objects $\bld k$ and $\bld l$ in ${\mathcal K}_+$.
\begin{corollary}\label{cor:K-positive-projective-commutative}
Let ${\mathcal K}$ be a well-structured index category and suppose that the inclusion
${\mathcal K}_+\to {\mathcal K}$ is homotopy cofinal and that $({\mathcal K},O{\mathcal K}_+)$ is very
well-structured. Then the positive projective ${\mathcal K}$-model structure on
${\mathcal S}^{{\mathcal K}}$ lifts to a simplicial model structure on ${\mathcal C}{\mathcal S}^{{\mathcal K}}$. \qed
\end{corollary}
\subsection{Change of operads}\label{operad-change-section}
In this section we analyze how our categories of structured diagram
spaces behave under change of operads. Thus, consider a map of operads
$\Phi\colon \mathcal D\to\mathcal E$ and the associated map of monads
$\Phi\colon \mathbb D\to\mathbb E$. This gives rise to a pair of
adjoint functors
\[
\Phi_*\colon {\mathcal S}^{{\mathcal K}}[\mathbb D]\rightleftarrows
{\mathcal S}^{{\mathcal K}}[\mathbb E]\thinspace\colon\Phi^*
\]
where the right adjoint $\Phi^*$ is defined by pulling an $\mathbb
E$-algebra structure on an object back to a $\mathbb D$-algebra
structure via $\Phi$. The $\mathbb E$-algebra $\Phi_*(A)$ associated to a
$\mathbb D$-algebra $A$ with structure map $\xi\colon\mathbb D(A)\to A$
can be represented by a (reflexive) coequalizer diagram in $\mathcal A$,
\[
\xymatrix{
\mathbb E\mathbb D(A)\ar@<0.5ex>[r]^{\partial_0}\ar@<-.5ex>[r]_{\partial_1}& \mathbb E(A)
\ar[r] &\Phi_*(A),
}
\]
where $\partial_0=\mathbb E(\xi)$ and $\partial_1=\mu_A\circ \mathbb
E(\Phi_A)$, see Section \ref{coequalizer-section} for further details. We say that $\Phi$ is a \emph{weak equivalence} if each of the maps $\Phi_n\colon \mathcal D(n)\to\mathcal E(n)$ is a weak equivalence. The next proposition shows that a weak equivalence of operads induces a Quillen equivalence whenever the lifted model structures on
${\mathcal S}^{{\mathcal K}}[\mathbb D]$ and ${\mathcal S}^{{\mathcal K}}[\mathbb E]$ are defined.
\begin{proposition}\label{prop:operad-change}
Let $({\mathcal K},{\mathcal A})$ be a well-structured relative index category and let
$\Phi\colon \mathcal D\to\mathcal E$ be a weak equivalence of operads. Suppose that either (i) both ${\mathcal D}$ and $\mathcal E$ are
$\Sigma$-free, or (ii) $({\mathcal K},{\mathcal A})$ is very well-structured. Then the adjoint functor pair $(\Phi_*,\Phi^*)$ defines a Quillen equivalence between the
${\mathcal A}$-relative ${\mathcal K}$-model structures on ${\mathcal S}^{{\mathcal K}}[\mathbb D]$ and
${\mathcal S}^{{\mathcal K}}[\mathbb E]$.
\end{proposition}
The main technical point in establishing this proposition is the
homotopical analysis of the unit $A\to\Phi^*\Phi_*(A)$ of the
adjunction. The proof of the below lemma requires the same kind of
analysis as the proofs of Lemmas \ref{structured-h-cofibration-pushout} and \ref{structured-acyclic-pushout}, and will be given in Section
\ref{subs:unit-lemma-section}.
\begin{lemma}\label{lem:unit-lemma}
Suppose that $\Phi\colon \mathcal D\to\mathcal E$ is a weak
equivalence and that $A$ is a $\mathbb D$-algebra which is cofibrant in the
${\mathcal A}$-relative ${\mathcal K}$-model structure on ${\mathcal S}^{{\mathcal K}}[\mathbb D]$. Then the
unit for the adjunction $A\to\Phi^*\Phi_*(A)$ is a level equivalence provided
that either (i) the operads $\mathcal D$ and
$\mathcal E$ are $\Sigma$-free, or (ii) $({\mathcal K},{\mathcal A})$ is very well-structured.
\end{lemma}
\begin{proof}[Proof of Proposition \ref{prop:operad-change}]
It is clear that $(\Phi_*,\Phi^*)$ is a Quillen adjunction since
$\Phi^*$ preserves fibrations and acyclic fibrations. Given a
cofibrant object $A$ in ${\mathcal S}^{{\mathcal K}}[\mathbb D]$ and a fibrant object
$B$ in ${\mathcal S}^{{\mathcal K}}[\mathbb E]$, we must show that a map $\Phi_*(A)\to
B$ is a ${\mathcal K}$-equivalence if and only if the adjoint map
$A\to \Phi^*(B)$ is. The latter map admits a factorization
\[
A\to\Phi^*\Phi_*(A)\to \Phi^*(B)
\]
so the result follows from Lemma \ref{lem:unit-lemma} and the 2 out
of 3 property for ${\mathcal K}$-equivalences.
\end{proof}
Specializing to a well-structured index category ${\mathcal K}$ such that
$({\mathcal K},O{\mathcal K}_+)$ is very well-structured, we use the above to rectify
$E_{\infty}$ objects in ${\mathcal S}^{{\mathcal K}}$ to strictly commutative monoids.
\begin{corollary}\label{cor:positive-projective-rectification}
Let ${\mathcal D}$ be an $E_{\infty}$ operad. Suppose that ${\mathcal K}$ is a well-structured index category such that ${\mathcal K}_+\to{\mathcal K}$ is homotopy cofinal and
$({\mathcal K},O{\mathcal K}_+)$ is very well-structured. Then the projective ${\mathcal K}$-model structure on
${\mathcal S}^{{\mathcal K}}[\mathbb D]$ is related to the positive projective ${\mathcal K}$-model structure on ${\mathcal C}{\mathcal S}^{{\mathcal K}}$ by a chain of Quillen equivalences.
\end{corollary}
\begin{proof}
Writing $\pi\colon \mathbb D\to \mathbb C$ for the canonical weak equivalence to the commutativity monad $\mathbb C$, we have a chain of adjoint functors
\[
\xymatrix@-1pc{
({\mathcal C}{\mathcal S}^{{\mathcal K}},\text{positive projective ${\mathcal K}$-model structure})
\ar@<1.0ex>[d]^{\pi^*}\\
({\mathcal S}^{{\mathcal K}}[\mathbb D],\text{positive projective ${\mathcal K}$-model structure})
\ar@<1.0ex>[u]^{\pi_*} \ar@<-1.0ex>[d]\\
({\mathcal S}^{{\mathcal K}}[\mathbb D],\text{projective ${\mathcal K}$-model structure}).
\ar@<-1.0ex>[u]
}
\]
According to Propositions \ref{prop:operad-well-structured-comparison}
and \ref{prop:operad-change}, these Quillen adjunctions specify the required Quillen equivalences.
\end{proof}
\begin{remark}\label{rem:levelwise-version}
The fact that by Lemma \ref{lem:unit-lemma} the unit $A\to\Phi^*\Phi_*(A)$ for the adjunction is a level equivalence implies that there is a levelwise version of Proposition \ref{prop:operad-change}. We shall not use this and leave the details to the interested reader.
\end{remark}
\section{Verification of structured diagram lemmas}
\label{sec:verification}
Let again ${\mathcal D}$ be an operad in ${\mathcal S}$ and let $\mathbb D$ be the associated monad on ${\mathcal S}^{{\mathcal K}}$. In order to prove the technical lemmas on structured diagram spaces stated in Section~\ref{sec:structured-section}, we shall analyze pushout diagrams in ${\mathcal S}^{{\mathcal K}}[\mathbb D]$ of the form
\begin{equation}\label{verification-structured-pushout}
\xymatrix@-1pc{
\mathbb D(X)\ar[r]^{\mathbb D(f)}\ar[d] &\mathbb D(Y)\ar[d]\\
A\ar[r]^{\bar f} &B
}
\end{equation}
for a map $f\colon X\to Y$ in ${\mathcal S}^{{\mathcal K}}$. For this we essentially follow Elmendorf-Mandell~\cite{Elmendorf-M_infinite-loop} by introducing a filtration of the induced map $\bar f$ such that the passage from the $(k-1)$th to the $k$th stage of the filtration can be expressed in terms of a certain
$\Sigma_k$-equivariant ${\mathcal K}$-space $U_k^{\mathbb D}(A)$ constructed from
$A$, together with data derived from the map $f$. Since we shall in fact need a refined version of the filtration considered in \cite{Elmendorf-M_infinite-loop} we have included an exposition of this material in Appendix \ref{app:analysis-operad-algebras}. In the below proposition we extract the facts about the functors $U_k^{\mathbb D}$ needed for the proofs of the above mentioned lemmas. Here $({\mathcal S}^{{\mathcal K}})^{\Sigma_k}/{\mathcal D}(k)$ denotes the category of
$\Sigma_k$-equivariant ${\mathcal K}$-spaces over the constant ${\mathcal K}$-space ${\mathcal D}(k)$ and $Q_{i-1}^i(f)$ is the domain of the $i$-fold iterated pushout-product map as in Section \ref{subs:h-cofibrations}. The statements in the proposition that are not obvious from the definitions appear as Lemma~\ref{freeshifted},
Lemma~\ref{terminal-projection}, and Proposition~\ref{prop:cat-pushout-filtration} in Appendix~\ref{app:analysis-operad-algebras}.
\begin{proposition}\label{prop:pushout-filtration}
There exists a sequence of functors $U_k^{\mathbb D}\colon {\mathcal S}^{{\mathcal K}}[\mathbb D]\to ({\mathcal S}^{{\mathcal K}})^{\Sigma_k}/{\mathcal D}(k)$ for $k\geq 0$, such that the following hold:
\begin{enumerate}[(i)]
\item
$U_0^{\mathbb D}$ is the forgetful functor to ${\mathcal S}^{{\mathcal K}}$.
\item
The functors $U_k^{\mathbb D}$ preserve filtered colimits.
\item
$U_k^{\mathbb D}(\mathbb D(X))$ is isomorphic to $\coprod_{n\geq 0}{\mathcal D}(n+k)\times_{\Sigma_n}X^{\boxtimes n}$ for any ${\mathcal K}$-space $X$.
\item
For a pushout diagram of the type in (\ref{verification-structured-pushout})
there is a natural sequence of $\Sigma_k$-equivariant ${\mathcal K}$-spaces
\[
U_k^{\mathbb D}(A)=F_0 U_k^{\mathbb D}(B)\to
F_1 U_k^{\mathbb D}(B)\to\dots\to F_iU_k^{\mathbb D}(B)\to\dots
\]
such that
$\colim_i F_iU_k^{\mathbb D}(B)=U_k^{\mathbb D}(B)$, the transfinite composition of the sequence equals $U_k^{\mathbb D}(\bar f)$, and
there are $\Sigma_k$-equivariant pushout diagrams in ${\mathcal S}^{{\mathcal K}}$,
\[
\xymatrix@-1pc{
U_{i+k}^{\mathbb D}(A)\boxtimes_{\Sigma_i}Q_{i-1}^i(f) \ar[rr]\ar[d]& &U_{i+k}^{\mathbb D}(A)\boxtimes_{\Sigma_i}Y^{\boxtimes i} \ar[d]\\
F_{i-1} U_k^{\mathbb D}(B) \ar[rr]& & F_iU_k^{\mathbb D}(B)
}
\]
for all $i\geq 1$.\qed
\end{enumerate}
\end{proposition}
For $k=0$, the filtration in (iv) is the filtration of $\bar f$ considered in
\cite{Elmendorf-M_infinite-loop} (for symmetric spectra).
\begin{example}\label{example:U_k-commutative}
If $\mathbb C$ is the monad associated to the commutativity operad, then
$U_k^{\mathbb C}(A)=A$ with trivial $\Sigma_k$-action (see also Example
\ref{ex:commutativity-example} for more details).
\end{example}
\subsection{The proofs of Lemmas \ref{structured-h-cofibration-pushout}
and \ref{structured-acyclic-pushout}}\label{structured-pushout-proofs}
\begin{proof}[Proof of Lemma \ref{structured-h-cofibration-pushout}]
Applying the filtration from Proposition \ref{prop:pushout-filtration} to
$B=U_0^{\mathbb D}(B)$ and writing $F_iB$ for the filtration terms
$F_iU_0^{\mathbb D}(B)$, we get a sequence of pushout diagrams
\[
\xymatrix@-1pc{ U_{i}^{\mathbb D}(A)\boxtimes_{\Sigma_i}Q_{i-1}^i(f)
\ar[rr] \ar[d]&&
U_{i}^{\mathbb D}(A)\boxtimes_{\Sigma_i}Y^{\boxtimes i}\ar[d]\\
F_{i-1}B\ar[rr] && F_iB }
\]
in ${\mathcal S}^{{\mathcal K}}$ such that $A=F_0B$ and $B=\colim_iF_iB$. Proposition
\ref{prop:h-cofibration-properties}(vii) implies that the upper
horizontal map in each of these diagrams is an $h$-cofibration, hence the
same holds for the maps $F_{i-1}B\to F_iB$ and therefore also for
the transfinite composition $A\to B$ by Proposition
\ref{prop:h-cofibration-properties}(ii).
\end{proof}
\begin{proof}[Proof of Lemma \ref{structured-acyclic-pushout}]
We begin with the ${\mathcal A}$-relative ${\mathcal K}$-model structure. Using the filtration from Proposition \ref{prop:pushout-filtration} and arguing as in the
proof of Lemma \ref{structured-h-cofibration-pushout}, we conclude from
Corollaries \ref{cor:pushout-K-or-level-h} and
\ref{cor:transfinite-K-or-level-equivalence} that it is sufficient
to show that
\[
U_i^{\mathbb D}(A)\boxtimes _{\Sigma_i}Q^i_{i-1}(f)\to U_i^{\mathbb
D}(A)\boxtimes_{\Sigma_i}Y^{\boxtimes i}
\]
is a ${\mathcal K}$-equivalence for all $i$. By the pushout-product
axiom, Proposition \ref{prop:K-pushout-product-axiom}, we know that
$Q_{i-1}^i(f)\to Y^{\boxtimes i}$ is an ${\mathcal A}$-relative cofibration and a
${\mathcal K}$-equivalence. Hence it follows from the monoid axiom, Proposition
\ref{prop:monoid-axiom}, that
\[
U_i^{\mathbb D}(A)\boxtimes Q_{i-1}^i(f)\to U_i^{\mathbb
D}(A)\boxtimes Y^{\boxtimes i}
\]
is also a ${\mathcal K}$-equivalence. Thus, it remains to show that the
stated conditions on $\mathcal D$ and $({\mathcal K},{\mathcal A})$ imply that the same holds
for the induced map of $\Sigma_i$-orbits. Suppose first that
$\mathcal D$ is $\Sigma$-free. We know from Proposition
\ref{prop:pushout-filtration} that there is a $\Sigma_i$-equivariant map
$U_i^{\mathbb D}(A)\to \mathcal D(i)$ onto the constant ${\mathcal K}$-space
$\mathcal D(i)$. From this we get the $\Sigma_i$-equivariant map
\[
U_i^{\mathbb D}(A)\boxtimes Y^{\boxtimes i}\to \mathcal
D(i)\boxtimes *\cong \mathcal D(i)\times(*\boxtimes *)\to \mathcal
D(i)\times *
\]
onto the constant ${\mathcal K}$-space ${\mathcal D}(i)$
and since the $\Sigma_i$-action on the latter is object-wise free,
the result follows from Lemma \ref{lem:orbit-lemma}.
Suppose then that $({\mathcal K},{\mathcal A})$ is very well-structured. We first observe that there
is a map $Y\to G^{{\mathcal K}}_{\bld{k}}({\mathcal K}(\bld{k})/H)$ for some object $\bld{k}$
in ${\mathcal A}$ and subgroup $H\subseteq {\mathcal A}(\bld{k})$. From this we get the
$\Sigma_i$-equivariant map
\[
U_i^{\mathbb D}(A)\boxtimes Y^{\boxtimes i}\to *\boxtimes
G^{{\mathcal K}}_{\bld{k}}({\mathcal K}(\bld k)/H)^{\boxtimes i} \cong \left(* \boxtimes
F^{{\mathcal K}}_{\bld{k}^{\concat i}}(*)\right)/H^{\times i}.
\]
Using the isomorphism \eqref{eq:left-Kan}, the value of the ${\mathcal K}$-space on the right at an object $\bld{m}$ in ${\mathcal K}$ can be identified with the
$H^{\times i}$-orbits of the set of connected components in
the comma category $(-\concat\bld{k}^{\concat i}\downarrow
\bld{m})$. The $(\Sigma_i\ltimes H^{\times i})$-action on the set of
connected components is free by the definition of a very well-structured
index category. Hence the $\Sigma_i$-action on the target is object-wise free and the result again follows from Lemma \ref{lem:orbit-lemma}.
For the ${\mathcal A}$-relative level model structures we use the analogous argument with ${\mathcal A}$-relative level equivalences instead of ${\mathcal K}$-equivalences.
\end{proof}
\subsection{The proof of Lemma \ref{lem:unit-lemma}}
\label{subs:unit-lemma-section}
As in Section \ref{operad-change-section} we consider a map of operads
$\Phi\colon \mathcal D\to\mathcal E$ and the adjoint functor pair
$
\Phi_*\colon {\mathcal S}^{{\mathcal K}}[\mathbb D]\rightleftarrows
{\mathcal S}^{{\mathcal K}}[\mathbb E]\thinspace\colon\Phi^*
$.
We first record the action of $\Phi_*$ on free $\mathbb D$-algebras which is a formal consequence of the fact that a composition of left adjoints is again a left adjoint.
\begin{lemma}
For a free $\mathbb D$-algebra $\mathbb D(X)$ we have
$\Phi_*\mathbb D(X)=\mathbb E(X)$.\qed
\end{lemma}
\begin{proof}[Proof of Lemma \ref{lem:unit-lemma}]
Recall that we assume operads to be reduced such that the unit
$\bld 1_{{\mathcal K}}$ for the monoidal structure on ${\mathcal S}^{{\mathcal K}}$ is the initial
object in ${\mathcal S}^{{\mathcal K}}[\mathbb D]$. We may assume without loss of
generality that $A$ is a cell complex in ${\mathcal S}^{{\mathcal K}}[\mathbb D]$
in the sense that it is the colimit of a $\lambda$-sequence
$\{A_{\alpha}\colon \alpha<\lambda\}$ (for an ordinal $\lambda$)
such that $A_0=\bld1_{{\mathcal K}}$ and $A_{\alpha}\to A_{\alpha+1}$ is a
pushout in ${\mathcal S}^{{\mathcal K}}[\mathbb D]$ of a generating cofibration
$\mathbb D(f_{\alpha})\colon \mathbb D(X_{\alpha})\to
\mathbb D(Y_{\alpha})$ (where $f_{\alpha}\colon X_{\alpha}\to Y_{\alpha}$
is a generating ${\mathcal A}$-relative cofibration in ${\mathcal S}^{{\mathcal K}}$). By Lemma
\ref{structured-cocomplete}, $A$ is also the colimit of the ${\mathcal K}$-spaces
$A_{\alpha}$ in the underlying category ${\mathcal S}^{{\mathcal K}}$.
Let us write $A'=\Phi_*(A)$ and
$A_{\alpha}'=\Phi_*(A_{\alpha})$. Since $\Phi_*$ preserves colimits,
$A'$ is the colimit of the $\lambda$-sequence $\{A'_{\alpha}\colon
\alpha<\lambda\}$ in ${\mathcal S}^{{\mathcal K}}[\mathbb E]$ and
$A_{\alpha}'\to A_{\alpha+1}'$ is a pushout of the generating cofibration
$\mathbb E(f_{\alpha})\colon \mathbb E(X_{\alpha})\to
\mathbb E(Y_{\alpha})$.
Again $A'$ is the colimit of the ${\mathcal K}$-spaces $A_{\alpha}'$ in the underlying
category ${\mathcal S}^{{\mathcal K}}$. Since by
Lemma~\ref{structured-h-cofibration-pushout} these are
$\lambda$-sequences of $h$-cofibrations, it suffices by
Proposition~\ref{prop:h-cofibration-properties}(v) to show that $A_{\alpha}\to
A_{\alpha}'$ is a level equivalence in ${\mathcal S}^{{\mathcal K}}$ for all $\alpha$.
In order to set up an inductive argument using Proposition
\ref{prop:pushout-filtration}, we shall in fact prove the following stronger
statement: for each $\alpha<\lambda$ the map $U_k^{\mathbb
D}(A_{\alpha})\to U_k^{\mathbb E}(A_{\alpha}')$ is a level
equivalence for all $k$. The result then follows by setting $k=0$.
Thus, given $\beta<\lambda$, we must show that if the statement holds
for all $\alpha<\beta$, then it also holds for $\beta$. For $\beta=0$
we conclude from Proposition~\ref{prop:pushout-filtration}(iii)
(using that $\mathbb D(\emptyset)=
\bld 1_{{\mathcal K}}$ and $\mathbb E(\emptyset)=\bld1_{{\mathcal K}}$) that the
maps in question can be identified with the maps $\mathcal D(k)\times
\bld 1_{{\mathcal K}}\to \mathcal E(k)\times \bld1_{{\mathcal K}}$ induced by $\Phi$.
These maps
are level equivalences by assumption. If $\beta$ is a limit ordinal,
then it follows from Proposition~\ref{prop:h-cofibration-properties}(v)
and the definition of a $\lambda$-sequence that the statement for
$\alpha<\beta$ implies the statement for $\beta$. Thus, it remains to
consider the case where $\beta$ is a successor ordinal,
$\beta=\alpha+1$. For the inductive step we use the filtration from
Proposition \ref{prop:pushout-filtration}(iv) and observe that there is a commutative diagram
\[
\xymatrix@-1pc{U_k^{\mathbb D}(A_{\alpha}) \ar@{=}[r]\ar[d]
&F_0U_k^{\mathbb D}(A_{\alpha+1})\ar[r] \ar[d] &
F_1U_k^{\mathbb D}(A_{\alpha+1})\ar[r] \ar[d] & F_2U_k^{\mathbb D}(A_{\alpha+1})\ar[r] \ar[d] & \dots \\
U_k^{\mathbb E}(A'_{\alpha}) \ar@{=}[r] &F_0U_k^{\mathbb
E}(A'_{\alpha+1})\ar[r] & F_1U_k^{\mathbb E}(A'_{\alpha+1})\ar[r]
& F_2U_k^{\mathbb E}(A'_{\alpha+1})\ar[r] & \dots }
\]
for each $k\geq 0$. Since the horizontal maps are $h$-cofibrations, by
Proposition \ref{prop:h-cofibration-properties}(ii) and (vii), it suffices to show
that the vertical maps are level equivalences for all $i$. It follows
from the naturality of the filtrations that the map in filtration degree $i$ can be identified with the map of pushouts
induced by the map of diagrams
\[
\xymatrix@-1pc{
F_{i-1}U_k^{\mathbb D}(A_{\alpha+1}) \ar[d] &
U_{i+k}^{\mathbb D}(A_{\alpha})\boxtimes_{\Sigma_i}Q^i_{i-1}(f)
\ar[l] \ar[r] \ar[d] & U_{i+k}^{\mathbb D}(A_{\alpha})\boxtimes_{\Sigma_i}Y^{\boxtimes i} \ar[d] \\
F_{i-1}U_k^{\mathbb E}(A'_{\alpha+1}) & U_{i+k}^{\mathbb E}(A'_{\alpha})\boxtimes_{\Sigma_i}Q^i_{i-1}(f)
\ar[l] \ar[r] & U_{i+k}^{\mathbb E}(A'_{\alpha})\boxtimes_{\Sigma_i}Y^{\boxtimes i}.
}
\]
By induction on $i$ we may assume that the vertical map on the left is
a level equivalence and since the horizontal maps on the right are
$h$-cofibrations, again by Proposition \ref{prop:h-cofibration-properties}(vii),
it is sufficient to show that the two other vertical maps are level equivalences. We know from the pushout-product axiom,
Proposition \ref{prop:K-pushout-product-axiom},
that the ${\mathcal K}$-spaces $Q_{i-1}^{i}(f)$ and $Y^{\boxtimes i}$ are ${\mathcal A}$-relative cofibrant. It therefore follows from the induction hypothesis on $\alpha$ and Proposition~\ref{prop:boxtimes-flat-invariance} that these two vertical maps
are level equivalences before passing to $\Sigma_i$-orbits. Using Lemma
\ref{lem:orbit-lemma} as in the proof of
Lemma \ref{structured-acyclic-pushout} we conclude that the induced maps
of $\Sigma_i$-orbits are also level equivalences.
\end{proof}
\section{Properness for \texorpdfstring{${\mathcal K}$}{K}-spaces}
Recall that a model category is \emph{left proper} if every pushout of a weak equivalence along a cofibration is a weak equivalence and \emph{right proper} if every pullback of a weak equivalence along a fibration is a weak equivalence. A model category is said to be \emph{proper} if it is both left and right proper. This is a desirable property that a model category may or may not have. In this section we discuss properness for our model categories of
${\mathcal K}$-spaces for a well-structured index category $({\mathcal K},{\mathcal A})$.
The main result is Corollary \ref{cor:proper} which states that the ${\mathcal A}$-relative ${\mathcal K}$-model structure on ${\mathcal S}^{{\mathcal K}}$ is proper and that the lifted model structure on ${\mathcal C}{\mathcal S}^{{\mathcal K}}$ is proper when $({\mathcal K},{\mathcal A})$ is very
well-structured.
\subsection{Right properness}
It is an immediate consequence of right properness for the category of spaces ${\mathcal S}$ that the ${\mathcal A}$-relative level model structure on ${\mathcal S}^{{\mathcal K}}$ is right proper. In order to establish right properness for the ${\mathcal A}$-relative ${\mathcal K}$-model structure we need the following lemma.
\begin{lemma}\label{lem:hocolim-pullback}
Let ${\mathcal C}$ be a small category and let ${\mathcal S}^{{\mathcal C}}$ be the associated diagram category. Then the functor $\hocolim_{{\mathcal C}}\colon {\mathcal S}^{{\mathcal C}}\to {\mathcal S}$ preserves pullbacks.
\end{lemma}
\begin{proof}
Consider a pullback square of ${\mathcal C}$-spaces
\[
\xymatrix@-1pc{
A\ar[r] \ar[d] & B\ar[d]\\
C\ar[r] & D.
}
\]
Since realization of simplicial spaces preserves pullback
diagrams it is enough to show that the diagram
\[\xymatrix@-1pc{\displaystyle\coprod_{\bld{c_0} \leftarrow \dots \leftarrow \bld{c_r}}A(\bld{c_r}) \ar[r] \ar[d]&
\displaystyle\coprod_{\bld{c_0} \leftarrow \dots \leftarrow \bld{c_r}}B(\bld{c_r}) \ar[d]\\
\displaystyle\coprod_{\bld{c_0} \leftarrow \dots \leftarrow \bld{c_r}}C(\bld{c_r}) \ar[r]&
\displaystyle\coprod_{\bld{c_0} \leftarrow \dots \leftarrow \bld{c_r}}D(\bld{c_r}) }\]
of simplicial replacements is a pullback diagram in each simplicial degree $r$ and this is clear from the definition.
\end{proof}
\begin{proposition}
The ${\mathcal A}$-relative ${\mathcal K}$-model structure on ${\mathcal S}^{{\mathcal K}}$ is right proper.
\end{proposition}
\begin{proof}
Given a pullback square of ${\mathcal K}$-spaces
\[
\xymatrix@-1pc{
A\ar[r]^{\bar f}\ar[d]_{\bar g} & B\ar[d]^g\\
C\ar[r]^f & D
}
\]
in which $g$ is an ${\mathcal A}$-relative ${\mathcal K}$-fibration and $f$ is a
${\mathcal K}$-equivalence, we must show that $\bar f$ is a
${\mathcal K}$-equivalence. Applying the homotopy colimit functor over
the subcategory ${\mathcal K}_{{\mathcal A}}$ to the diagram we get a diagram of spaces
\[
\xymatrix@-1pc{
\hocolim_{{\mathcal K}_{{\mathcal A}}}A\ar[r] \ar[d] &
\hocolim_{{\mathcal K}_{{\mathcal A}}} B\ar[d]\\
\hocolim_{{\mathcal K}_{{\mathcal A}}}C\ar[r] &
\hocolim_{{\mathcal K}_{{\mathcal A}}}D
}
\]
which we claim is homotopy cartesian. Indeed, since we know from Lemma
\ref{lem:hocolim-pullback} that this is a pullback diagram it suffices to show that the fibers of each of the vertical maps are weakly equivalent to the corresponding homotopy fibers. The latter can be deduced from Lemma \ref{lem:puppe-lemma} and the definition of an ${\mathcal A}$-relative
${\mathcal K}$-fibration. The fact that the diagram is homotopy cartesian in turn implies that the vertical maps induce weak equivalences of the horizontal homotopy fibers which, by the homotopy cofinality condition (iv) in the definition of a well-structured relative index category, implies the statement of the proposition.
\end{proof}
Recall that homotopy cartesian squares can be conveniently treated in
right proper model categories (see e.g.\ ~\cite[II.\S
8]{Goerss_J-simplicial} for details). The proof of the previous
proposition can now be reinterpreted to give the following corollary.
\begin{corollary}
A square diagram in ${\mathcal S}^{{\mathcal K}}$ is homotopy cartesian with respect to the
${\mathcal A}$-relative ${\mathcal K}$-model structure if and only if the associated
square of homotopy colimits is homotopy cartesian in ${\mathcal S}$. \qed
\end{corollary}
Consider now an operad ${\mathcal D}$ and the category of algebras ${\mathcal S}^{{\mathcal K}}[\mathbb D]$ for the corresponding monad $\mathbb D$. Since pullbacks in
${\mathcal S}^{{\mathcal K}}[\mathbb D]$ are defined in terms of the underlying ${\mathcal K}$-spaces, right properness of a model structure on ${\mathcal S}^{{\mathcal K}}$ carries over to the lifted model structure on ${\mathcal S}^{{\mathcal K}}[\mathbb D]$ when the latter is defined.
\begin{corollary}
The ${\mathcal A}$-relative ${\mathcal K}$-model structures on $ {\mathcal S}^{{\mathcal K}}[\mathbb D]$ considered in Proposition \ref{prop:structured-lift-proposition} are right proper.\qed
\end{corollary}
\subsection{Left properness}
It is an immediate consequence of the left properness of ${\mathcal S}$ that the ${\mathcal A}$-relative level model structure on ${\mathcal S}^{{\mathcal K}}$ is left proper. By parts (i) and (iv) of Proposition \ref{prop:h-cofibration-properties} the same then holds for the
${\mathcal A}$-relative ${\mathcal K}$-model structure.
\begin{proposition}
The ${\mathcal A}$-relative ${\mathcal K}$-model structure on ${\mathcal S}^{{\mathcal K}}$ is left proper.\qed
\end{proposition}
Next we consider left properness for the category ${\mathcal C}{\mathcal S}^{{\mathcal K}}$ of commutative ${\mathcal K}$-space monoids.
The proof in this case is based on the following lemma where
we use the notation $B\boxtimes _AC$ for the pushout of a diagram of
commutative ${\mathcal K}$-space monoids $B\leftarrow A\to C$.
\begin{lemma}\label{lem:AB-lemma}
Let $({\mathcal K},{\mathcal A})$ be a very well-structured relative index category and let
$A\to B$ be a cofibration in the ${\mathcal A}$-relative ${\mathcal K}$-model structure on ${\mathcal C}{\mathcal S}^{{\mathcal K}}$.
Then the functor $B\boxtimes_A(-)$ preserves
${\mathcal K}$-equivalences of commutative ${\mathcal K}$-space monoids under $A$.
\end{lemma}
\begin{proof}
Let $C\to C'$ be a ${\mathcal K}$-equivalence of commutative $\mathcal
K$-space monoids under $A$. Suppose first that $A\to B$ has the form
$\mathbb C(X)\to \mathbb C(Y)$ for a generating ${\mathcal A}$-relative
cofibration $f\colon X\to Y$ in $\mathcal S^{\mathcal K}$, where as usual
$\mathbb C$ denotes the monad associated to the commutativity operad.
We write
$D$ and $D'$ for the pushouts $\mathbb C(Y)\boxtimes_{\mathbb
C(X)}C$ and $\mathbb C(Y)\boxtimes_{\mathbb C(X)}C'$, and consider
the associated filtration terms $F_iD = F_iU_0^{\mathbb C}(D)$ and
$F_iD'= F_iU_0^{\mathbb C}(D')$ from Proposition
\ref{prop:pushout-filtration}. Since these are filtrations by
$h$-cofibrations of ${\mathcal K}$-spaces, it suffices to prove that
$F_iD\to F_iD'$ is a ${\mathcal K}$-equivalence for all $i$. It
follows from Proposition~\ref{prop:pushout-filtration} that $F_{i+1}D\to
F_{i+1}D'$ can be identified with the map obtained from the diagram
\[
\xymatrix@-1pc{
F_iD \ar[d]& C\boxtimes Q_{i-1}^i(f)/\Sigma_i \ar[l] \ar[r] \ar[d]& C\boxtimes Y^{\boxtimes i}/\Sigma_i \ar[d]\\
F_iD' & C'\boxtimes Q_{i-1}^i(f)/\Sigma_i \ar[l] \ar[r] &
C'\boxtimes Y^{\boxtimes i}/\Sigma_i }
\]
by evaluating the pushouts horizontally. Here we use that by Example
\ref{example:U_k-commutative} the functor $U_i^{\mathbb C}$
takes a commutative ${\mathcal K}$-space monoid to its underlying
${\mathcal K}$-space with trivial $\Sigma_i$-action. Proceeding by
induction, we assume that $F_iD\to F_iD'$ is a
${\mathcal K}$-equivalence and it remains to show that so are the other two
vertical maps. We know from Proposition
\ref{prop:boxtimes-flat-invariance} that these maps are
$\mathcal K$-equivalences before taking $\Sigma_i$-orbits and arguing as in the second half of the proof of Lemma \ref{structured-acyclic-pushout} the
same then holds for the maps of $\Sigma_i$-orbits.
In the general case we may assume that $A\to B$ is a relative
cell complex in the sense that there exists a $\lambda$-sequence
$\{B_{\alpha}\colon \alpha<\lambda\}$ of commutative $\mathcal
K$-space monoids (for some ordinal~$\lambda$) such that $A=B_0$, $B
\cong \colim_{\alpha < \lambda}B_{\alpha}$, and $B_{\alpha}\to
B_{\alpha+1}$ is obtained by cobase change from a map of the form
$\mathbb C(X_{\alpha})\to \mathbb C(Y_{\alpha})$ where $X_{\alpha}\to
Y_{\alpha}$ is a generating ${\mathcal A}$-relative cofibration in
$\mathcal S^{{\mathcal K}}$. Lemma~\ref{structured-cocomplete} and
Proposition~\ref{prop:D-h-cofibration} imply that $B_{\alpha}\boxtimes_A C$ and
$B_{\alpha}\boxtimes_A C'$ are $\lambda$-sequences of $h$-cofibrations
of the underlying ${\mathcal K}$-spaces. It follows from an inductive
argument based on the special case considered in the beginning of the
proof that $B_{\alpha}\boxtimes _AC\to B_{\alpha}\boxtimes_A C'$ is a
${\mathcal K}$-equivalence for each $\alpha$. This implies the statement
of the lemma.
\end{proof}
\begin{proposition}\label{prop:left-properness}
Let $({\mathcal K},{\mathcal A})$ be a very well-structured relative index category. Then the
${\mathcal A}$-relative ${\mathcal K}$-model structure on
$\mathcal C\mathcal S^{{\mathcal K}}$ is left proper.
\end{proposition}
\begin{proof}
Let $A\to B$ be an ${\mathcal A}$-relative cofibration and let $A\to C$ be a
${\mathcal K}$-equivalence, both in $\mathcal C\mathcal S^{{\mathcal K}}$. The
cobase change of $A \to C$ along $A \to B$ can be identified with
the map $B\boxtimes_A A\to B\boxtimes _A C$ and the result therefore
follows from Lemma \ref{lem:AB-lemma}.
\end{proof}
\begin{corollary}\label{cor:proper}
\begin{enumerate}[(i)]
\item
The ${\mathcal A}$-relative ${\mathcal K}$-model structure on ${\mathcal S}^{{\mathcal K}}$ is proper.
\item
If $({\mathcal K},{\mathcal A})$ is very well-structured, then the ${\mathcal A}$-relative ${\mathcal K}$-model structure on ${\mathcal C}{\mathcal S}^{{\mathcal K}}$ is proper.
\qed
\end{enumerate}
\end{corollary}
\begin{remark}
In general we do not know under which conditions on the operad the model structures on ${\mathcal S}^{{\mathcal K}}[\mathbb D]$ are left proper. A proof based on a generalization of Lemma \ref{lem:AB-lemma} would require a more careful analysis of the functors $U_k^{\mathbb D}$.
\end{remark}
\section{Cofibrancy of structured diagram spaces}
\label{sec:structured-cofibrant}
Let again $\mathcal D$ be an operad in ${\mathcal S}$ and let $\mathbb D$ be the associated monad on $\mathcal S^{{\mathcal K}}$. In this section we analyze to what
extent the forgetful functor from $\mathcal S^{{\mathcal K}}[\mathbb D]$
to $\mathcal S^{{\mathcal K}}$ preserves ${\mathcal A}$-relative cofibrancy for a (very)
well-structured relative index category $({\mathcal K},{\mathcal A})$. Suppose first that ${\mathcal D}$ is $\Sigma$-free.
As a motivating example consider for an object $\bld k$ in ${\mathcal A}$
the cofibrant object $\mathbb D(F^{{\mathcal K}}_{\mathbf k}(*))$
in the ${\mathcal A}$-relative ${\mathcal K}$-model structure on $\mathcal
S^{{\mathcal K}}[\mathbb D]$. This has as its underlying $\mathcal
K$-space the coproduct of the ${\mathcal K}$-spaces $\mathcal
D(n)\times_{\Sigma_n}F^{{\mathcal K}}_{\mathbf k^{\sqcup n}}(*)$ for $n\geq 0$. In particular, for $n=0$ this is the free ${\mathcal K}$-space $F_{\bld 0}(*)$. In order for the latter to be cofibrant we introduce the following assumption on our index categories; compare also to Proposition \ref{prop:well-structured-comparison}.
Here and in the following the degree functor on ${\mathcal K}$ is supposed to be fixed.
\begin{coarse-assumptions}\label{assump:coarse}
Let $({\mathcal K},{\mathcal A})$ be a well-structured relative index category
and let $({\mathcal K},{\mathcal B})$ be a well-structured relative index category such that
${\mathcal A}\subseteq {\mathcal B}$ and ${\mathcal B}$ contains the unit $\bld 0$ for the monoidal structure.
\end{coarse-assumptions}
Recall from Section \ref{subs:G-space-model} the notion of the \emph{coarse} model structure associated to the category of $G$-spaces for a discrete group $G$. It follows from the definition of the generating cofibrations that the $G$-action on a cofibrant object is free. In the topological setting a cofibrant object is also Hausdorff (and not just weak Hausdorff).
\begin{proposition}\label{prop:coarse-underlying-cofibrant}
Let $\mathcal D$ be an operad in $\mathcal S$ such that $\mathcal
D(n)$ is cofibrant in the coarse model structure on ${\mathcal S}^{\Sigma_n}$ for all
$n$, and let ${\mathcal K}$, ${\mathcal A}$, and ${\mathcal B}$ be as in Coarse
Assumptions~\ref{assump:coarse}.
Suppose that $A$ is a $\mathbb D$-algebra which is cofibrant in the
${\mathcal A}$-relative ${\mathcal K}$-model structure on $\mathcal S^{{\mathcal K}}[\mathbb
D]$. Then the underlying ${\mathcal K}$-space of $A$ is cofibrant in
the ${\mathcal B}$-relative ${\mathcal K}$-model structure on $\mathcal S^{\mathcal K}$.
\end{proposition}
We prove the Proposition in Section \ref{subs:coarse-underlying} below.
Notice, that in the simplicial setting this proposition applies to all
$\Sigma$-free operads. In the topological setting it applies for
instance to operads $\mathcal D$ such that $\mathcal D(n)$ is a free
$\Sigma_n$-equivariant CW complex.
\begin{corollary}\label{cor:projective-underlying-cofibrant}
Let ${\mathcal K}$ be a well-structured index category and let ${\mathcal D}$ be an operad in
${\mathcal S}$ such that $\mathcal D(n)$ is cofibrant in the coarse model structure on ${\mathcal S}^{\Sigma_n}$ for all $n$. Suppose that $A$ is a $\mathbb D$-algebra which is cofibrant in the projective ${\mathcal K}$-model structure on
${\mathcal S}^{{\mathcal K}}[\mathbb D]$. Then the underlying ${\mathcal K}$-space of $A$ is cofibrant in the projective ${\mathcal K}$-model structure on ${\mathcal S}^{{\mathcal K}}$. \qed
\end{corollary}
For example, this applies to the associativity operad and the corresponding category of (not necessarily commutative) ${\mathcal K}$-space monoids.
Now we drop the assumption that ${\mathcal D}$ be $\Sigma$-free and assume instead that $({\mathcal K},{\mathcal A})$ is very well-structured. By Proposition
\ref{prop:structured-lift-proposition} this ensures that the ${\mathcal A}$-relative
${\mathcal K}$-model structure on ${\mathcal S}^{{\mathcal K}}$ lifts to
${\mathcal S}^{{\mathcal K}}[\mathbb D]$ for any operad $\mathcal D$. However, for the forgetful functor to preserve cofibrancy we need the additional assumption stated below. Recall the canonical homomorphism
$\Sigma_n\ltimes {\mathcal A}(\bld k)^{\times n}\to{\mathcal K}(\bld k^{\sqcup n})$ discussed after Definition~\ref{def:very-well-structured}.
\begin{fine-assumptions}\label{assump:fine}
Let $({\mathcal K},{\mathcal A})$ be a very well-structured relative index category
and let $({\mathcal K},{\mathcal B})$ be a well-structured relative index category such that
${\mathcal A}\subseteq {\mathcal B}$, the group $\Sigma_n\ltimes {\mathcal A}(\bld k)^{\times n}$ maps into ${\mathcal B}(\bld k^{\sqcup n})$ for all objects $\bld k$ in ${\mathcal A}$ and all $n\geq 1$, and ${\mathcal B}$ contains the unit $\bld 0$ for the monoidal structure
\end{fine-assumptions}
Recall from Section \ref{subs:G-space-model} the notion of the \emph{fine} model structure associated to the category of $G$-spaces for a group $G$.
\begin{proposition}\label{prop:fine-underlying-cofibrant}
Let $\mathcal D$ be an operad such that $\mathcal D(n)$ is
cofibrant in the fine model structure on ${\mathcal S}^{\Sigma_n}$ for all $n$, and let
${\mathcal K}$, ${\mathcal A}$, and ${\mathcal B}$ be as in Fine Assumptions \ref{assump:fine}. Suppose that $A$ is a
$\mathbb D$-algebra which is cofibrant in the ${\mathcal A}$-relative ${\mathcal K}$-model structure on $\mathcal S^{{\mathcal K}}[\mathbb D]$. Then the underlying ${\mathcal K}$-space of $A$ is cofibrant in the ${\mathcal B}$-relative ${\mathcal K}$-model structure on ${\mathcal S}^{{\mathcal K}}$.
\end{proposition}
The proof will be given in Section \ref{subs:fine-underlying}.
Notice, that in the simplicial setting this proposition applies to all
operads. In the topological setting it applies for instance to operads
${\mathcal D}$ for which ${\mathcal D}(n)$ is a $\Sigma_n$-equivariant CW
complex.
\begin{corollary}\label{cor:comm-underlying-cofibrant}
Let ${\mathcal K}$, ${\mathcal A}$, and ${\mathcal B}$ be as in Fine Assumptions
\ref{assump:fine}. Suppose that $A$ is a commutative ${\mathcal K}$-space monoid which is cofibrant in the ${\mathcal A}$-relative ${\mathcal K}$-model structure on
$\mathcal C\mathcal S^{{\mathcal K}}$. Then the underlying $\mathcal
K$-space of $A$ is cofibrant in the ${\mathcal B}$-relative ${\mathcal K}$-model
structure on $\mathcal S^{{\mathcal K}}$.\qed
\end{corollary}
The last result about preservation of cofibrancy has no assumptions on
the operads. Instead it uses the notion of an $h$-cofibration from Section
\ref{subs:h-cofibrations}.
\begin{proposition}\label{prop:D-h-cofibration}
Let $\mathcal D$ be an operad and let $A\to B$ be a cofibration in
any of the lifted model structures on ${\mathcal S}^{{\mathcal K}}[\mathbb
D]$ considered in Proposition \ref{prop:structured-lift-proposition}. Then the underlying map of ${\mathcal K}$-spaces is an $h$-cofibration.
\end{proposition}
\begin{proof}
This follows from Lemma \ref{structured-h-cofibration-pushout},
Lemma \ref{structured-cocomplete}, and Proposition
\ref{prop:h-cofibration-properties}, by expressing $A\to B$ as a
retract of a relative cell complex (that is, a transfinite composition of maps
obtained by attaching generating cofibrations).
\end{proof}
\subsection{The proof of Proposition \ref{prop:coarse-underlying-cofibrant}}
\label{subs:coarse-underlying}
In order to prove Proposition \ref{prop:coarse-underlying-cofibrant} we shall
temporarily work in the category of $\Sigma_k$-equivariant $\mathcal K$-spaces for varying $k$ and for this we need an equivariant
version of the ${\mathcal A}$-relative level model structure on
$\mathcal S^{\mathcal K}$. In general, given a finite group $G$, we write
$(\mathcal S^{{\mathcal K}})^G$ for the category of $G$-${\mathcal K}$-spaces, that is,
${\mathcal K}$-spaces with $G$-action. If we view $G$ as a category with a
single object this is the same thing as functors from
$ G\times{\mathcal K}$ to $\mathcal S$ which in turn is the same as ${\mathcal K}$-diagrams in ${\mathcal S}^G$.
Proceeding as in the definition of the \emph{coarse} model structure on
${\mathcal S}^G$ considered in Section \ref{subs:G-space-model} we now define the \emph{$G$-coarse ${\mathcal A}$-relative level model structure} on $({\mathcal S}^{{\mathcal K}})^G$.
In this model structure a map of
$G$-${\mathcal K}$-spaces is a weak equivalence (respectively a fibration) if and only if the underlying map of ${\mathcal K}$-spaces is an ${\mathcal A}$-relative level equivalence (respectively an ${\mathcal A}$-relative level fibration) as defined in Section
\ref{subs:level-model-structures}. Arguing as for the ${\mathcal A}$-relative level model structure on $\mathcal S^{{\mathcal K}}$, one checks that this defines a cofibrantly generated model structure on $(\mathcal S^{{\mathcal K}})^G$. The generating
cofibrations (respectively the generating acyclic cofibrations) are the maps
of the form $G\times X\to G\times Y$ where $X\to Y$ is a generating
cofibration (respectively a generating acyclic cofibration) for the ${\mathcal A}$-relative level model structure on $\mathcal S^{{\mathcal K}}$. We shall use the term
\emph{$G$-coarse ${\mathcal A}$-relative cofibration} for a cofibration in this model structure.
The properties of the cofibrations stated in the next lemma will be needed later. They are easy consequences of the fact that any such cofibration is a retract of a relative cell complex constructed from the generating cofibrations.
\begin{lemma}\label{lem:coarse-cofibration-properties}
Let $H$ and $G$ be finite groups.
\begin{enumerate}[(i)]
\item If $X\to Y$ is an $(H\times G)$-coarse ${\mathcal A}$-relative cofibration, then the map of $G$-orbits $X/G\to Y/G$ is an $H$-coarse ${\mathcal A}$-relative cofibration.
\item If $X\to Y$ is a $G$-coarse ${\mathcal A}$-relative cofibration and $H$ is a subgroup of $G$, then, restricting the action, $X\to Y$ is an
$H$-coarse ${\mathcal A}$-relative cofibration. \qed
\end{enumerate}
\end{lemma}
The proofs of the next two lemmas are based on the following elementary observation: Let $f\colon X\to Y$ be a map of $G$-${\mathcal K}$-spaces and write $f'\colon X'\to Y'$ for the underlying map of ${\mathcal K}$-spaces with trivial $G$-action. Then there is a commutative diagram of $G$-${\mathcal K}$-spaces
\begin{equation}\label{eq:G-observation}
\xymatrix{
G\times X' \ar[r]^{1_G\times f'} \ar[d]^{\cong} & G\times Y'\ar[d]^{\cong}\\
G\times X \ar[r]^{1_G\times f} & G\times Y
}
\end{equation}
where the vertical maps are the $G$-equivariant isomorphisms
$(g,x)\mapsto (g,gx)$.
\begin{lemma}\label{lem:coarse-equivariant-lemma}
Let $G$ be a finite group and suppose that $f$ is a $G$-coarse ${\mathcal A}$-relative cofibration in $({\mathcal S}^{{\mathcal K}})^G$ and that $i$ is a cofibration in the fine model structure on ${\mathcal S}^G$. Then the pushout product $f\Box i$ is again a
$G$-coarse ${\mathcal A}$-relative cofibration in $(\mathcal S^{{\mathcal K}})^G$.
\end{lemma}
\begin{proof}
By~\cite[Lemma 4.2.4]{Hovey-model} it suffices to check this
when $f$ and $i$ are generating cofibrations for the respective
model structures. Thus we may assume that $f$ has the form $1_G\times g$ for a generating ${\mathcal A}$-relative cofibration $g$ in ${\mathcal S}^{{\mathcal K}}$. Then $f\Box i$ can be identified with $1_G\times (g\Box i)$. It follows from Proposition \ref{prop:level-S-model} that, forgetting the equivariant structure, $g\Box i$ is an ${\mathcal A}$-relative cofibration. Now apply the observation above the lemma to get the result.
\end{proof}
As a consequence of the lemma we see that if $U$ is cofibrant in
the $G$-coarse ${\mathcal A}$-relative level model structure on
$(\mathcal S^{{\mathcal K}})^G$ and $K\to L$ is a cofibration in the fine model structure on ${\mathcal S}^G$, then the induced map $U\times K\to U\times L$ is a
$G$-coarse ${\mathcal A}$-relative cofibration (since $\emptyset \to U$ is a $G$-coarse ${\mathcal A}$-relative cofibration).
Just as in the non-equivariant setting, the category $(\mathcal
S^{{\mathcal K}})^G$ is closed symmetric monoidal with monoidal product
$X\boxtimes Y$ defined as the usual Kan extension along the monoidal
structure map ${\mathcal K}\times {\mathcal K}\to {\mathcal K}$. Thus, the underlying ${\mathcal K}$-space
of $X\boxtimes Y$ is the $\boxtimes$-product of the underlying
${\mathcal K}$-spaces of $X$ and $Y$.
\begin{lemma}\label{lem:UX-coarse-cofibrant}
Let $G$ be a finite group. If $U$ is a $G$-${\mathcal K}$-space which
is cofibrant in the $G$-coarse ${\mathcal A}$-relative level model structure on
$(\mathcal S^{{\mathcal K}})^G$ and $X$ is a $G$-${\mathcal K}$-space whose
underlying ${\mathcal K}$-space is ${\mathcal A}$-relative cofibrant, then $U\boxtimes X$ is
again cofibrant in the $G$-coarse ${\mathcal A}$-relative level model structure on
$(\mathcal S^{{\mathcal K}})^G$.
\end{lemma}
\begin{proof}
We may assume without loss of generality that $U$ is a $G$-coarse cell complex in the sense that it can be identified
with the colimit of a $\lambda$-sequence of $G$-${\mathcal K}$-spaces
$\{U_{\alpha}\colon\alpha<\lambda\}$ (for some ordinal~$\lambda$)
such that $U_0=\emptyset$ and $U_{\alpha}\to U_{\alpha+1}$ is
obtained by cobase change of a generating cofibration $G\times
X_{\alpha}\to G\times Y_{\alpha}$ (where $X_{\alpha}\to Y_{\alpha}$
is a generating ${\mathcal A}$-relative cofibration in $\mathcal S^{\mathcal
K}$). It follows that $U\boxtimes X$ can be identified with the
colimit of the $\lambda$-sequence $\{U_{\alpha}\boxtimes
X\colon\alpha<\lambda\}$ and it suffices to show that each of the
maps $U_{\alpha}\boxtimes X\to U_{\alpha+1}\boxtimes X$ is a
$G$-coarse ${\mathcal A}$-relative cofibration. This map is obtained by cobase change from the map $G\times X_{\alpha}\boxtimes X\to G\times Y_{\alpha}\boxtimes X$ and the conclusion now follows from the observation before Lemma \ref{lem:coarse-equivariant-lemma} (letting $X_{\alpha}\boxtimes X\to Y_{\alpha}\boxtimes X$ be the map $f$ in \eqref{eq:G-observation}).
\end{proof}
We now turn to the proof of Proposition \ref{prop:coarse-underlying-cofibrant}
and assume that ${\mathcal A}$ and ${\mathcal B}$ are as in Coarse Assumptions \ref{assump:coarse}. Then the above lemmas apply as well to the well-structured relative index category $({\mathcal K},{\mathcal B})$.
The key step is again to analyze pushout diagrams of the form
\eqref{verification-structured-pushout}.
\begin{lemma}\label{lem:structured-coarse-pushout-lemma}
Let $\mathcal D$ be an operad in $\mathcal S$ such that $\mathcal
D(n)$ is cofibrant in the coarse model structure on ${\mathcal S}^{\Sigma_n}$ for all $n$. Suppose that the map $f$ in \eqref{verification-structured-pushout} is a generating ${\mathcal A}$-relative cofibration and that $U_k^{\mathbb D}(A)$ is cofibrant in the $\Sigma_k$-coarse ${\mathcal B}$-relative level model structure on
$(\mathcal S^{\mathcal K})^{\Sigma_k}$ for all $k\geq 0$. Then the induced map
\[
U_k^{\mathbb D}(\bar f)\colon U_k^{\mathbb D}(A)\to U^{\mathbb
D}_k(B)
\]
is a $\Sigma_k$-coarse ${\mathcal B}$-relative cofibration for all $k\geq 0$.
\end{lemma}
\begin{proof}
Applying the filtration from Proposition \ref{prop:pushout-filtration} it
suffices to show that
\begin{equation}\label{eq:coarse-map}
U_{i+k}^{\mathbb D}(A)\boxtimes_{\Sigma_i}Q_{i-1}^i(f) \to
U_{i+k}^{\mathbb D}(A)\boxtimes_{\Sigma_i}Y^{\boxtimes i}
\end{equation}
is a $\Sigma_k$-coarse ${\mathcal B}$-relative cofibration for all $i$ and $k$. By definition, the generating cofibration $f$ has the form
$G_{\bld k}^{{\mathcal K}}({\mathcal K}(\bld k)/H\times h)$ for an object $\bld k$ in ${\mathcal A}$, a subgroup $H\subseteq {\mathcal A}(\bld k)$, and a generating cofibration $h$ in ${\mathcal S}$. Before passing to $\Sigma_i$-orbits, the map \eqref{eq:coarse-map} can therefore be identified with the map
\[
U_{i+k}^{\mathbb D}(A)\boxtimes G_{\bld k^{\sqcup i}}^{{\mathcal K}}
({\mathcal K}(\bld k^{\sqcup i})/H^{\times i})\times h^{\Box i}
\]
where $h^{\Box i}$ is the iterated pushout-product in ${\mathcal S}$. We can view this as a map of $(\Sigma_i\times \Sigma_k)$-${\mathcal K}$-spaces by restricting the
$\Sigma_{i+k}$-action on $U_{i+k}^{\mathbb D}(A)$ and extending the obvious
$\Sigma_i$-actions on the two other factors to $(\Sigma_i\times \Sigma_k)$-actions by letting $\Sigma_k$ act trivially. Using Lemma \ref{lem:coarse-cofibration-properties}(ii) (with ${\mathcal B}$ instead of ${\mathcal A}$) we see that the assumptions on
$U_{i+k}^{\mathbb D}(A)$ imply that the latter is cofibrant in the
$(\Sigma_i\times \Sigma_k)$-coarse ${\mathcal B}$-relative model structure. Hence the $\boxtimes$-product of the first two factors is also cofibrant in this model structure by Lemma \ref{lem:UX-coarse-cofibrant}. The argument given in the proof of Proposition \ref{prop:h-cofibration-properties}(vii) implies that
$h^{\Box i}$ is a cofibration in the fine model structure on ${\mathcal S}^{\Sigma_i}$.
Combined with Lemma \ref{lem:coarse-equivariant-lemma}
this in turn implies that the above map is a
$(\Sigma_i\times\Sigma_k)$-coarse ${\mathcal B}$-relative cofibration. Finally,
Lemma \ref{lem:coarse-cofibration-properties}(i) then gives that the induced map of $\Sigma_i$-orbits is a $\Sigma_k$-coarse ${\mathcal B}$-relative cofibration.
\end{proof}
\begin{proof}[Proof of Proposition \ref{prop:coarse-underlying-cofibrant}]
We may assume without loss of generality that $A$ is a cell complex
in the ${\mathcal A}$-relative ${\mathcal K}$-model structure on $\mathcal
S^{{\mathcal K}}[\mathbb D]$. Thus, $A$ may be identified with the colimit of
a $\lambda$-sequence of $\mathbb D$-algebras
$\{A_{\alpha}\colon\alpha<\lambda\}$ (for some ordinal~$\lambda$)
such that $A_0=\bld 1_{{\mathcal K}}$ and $A_{\alpha}\to A_{\alpha+1}$ is
obtained by cobase change of a generating cofibration $\mathbb
D(X_{\alpha})\to \mathbb D(Y_{\alpha})$ where $X_{\alpha}\to
Y_{\alpha}$ is a generating ${\mathcal A}$-relative cofibration in $\mathcal
S^{{\mathcal K}}$. Since the underlying ${\mathcal K}$-space of the unit $\bld
1_{{\mathcal K}}$ can be identified with $F^{{\mathcal K}}_{\mathbf 0}(*)$, it
suffices by Lemma \ref{structured-cocomplete} to show that each of
the morphisms $A_{\alpha}\to A_{\alpha+1}$ defines a ${\mathcal B}$-relative
cofibration in $\mathcal S^{{\mathcal K}}$. In order to set up an inductive
argument based on Lemma \ref{lem:structured-coarse-pushout-lemma} we
in fact prove the stronger statement that (i) $U_k^{\mathbb
D}(A_{\alpha})$ is cofibrant in the $\Sigma_k$-coarse
${\mathcal B}$-relative level model structure on $(\mathcal
S^{{\mathcal K}})^{\Sigma_k}$ for all $k$, and (ii) $U_k^{\mathbb
D}(A_{\alpha})\to U_k^{\mathbb D}(A_{\alpha+1})$ is a
$\Sigma_k$-coarse ${\mathcal B}$-relative cofibration for all $k$. Letting
$k=0$ then gives the result. In order to start the induction we
observe that by Proposition~\ref{prop:pushout-filtration}(iii)
(identifying $\bld 1_{{\mathcal K}}$ with $\mathbb D(\emptyset)$), the
underlying $\mathcal K$-space of $U_k^{\mathbb D}(\bld 1_{{\mathcal K}})$ is
isomorphic to $F^{{\mathcal K}}_{\mathbf 0}(\mathcal D(k))$ and therefore
cofibrant in the $\Sigma_k$-coarse ${\mathcal B}$-relative model structure on
$(\mathcal S^{{\mathcal K}})^{\Sigma_k}$ by the assumption on $\mathcal D$.
Proceeding by induction we consider an ordinal $\beta$ with
$\beta+1<\lambda$ such that (i) and (ii) hold for all
$\alpha<\beta$. If $\beta$ is a successor ordinal, $\beta=\alpha+1$,
it is immediately clear from Lemma
\ref{lem:structured-coarse-pushout-lemma} that (i) and (ii) also
hold for $\beta$. If $\beta$ is a limit ordinal, then it follows
from the definition of a $\lambda$-sequence and the fact that the
functor $U^{\mathbb D}_k$ preserves filtered colimits that
$U_k^{\mathbb D}(A_{\beta})$ can be identified with
$\colim_{\alpha<\beta} U_k^{\mathbb D}(A_{\alpha})$. By the
induction hypothesis $\bld 1_{{\mathcal K}}\to U_k^{\mathbb D}(A_{\beta})$ is
therefore a transfinite composition of $\Sigma_k$-coarse
${\mathcal B}$-relative cofibrations, hence itself a $\Sigma_k$-coarse
${\mathcal B}$-relative cofibration which implies that (i) holds for
$\beta$. By Lemma~\ref{lem:structured-coarse-pushout-lemma}, (ii)
then also holds for $\beta$.
\end{proof}
\subsection{The proof of Proposition \ref{prop:fine-underlying-cofibrant} }\label{subs:fine-underlying}
In this section $({\mathcal K},{\mathcal A})$ denotes a very well-structured relative index category. We begin by defining, for a finite group $G$, a cofibrantly generated \emph{$(G\times {\mathcal A})$-relative level model structure} on $(\mathcal S^{{\mathcal K}})^G$: The weak equivalences in this model structure are the
${\mathcal A}$-relative level equivalences of the underlying ${\mathcal K}$-spaces and the fibrations are defined as in Section \ref{subs:level-model} with $H$ now being a subgroup of $G \times{\mathcal A}(\bld{k})$. In order to see that this defines a cofibrantly generated model structure we notice that for each object $\bld k$ in ${\mathcal A}$ the functor
\[
G_{\bld k}^{{\mathcal K}}\colon {\mathcal S}^{G\times {\mathcal K}(\bld k)}\to {\mathcal S}^{G\times {\mathcal K}},\quad
G_{\bld k}^{{\mathcal K}}(L)={\mathcal K}(\bld k,-)\times_{{\mathcal K}(\bld k)}L
\]
is left adjoint to the evaluation functor $\Ev_{\bld k}$. We obtain the generating (acyclic) cofibrations by applying the functors
$G_{\bld k}^{{\mathcal K}}$ to the generating (acyclic) cofibrations for the
$(G\times {\mathcal A}(\bld k))$-relative mixed model structure on $\mathcal
S^{G\times {\mathcal K}(\bld{k})}$, cf.\ Section
\ref{subs:G-space-model}. We shall use the term \emph{$(G\times{\mathcal A})$-relative cofibration} for a cofibration in this model structure.
As in the non-equivariant case there is a more explicit description of the
$(G\times {\mathcal A})$-relative cofibrations. Given a $G$-${\mathcal K}$-space $X$ and an object $\bld k$ in ${\mathcal K}$, the latching space $L_{\bld k}(X)$ of the underlying ${\mathcal K}$-space is defined as in Section \ref{subs:level-model}. Recall the notion of a relative equivariant cofibration introduced in Section
\ref{subs:G-space-model}. The arguments in the proof of Proposition
\ref{prop:latching-characterization} easily generalizes to give the next result.
\begin{proposition}\label{prop:G-latching}
A map of $G$-${\mathcal K}$-spaces $X\to Y$ is a $(G\times{\mathcal A})$-relative cofibration if and only if the $(G\times {\mathcal K}(\bld k))$-equivariant latching map
\[
L_{\bld{k}}(Y) \cup_{L_{\bld{k}}(X)}X(\bld{k}) \to Y(\bld{k})
\]
is a $(G\times{\mathcal A}(\bld k))$-relative cofibration for all objects $\bld k$ in ${\mathcal A}$ and an isomorphism for all objects $\bld k$ not in ${\mathcal A}$.\qed
\end{proposition}
Using this characterization it is easy to see that the
$(G\times {\mathcal A})$-relative analogues of the statements in Lemma
\ref{lem:coarse-cofibration-properties} hold. In the following we deduce some further properties of the $(G\times {\mathcal A})$-relative cofibrations needed for the proof of
Proposition~\ref{prop:fine-underlying-cofibrant}. For this it is useful to observe that in general, given a discrete group $H$ and a normal subgroup $N$, a discrete $H$-space is cofibrant in the $N$-relative model structure on ${\mathcal S}^H$ if and only if each of the isotropy subgroups of $H$ is contained in $N$.
\begin{proposition}\label{G-pushout-product-axiom}
The pushout-product axiom holds for the $(G\times {\mathcal A})$-relative level model structure on $(\mathcal S^{{\mathcal K}})^G$.
\end{proposition}
\begin{proof}
By \cite[Lemma 3.5]{Schwede_S-algebras} it suffices to consider the generating (acyclic) cofibrations. Let
$\bld k_1$ and $\bld k_2$ be objects in ${\mathcal A}$, and let $h_s$ be a generating cofibration for the ($G\times {\mathcal A}(\bld k_s)$)-relative model structure on $S^{G\times {\mathcal K}(\bld k_s)}$ for $s=1,2$. Then we have the identification
\[
G_{\bld k_1}^{{\mathcal K}}(h_1)\Box G_{\bld k_2}^{{\mathcal K}}(h_2)
\cong G^{{\mathcal K}}_{\bld k_1\sqcup \bld k_2}({\mathcal K}(\bld k_1\sqcup \bld k_2)
\times _{{\mathcal K}(\bld k_1)\times {\mathcal K}(\bld k_2)}h_1\Box h_2)
\]
and we must check that in the last expression we apply
$G_{\bld k_1\sqcup \bld k_2}^{{\mathcal K}}$ to a ($G\times{\mathcal A}(\bld k_1\sqcup \bld k_2)$)-relative cofibration. Writing $h_s=(G\times {\mathcal K}(\bld k_s))/H\times i_s$ for
$H_s\subseteq G\times {\mathcal A}(\bld k_s)$ and $i_s$ a generating cofibration in
${\mathcal S}$, we get
\[
{\mathcal K}(\bld k_1\sqcup \bld k_2)
\times _{{\mathcal K}(\bld k_1)\times {\mathcal K}(\bld k_2)}h_1\Box h_2\cong
(G\times G\times K(\bld k_1\sqcup \bld k_2))/(H_1\times H_2)
\times i_1\Box i_2.
\]
Now apply the ($G\times {\mathcal K}(\bld k_1\sqcup \bld
k_2)$)-equivariant projection
\[
(G\times G\times K(\bld k_1\sqcup \bld k_2))/(H_1\times H_2)\to
K(\bld k_1\sqcup \bld k_2)/{\mathcal A}(\bld k_1\sqcup \bld k_2)
\]
to conclude that the isotropy groups of the elements in the domain are contained in
$G\times {\mathcal A}(\bld k_1\sqcup\bld k_2)$. This gives the result for the generating cofibrations.
For the second part of the pushout-product axiom (concerning the acyclic cofibrations) we first observe that a ($G\times{\mathcal A})$-relative cofibration defines an ${\mathcal A}$-relative cofibration of the underlying ${\mathcal K}$-spaces. This can for instance be deduced from the ($G\times {\mathcal A}$)-relative version of Lemma
\ref{lem:coarse-cofibration-properties}(ii).
Since the weak equivalences are defined in terms of the underlying
${\mathcal K}$-spaces the conclusion now follows from the pushout-product axiom for the ${\mathcal A}$-relative level model structure, Proposition
\ref{prop:K-pushout-product-axiom}.
\end{proof}
\begin{lemma}\label{lem:iterated-pushout-product-lemma}
Let $({\mathcal K},{\mathcal A})$ and $({\mathcal K},{\mathcal B})$ be as in Fine Assumptions \ref{assump:fine}, and suppose that $f\colon X\to Y$ is a generating ${\mathcal A}$-relative
cofibration in ${\mathcal S}^{{\mathcal K}}$. Then the $i$-fold iterated pushout-product
$
f^{\Box i}\colon Q_{i-1}^i(f)\to Y^{\boxtimes i}
$
is a $(\Sigma_i\times {\mathcal B})$-relative cofibration
\end{lemma}
\begin{proof}
As usual, $f=G_{\bld k}^{{\mathcal K}}({\mathcal K}(\bld k)/H\times h)$ for an object $\bld k$ in
${\mathcal A}$, a subgroup $H\subseteq {\mathcal A}(\bld k)$, and a generating cofibration $h$ for ${\mathcal S}$. Then $f^{\Box i}$ can be identified with the map
\[
G_{\bld k^{\sqcup i}}^{{\mathcal K}}({\mathcal K}(\bld k^{\sqcup i})/H^{\times i}\times h^{\Box i})=
{\mathcal K}(\bld k^{\sqcup i},-)/H^{\times i}\times h^{\Box i}
\]
where $h^{\Box i}$ is the iterated pushout-product in ${\mathcal S}$. Thus, it suffices to show that the $\Sigma_i$-${\mathcal K}$-space $Z={\mathcal K}(\bld k^{\sqcup i},-)/H^{\times i}$ is $(\Sigma_i\times {\mathcal B})$-relative cofibrant (where $\Sigma_i$ acts via the canonical right $\Sigma_i$-action on $\bld k^{\sqcup i}$ specified by the symmetric monoidal structure). By
Proposition~\ref{prop:G-latching} this is equivalent to the latching map $L_{\bld l}(Z)\to Z(\bld l)$ being a $(\Sigma_i\times {\mathcal B}(\bld l))$-relative cofibration for all objects $\bld l$ in ${\mathcal B}$ and an isomorphism for all $\bld l$ not in ${\mathcal B}$. It is clear from the definition that the latching map is an isomorphism if $\bld l$ and $\bld k^{\sqcup i}$ have different degrees, $\lambda(\bld l)\neq \lambda(\bld k^{\sqcup i})$. If $\lambda(\bld l)=\lambda(\bld k^{\sqcup i})$, then $L_{\bld l}(Z)=\emptyset$ and we claim that $Z(\bld l)$ is $(\Sigma_i\times {\mathcal B}(\bld l))$-relative cofibrant. This is clear if $Z(\bld l)$ is empty and otherwise there is an isomorphism $\bld l\to \bld k^{\sqcup i}$ in ${\mathcal K}$ so that it suffices to consider the case $\bld l=\bld k^{\sqcup i}$ (this uses the normality assumption on ${\mathcal B}$). Hence it only remains to show that the elements of $Z(\bld k^{\sqcup i})$ have isotropy groups contained in $\Sigma_i\times \mathcal B(\bld k^{\sqcup i})$. Let $z$ be an element in $Z(\bld k^{\sqcup i})$ represented by a morphism $\beta$ in ${\mathcal K}(\bld k^{\sqcup i})$. The condition for an element $(\sigma,\alpha)$ in
$\Sigma_i\times {\mathcal K}(\bld k^{\sqcup i})$ to belong to the isotropy group of $z$ is that there are elements $f_1\,\dots, f_i$ in $H$ such that
$\alpha\circ\beta=\beta \circ\sigma_*\circ(f_1\sqcup\dots\sqcup f_i)$. Since
${\mathcal B}(\bld k^{\sqcup i})$ is a normal subgroup of ${\mathcal K}(\bld k^{\sqcup i})$ and
$\sigma_*\circ(f_1\sqcup\dots\sqcup f_i)$ belongs to ${\mathcal B}(\bld k^{\sqcup i})$, we conclude that $\alpha$ also belongs to ${\mathcal B}(\bld k^{\sqcup i})$.
\end{proof}
Using the above lemmas we get an analogue of Lemma
\ref{lem:structured-coarse-pushout-lemma} in the current setting.
\begin{lemma}\label{lem:structured-fine-pushout-lemma}
Let $\mathcal D$ be an operad in $\mathcal S$ such that $\mathcal
D(n)$ is cofibrant in the fine model structure on ${\mathcal S}^{\Sigma_n}$
for all $n$. Suppose that the map $f$ in
\eqref{verification-structured-pushout} is a generating ${\mathcal A}$-relative
cofibration and that $U_k^{\mathbb D}(A)$ is cofibrant in the
($\Sigma_k\times {\mathcal B}$)-relative model structure on
$(\mathcal S^{{\mathcal K}})^{\Sigma_k}$ for all $k\geq 0$. Then the induced map
\[
U_k^{\mathbb D}(\bar f)\colon U_k^{\mathbb D}(A)\to U^{\mathbb
D}_k(B)
\]
is a ($\Sigma_k\times {\mathcal B}$)-relative cofibration for all $k\geq 0$.
\end{lemma}
\begin{proof}
Applying the filtration from Proposition \ref{prop:pushout-filtration} it
suffices to show that
\begin{equation}\label{equivariant-flat-equation}
U_{i+k}^{\mathbb D}(A)\boxtimes_{\Sigma_i}Q_{i-1}^i(f) \to
U_{i+k}^{\mathbb D}(A)\boxtimes_{\Sigma_i}Y^{\boxtimes i}
\end{equation}
is a $(\Sigma_k\times {\mathcal B})$-relative cofibration for all $i$ and $k$. We know from Lemma \ref{lem:iterated-pushout-product-lemma} that the iterated pushout-product
$Q_{i-1}^i(f)\to Y^{\boxtimes i}$ is a $(\Sigma_i\times {\mathcal B})$-relative cofibration. Letting $\Sigma_k$ act trivially we may view this map as a
$(\Sigma_i\times\Sigma_k\times {\mathcal B})$-relative cofibration.
Since $U_{i+k}^{\mathbb D}(A)$ is $(\Sigma_{i+k}\times {\mathcal B})$-relative cofibrant by assumption, the relative version of
Lemma \ref{lem:coarse-cofibration-properties}(ii) implies that it is also
$(\Sigma_i\times \Sigma_k\times {\mathcal B})$-relative cofibrant. By the pushout-product axiom for the latter model structure, Proposition \ref{G-pushout-product-axiom}, it follows that (\ref{equivariant-flat-equation}) is a
$(\Sigma_i\times\Sigma_k\times {\mathcal B})$-relative cofibration before taking
$\Sigma_i$-orbits and by the relative version of Lemma \ref{lem:coarse-cofibration-properties}(i) the map of
$\Sigma_i$-orbits is therefore a ($\Sigma_k\times {\mathcal B}$)-relative cofibration as
claimed.
\end{proof}
\begin{proof}[Proof of Proposition \ref{prop:fine-underlying-cofibrant}]
Using Lemma~\ref{lem:structured-fine-pushout-lemma} instead of
Lemma~\ref{lem:structured-coarse-pushout-lemma}, the proof of the proposition is now completely analogous to the proof of
Proposition~\ref{prop:coarse-underlying-cofibrant}
\end{proof}
\section{Diagram spaces and graded spaces}
In this section ${\mathcal K}$ denotes a well-structured index category as specified in Definition~\ref{def:well-structured-index}. We shall consider the corresponding projective ${\mathcal K}$-model structure on ${\mathcal S}^{{\mathcal K}}$ from Definition~\ref{def:K-projective-model}.
By a \emph{space graded over the classifying space $B{\mathcal K}$}
we understand a space $X$ together with a map $X\to B{\mathcal K}$. The purpose of this section is to relate the category of (structured) ${\mathcal K}$-spaces to the category of (structured) spaces graded over $B{\mathcal K}$.
\subsection{Diagram spaces and graded spaces}
Let ${\mathcal K}$ be a well-structured index category with classifying space
$B{\mathcal K}$. We write ${\mathcal S}/B{\mathcal K}$ for the category of spaces over $B{\mathcal K}$, equipped with the standard model structure in which a map is a weak equivalence, fibration, or cofibration, if and only if the underlying map in ${\mathcal S}$ is, see \cite[Theorem 7.6.4]{Hirschhorn-model}.
\begin{theorem}\label{thm:K-spaces-over-BK}
There is a chain of Quillen equivalences relating the projective
${\mathcal K}$-model structure on ${\mathcal S}^{{\mathcal K}}$ to the standard model structure on
${\mathcal S}/B{\mathcal K}$.
\end{theorem}
We first describe the chain of adjunctions in the theorem. Given an
object $\bld{k}$ in ${\mathcal K}$, we write $({\mathcal K}\downarrow \bld{k})$ for the
comma category of objects over $\mathbf k$, and we write $E{\mathcal K}$ for the
${\mathcal K}$-space $B({\mathcal K}\downarrow -)$. There is a pair of adjoint functors
\begin{equation}\label{eq:EK-adjunction}
\xymatrix{
{\mathcal S}^{{\mathcal K}}/E{\mathcal K} \ar@<0.5ex>[r] &
{\mathcal S}^{{\mathcal K}} \ar@<0.5ex>[l]
}
\end{equation}
induced by composition with and pullback along $E{\mathcal K} \to *$. It is
immediate from the definitions that this is a Quillen adjunction. The
fact that $E{\mathcal K}$ is levelwise contractible combined with the right properness of ${\mathcal S}$ has the following implication.
\begin{lemma}\label{lem:EK-Q-equivalence}
The adjunction \eqref{eq:EK-adjunction} is a Quillen equivalence
with respect to the projective ${\mathcal K}$-model structure
on ${\mathcal S}^{{\mathcal K}}$.\qed
\end{lemma}
The obvious forgetful functors $({\mathcal K}\downarrow \mathbf k)\to{\mathcal K}$ give rise to
a map of ${\mathcal K}$-spaces $\pi\colon E{\mathcal K}\to B{\mathcal K}$ (here we write $B{\mathcal K}$ for the constant ${\mathcal K}$-space $\const_{{\mathcal K}}B{\mathcal K}$). Evaluating the colimit over ${\mathcal K}$, this induces an isomorphism of spaces $\colim_{{\mathcal K}}E{\mathcal K}\cong B{\mathcal K}$.
There is a chain of adjoint functors
\begin{equation}\label{eq:EKBKadjunction}
\xymatrix{
\colim_{{\mathcal K}}\colon {\mathcal S}^{{\mathcal K}}/E{\mathcal K} \ar@<0.5ex>[r] &
{\mathcal S}^{{\mathcal K}}/B{\mathcal K} \ar@<0.5ex>[l]
\ar@<0.5ex>[r] & {\mathcal S}/B{\mathcal K} \ar@<0.5ex>[l]
\thinspace \thinspace\colon\! \pi^*
}
\end{equation}
where in the first adjunction the left adjoint is defined by composing
with $\pi$ and the right adjoint takes a ${\mathcal K}$-space $Y$ over $B{\mathcal K}$
to the pullback $E{\mathcal K}\times_{B{\mathcal K}}Y$. The second adjunction is induced
by the usual $(\colim_{{\mathcal K}},\const_{{\mathcal K}})$ adjunction relating ${\mathcal S}^{{\mathcal K}}$
and ${\mathcal S}$. It is clear that the left adjoint in the first adjunction preserves (acyclic) cofibrations and that the right adjoint of the second adjunction preserves (acyclic) fibrations. Thus, these are both Quillen adjunctions and the same therefore holds for their composition.
\begin{lemma}\label{lem:EKBK-Q-equivalence}
The adjunction $(\colim_{{\mathcal K}},\pi^*)$ in \eqref{eq:EKBKadjunction}
is a Quillen equivalence with respect to the projective
${\mathcal K}$-model structure on ${\mathcal S}^{{\mathcal K}}/E{\mathcal K}$ and the standard model structure on ${\mathcal S}/B{\mathcal K}$.
\end{lemma}
\begin{proof}
We use the criterion of~\cite[Corollary 1.3.16]{Hovey-model} and must check that the left adjoint reflects weak equivalences between cofibrant objects and that the derived counit of the adjunction is a weak equivalence. Here the first condition is an immediate consequence of
Lemma~\ref{lem:projective-hocolim}.
That the derived counit is an equivalence means that the composition in the upper row of the diagram
\[
\xymatrix@-1pc{
\colim_{{\mathcal K}}(E{\mathcal K}\times_{B{\mathcal K}}Y)^{\textrm{cof}}
\ar[r] &\colim_{{\mathcal K}}(E{\mathcal K}\times_{B{\mathcal K}}Y) \ar[r]& Y\\
\hocolim_{{\mathcal K}}(E{\mathcal K}\times_{B{\mathcal K}}Y)^{\textrm{cof}}
\ar[r]^-{\simeq} \ar[u]_{\simeq}&\hocolim_{{\mathcal K}}(E{\mathcal K}\times_{B{\mathcal K}}Y) \ar[r]\ar[u]&
\hocolim_{{\mathcal K}}Y\ar[u]
}
\]
is a weak equivalence of spaces over $B{\mathcal K}$ for any fibration $Y\to B{\mathcal K}$ in ${\mathcal S}$. Here $(-)^{\textrm{cof}}$ denotes cofibrant replacement in the projective ${\mathcal K}$-model structure and we again write $B{\mathcal K}$ and $Y$ for the corresponding constant ${\mathcal K}$-spaces. Using Lemma \ref{lem:projective-hocolim} again, we see that it suffices to show that the composition in the upper row of the diagram
\[
\xymatrix@-1pc{
\hocolim_{{\mathcal K}}E{\mathcal K}\times _{B{\mathcal K}}Y \ar[r] \ar[d]& \hocolim_{{\mathcal K}}Y
\ar[r]^-{\textrm{pr}} \ar[d] & \ar[d] Y\\
\hocolim_{{\mathcal K}}E{\mathcal K} \ar[r] & \hocolim_{{\mathcal K}}B{\mathcal K} \ar[r]^-{\textrm{pr}} & B{\mathcal K}
}
\]
is a weak equivalence. We know from Lemma \ref{lem:hocolim-pullback}
that homotopy
colimits preserve pullbacks; hence the left hand square is a pullback
diagram. Since the right hand square is clearly a pullback diagram it
follows that the outer square is a pullback diagram as well. By right
properness of the model structure on ${\mathcal S}$ we have thereby reduced the
problem to showing that the composite map in the bottom row of the
diagram is a weak homotopy equivalence. This composition can be
identified with the canonical map
\[
\hocolim_{{\mathcal K}}E{\mathcal K}\to \colim_{{\mathcal K}}E{\mathcal K}=B{\mathcal K}
\]
and the result follows from Lemma \ref{lem:projective-hocolim} and
the fact that $E{\mathcal K}$ is cofibrant by
Proposition~\ref{prop:latching-characterization}
(see also~\cite[Proposition 14.8.9]{Hirschhorn-model}).
\end{proof}
\noindent\textit{Proof of Theorem \ref{thm:K-spaces-over-BK}.}
Combining Lemmas \ref{lem:EK-Q-equivalence} and \ref{lem:EKBK-Q-equivalence} we get the required chain of Quillen
equivalences
\[
\xymatrix{
{\mathcal S}^{{\mathcal K}} \ar@<-0.5ex>[r] &
{\mathcal S}^{{\mathcal K}}/E{\mathcal K} \ar@<-0.5ex>[l]
\ar@<0.5ex>[r]^-{\colim_{{\mathcal K}}} & {\mathcal S}/B{\mathcal K} \ar@<0.5ex>[l]^-{\pi^*}
}
\eqno\qed
\]
\subsection{Structured diagram spaces and graded spaces}
\label{subs:structured-comma}
We begin with some elementary remarks on adjoint functors between categories of algebras. Consider in general a pair of symmetric monoidal categories $({\mathcal A},\otimes,1_{{\mathcal A}})$ and $({\mathcal B},\Box,1_{{\mathcal B}})$, and
a strong symmetric monoidal functor $V\colon {\mathcal A}\to{\mathcal B}$. Suppose that
$V$ is the left adjoint of an adjoint functor pair $V\colon{\mathcal A}\rightleftarrows {\mathcal B} \thinspace\colon\! U$. Then the right adjoint $U$ inherits the structure of a (lax) symmetric monoidal functor with monoidal structure maps
\[
U(B)\otimes U(B')\to UV(U(B)\otimes U(B'))\xleftarrow{\cong} U(VU(B)\Box VU(B'))
\to U(B\Box B').
\]
Here the first and last arrows are induced respectively by the unit and counit of the adjunction. Similarly, the monoidal unit of $U$ is inherited from the monoidal unit of V using the counit of the adjunction,
\[
1_{{\mathcal B}}\to UV(1_{{\mathcal B}})\xleftarrow{\cong}U(1_{{\mathcal A}}).
\]
With this definition it is clear that the unit and counit of the adjunction are monoidal natural transformations.
Now let ${\mathcal D}$ be an operad in ${\mathcal S}$ and suppose we are given a strong symmetric monoidal functor
$F\colon({\mathcal S},\times,*)\to({\mathcal A},\otimes 1_{{\mathcal A}})$. As we explain in Appendix
\ref{app:analysis-operad-algebras}, the operad ${\mathcal D}$ then gives rise to a monad $\mathbb D$ on ${\mathcal A}$. We write ${\mathcal A}[\mathbb D]$ for the category of $\mathbb D$-algebras in ${\mathcal A}$. Using the strong symmetric monoidal composition $V\circ F$ we also get a monad $\mathbb D$ on ${\mathcal B}$ with a corresponding category of algebras ${\mathcal B}[\mathbb D]$. The symmetric monoidal structure of $V$ gives rise to natural transformations $V\to \mathbb DV$ and $\mathbb DV\to V\mathbb D$, and consequently $V$ takes
$\mathbb D$-algebras in ${\mathcal A}$ to $\mathbb D$-algebras in ${\mathcal B}$. Similarly, using the unit of the adjunction, the (lax) monoidal structure of $U$ gives rise to natural transformations $U\to \mathbb D U$ and $\mathbb D U\to U\mathbb D $. Using this, $V$ and $U$ lift to an adjoint pair of functors
\[
\xymatrix{
V\colon {\mathcal A}[\mathbb D] \ar@<0.5ex>[r] &
{\mathcal B}[\mathbb D] \ar@<0.5ex>[l] \thinspace\colon \! U.
}
\]
Now we specialize to the adjunction
\[
\xymatrix{
\colim_{{\mathcal K}} \colon {\mathcal S}^{{\mathcal K}} \ar@<0.5ex>[r] &
{\mathcal S} \thinspace \colon \! \! \const_{{\mathcal K}} \ar@<0.5ex>[l]
}
\]
where ${\mathcal K}$ denotes a well-structured index category. Let
$\mathcal D$ be an operad in ${\mathcal S}$ and let us again write $\mathbb D$ both for the associated monad on ${\mathcal S}$ and the associated monad on ${\mathcal S}^{{\mathcal K}}$. Using that the functor $\colim_{{\mathcal K}}$ is strong symmetric monoidal, the above discussion gives a pair of adjoint functors
\begin{equation}\label{eq:structured-colim-const}
\xymatrix{ \colim_{{\mathcal K}} \colon {\mathcal S}^{{\mathcal K}}[\mathbb
D] \ar@<0.5ex>[r] & {\mathcal S}[\mathbb D] \thinspace \colon \! \!
\const_{{\mathcal K}} \ar@<0.5ex>[l] }.
\end{equation}
The next result is a structured version of Proposition \ref{prop:colim-const-Q-adjunction}.
\begin{proposition}\label{prop:structured-colim-const}
Let ${\mathcal D}$ be a $\Sigma$-free operad in ${\mathcal S}$ and let ${\mathcal S}^{{\mathcal K}}[\mathbb D]$ be equipped with the projective ${\mathcal K}$-model structure and ${\mathcal S}[\mathbb D]$ with the standard model structure. Then the Quillen adjunction
\eqref{eq:structured-colim-const} is a Quillen equivalence if and only if $B{\mathcal K}$ is contractible.
\end{proposition}
\begin{proof}
First we reduce to the case where the spaces ${\mathcal D}(n)$ of the operad are cofibrant in the coarse model structure on ${\mathcal S}^{\Sigma_n}$ for all $n$. This is automatic in the simplicial setting and in the topological setting we may replace ${\mathcal D}$ by the geometric realization of its singular complex
$\bar{\mathcal D}=|\Sing {\mathcal D}|$ which has this property. Indeed, the canonical map
$\bar{\mathcal D}\to {\mathcal D}$ is a weak equivalence of operads so that by
Proposition~\ref{prop:operad-change} it suffices to prove the result for
$\bar{\mathcal D}$. With this assumption on ${\mathcal D}$,
Corollary~\ref{cor:projective-underlying-cofibrant} implies that the underlying
${\mathcal K}$-space of a cofibrant $\mathbb D$-algebra is cofibrant in the projective model structure on ${\mathcal S}^{{\mathcal K}}$. From here the argument proceeds exactly as in the proof of the non-structured statement in
Proposition~\ref{prop:colim-const-Q-adjunction}.
\end{proof}
Combining the above proposition with
Corollary~\ref{cor:positive-projective-rectification} allows us to rectify $E_{\infty}$ spaces to strictly commutative ${\mathcal K}$-space monoids provided that $B{\mathcal K}$ is contractible.
\begin{theorem}\label{theorem:K-E-infinity-rectification}
Let ${\mathcal D}$ be an $E_{\infty}$ operad and let ${\mathcal K}$ be a well-structured index category with contractible classifying space. Suppose that ${\mathcal K}_+\to{\mathcal K}$ is homotopy cofinal and that $({\mathcal K},O{\mathcal K}_+)$ is very well-structured. Then the positive projective ${\mathcal K}$-model structure on ${\mathcal C}{\mathcal S}^{{\mathcal K}}$ is related to the standard model structure on ${\mathcal S}[\mathbb D]$ by a chain of Quillen equivalences.\qed
\end{theorem}
In order to investigate the situation when $B{\mathcal K}$ is not contractible we specialize further and assume from now on that the
symmetric monoidal category ${\mathcal K}$ is in fact permutative.
The reason for this is the following
lemma whose proof is analogous to the proof for the special case
${\mathcal K}={\mathcal I}$ considered in~\cite[Lemma~6.7]{Schlichtkrull-Thom_symm}. Let
$\mathcal E$ be the Barratt-Eccles operad, i.e., the operad whose
$k$th space is the classifying space of the translation category of
$\Sigma_k$. We write $\mathbb E$ for the associated monad.
\begin{lemma}\label{lem:Barratt-Eccles-action}
If ${\mathcal K}$ is permutative, then the ${\mathcal K}$-spaces $E{\mathcal K}$ and $B{\mathcal K}$ have canonical $\mathbb E$-algebra structures such that $\pi\colon E{\mathcal K}\to B{\mathcal K}$
is a map of $\mathbb E$-algebras. \qed
\end{lemma}
Here we again view $B{\mathcal K}$ as a constant ${\mathcal K}$-space. By an operad augmented over ${\mathcal E}$ we understand an operad ${\mathcal D}$ equipped with a map of operads ${\mathcal D}\to {\mathcal E}$. In the following we assume that ${\mathcal D}$ is $\Sigma$-free such that Proposition \ref{prop:structured-lift-proposition} provides a projective ${\mathcal K}$-model structure on ${\mathcal S}^{{\mathcal K}}[\mathbb D]$.
It is easy to see that the adjunction \eqref{eq:EK-adjunction}
lifts to an adjunction of structured diagram spaces.
\begin{lemma}\label{lem:structured-EK-Q-equivalence}
Let ${\mathcal D}$ be a $\Sigma$-free operad augmented over ${\mathcal E}$. Then the adjoint functors ${\mathcal S}^{{\mathcal K}}[\mathbb D]/E{\mathcal K}
\rightleftarrows{\mathcal S}^{{\mathcal K}}[\mathbb D]$ form a Quillen equivalence
with respect to the projective ${\mathcal K}$-model structures. \qed
\end{lemma}
With $\mathbb D$ as above we also have a structured version of the adjunctions in \eqref{eq:EKBKadjunction},
\begin{equation}\label{eq:structured-EKBKadjunction}
\xymatrix{
\colim_{{\mathcal K}}\colon {\mathcal S}^{{\mathcal K}}[\mathbb D]/E{\mathcal K} \ar@<0.5ex>[r] &
{\mathcal S}^{{\mathcal K}}[\mathbb D]/B{\mathcal K} \ar@<0.5ex>[l]
\ar@<0.5ex>[r] & {\mathcal S}[\mathbb D]/B{\mathcal K} \ar@<0.5ex>[l]
\thinspace \thinspace\colon\! \pi^*.
}
\end{equation}
When ${\mathcal S}^{{\mathcal K}}[\mathbb D]$ is equipped with the projective
${\mathcal K}$-model structure and ${\mathcal S}[\mathbb D]$ with the standard model structure discussed in Remark~\ref{rem:model-str-D-spaces}, the arguments for the adjunctions in \eqref{eq:EKBKadjunction} also apply here to show that the composition is a Quillen adjunction.
\begin{lemma}\label{lem:structured-EKBK-Q-equivalence}
The adjunction $(\colim_{{\mathcal K}},\pi^*)$ in \eqref{eq:structured-EKBKadjunction}
is a Quillen equivalence with respect to the projective
${\mathcal K}$-model structure on ${\mathcal S}^{{\mathcal K}}[\mathbb D]/E{\mathcal K}$ and the standard model structure on ${\mathcal S}[\mathbb D]/B{\mathcal K}$.
\end{lemma}
\begin{proof}
Arguing as in the proof of Proposition~\ref{prop:structured-colim-const} we first reduce to the case where the spaces ${\mathcal D}(n)$ of the operad are cofibrant in the coarse model structure on ${\mathcal S}^{\Sigma_n}$ for all $n$.
Using Corollary~\ref{cor:projective-underlying-cofibrant}, the argument then proceeds exactly as in the proof of Lemma \ref{lem:EKBK-Q-equivalence}.
\end{proof}
Combining Lemmas \ref{lem:structured-EK-Q-equivalence} and
\ref{lem:structured-EKBK-Q-equivalence} we get a
structured version of Theorem~\ref{thm:K-spaces-over-BK}.
\begin{theorem}\label{thm:structured-K-spaces-over-BK}
Let ${\mathcal K}$ be a well-structured index category that is permutative as a
symmetric monoidal category and let ${\mathcal D}$ be a $\Sigma$-free operad augmented over ${\mathcal E}$. Then there is a chain of Quillen equivalences
\[
\xymatrix{
{\mathcal S}^{{\mathcal K}}[\mathbb D] \ar@<-0.5ex>[r] &
{\mathcal S}^{{\mathcal K}}[\mathbb D]/E{\mathcal K} \ar@<-0.5ex>[l]
\ar@<0.5ex>[r]^-{\colim_{{\mathcal K}}} & {\mathcal S}[\mathbb D]/B{\mathcal K} \ar@<0.5ex>[l]^-{\pi^*}
}
\]
relating the projective ${\mathcal K}$-model structure on ${\mathcal S}^{{\mathcal K}}[\mathbb D]$ to the standard model structure on ${\mathcal S}[\mathbb D]/B{\mathcal K}$.\qed
\end{theorem}
For instance, this theorem applies to the associativity operad in which case it relates the corresponding category of (not necessary commutative)
${\mathcal K}$-space monoids to the category of monoids (in ${\mathcal S}$) over $B{\mathcal K}$. Applying the theorem to the operad ${\mathcal E}$ itself and combining the result with Corollary~\ref{cor:positive-projective-rectification}, we finally get the next theorem where $\mathbb E$ again denotes the monad associated to ${\mathcal E}$.
\begin{theorem}\label{thm:comm-graded-theorem}
Let ${\mathcal K}$ be a well-structured index category whose underlying symmetric monoidal category is permutative. Suppose that ${\mathcal K}_+\to {\mathcal K}$ is homotopy cofinal and that $({\mathcal K},O{\mathcal K}_+)$ is very well-structured. Then there is a chain of Quillen equivalences relating the positive projective ${\mathcal K}$-model structure on ${\mathcal C}{\mathcal S}^{{\mathcal K}}$ to the standard model structure on ${\mathcal S}[\mathbb E]/B{\mathcal K}$. \qed
\end{theorem}
\section{Diagram spaces and symmetric spectra}
\label{sec:diagram-spaces-symmetric-spectra}
Let ${\mathcal K}$ be a small symmetric monoidal category and assume that we are given a strong symmetric monoidal functor $H \colon {\mathcal K}^{{\mathrm{op}}} \to {\mathrm{\textrm{Sp}^{\Sigma}}}$. The examples to keep in mind are the functors $F_{-}(S^{-})$ in
Lemmas~\ref{lem:I-and-free-spsym} and \ref{lem:J-and-free-spsym}. In general, such a functor $H$ gives rise to an adjunction
\begin{equation}\label{eq:k-space-spsym-adj}
{\mathbb S}^{{\mathcal K}}[-] \colon {\mathcal S}^{{\mathcal K}} \rightleftarrows {\mathrm{\textrm{Sp}^{\Sigma}}} \colon
\Omega^{{\mathcal K}}
\end{equation}
relating the categories of ${\mathcal K}$-spaces and symmetric spectra. The left adjoint takes a ${\mathcal K}$-space $X$ to the coend
\[{\mathbb S}^{{\mathcal K}}[X]=\int^{\bld k\in {\mathcal K}}H(\bld k)\wedge X(\bld k)_+
\]
of the diagram $H \wedge X_+ \colon {\mathcal K}^{{\mathrm{op}}} \times{\mathcal K} \to {\mathrm{\textrm{Sp}^{\Sigma}}}$,
where $X(\bld k)_+$ denotes the union of $X(\bld k)$ with a disjoint base point. The right adjoint takes a symmetric spectrum $E$ to the ${\mathcal K}$-space
$\Omega^{{\mathcal K}}(E)$ defined by
\[
\Omega^{\cK}(E)(\bld{k}) =\Map_{{\mathrm{\textrm{Sp}^{\Sigma}}}}(H(\bld{k}),E).
\]
We recall that ${\mathrm{\textrm{Sp}^{\Sigma}}}$ is enriched over based spaces with
$\Map_{{\mathrm{\textrm{Sp}^{\Sigma}}}}(D,E)$ defined as the appropriate subspace of the product of the mapping spaces $\Map(D(n),E(n))$.
The next lemma is a consequence of $H$ being strong symmetric monoidal.
\begin{lemma}\label{lem:S_J_Omega_J-monoidal}
The functor ${\mathbb S}^{{\mathcal K}}[-]$ is strong symmetric monoidal and $\Omega^{\cK}$ is lax symmetric monoidal.\qed
\end{lemma}
Now let ${\mathcal D}$ be an operad in ${\mathcal S}$ and let us write $\mathbb D$ both for the associated monad on ${\mathcal S}^{{\mathcal K}}$ and the associated monad on ${\mathrm{\textrm{Sp}^{\Sigma}}}$.
By the discussion at the beginning of Section \ref{subs:structured-comma}, the fact that $\mathbb S^{{\mathcal K}}[-]$ is strong symmetric monoidal implies that the adjunction \eqref{eq:k-space-spsym-adj} lifts to the corresponding categories of $\mathbb D$-algebras.
\begin{proposition}\label{prop:structured-adjunction-SK-Spsym}
Let ${\mathcal D}$ be an operad in ${\mathcal S}$. Then the adjunction $({\mathbb S}^{{\mathcal K}}[-], \Omega^{{\mathcal K}})$ lifts to an adjunction relating the associated categories of
$\mathbb D$-algebras,
\[
{\mathbb S}^{{\mathcal K}}[-] \colon {\mathcal S}^{{\mathcal K}}[\mathbb D] \rightleftarrows {\mathrm{\textrm{Sp}^{\Sigma}}}[\mathbb D] \colon\Omega^{{\mathcal K}}.
\eqno\qed
\]
\end{proposition}
This applies in particular to give an adjunction between the categories of (commutative) ${\mathcal K}$-space monoids and (commutative) symmetric ring spectra.
It remains to analyze the homotopical properties of these adjunctions. For this we need the next two lemmas.
\begin{lemma}\label{lem:mS_of_FK}
There are natural isomorphisms
\[
{\mathbb S}^{{\mathcal K}}[F^{{\mathcal K}}_{\bld{k}}(K)] \cong H(\bld{k})\wedge K_+
\qquad \text{and} \qquad
{\mathbb S}^{{\mathcal K}}[G^{{\mathcal K}}_{\bld{k}}(L)] \cong H(\bld{k})\wedge_{{\mathcal K}(\bld{k})} L_+
\]
for any space $K$ and any ${\mathcal K}(\bld k)$-space $L$.
\end{lemma}
\begin{proof}
By the uniqueness of adjoint functors, the second isomorphism follows from the chain of natural isomorphisms
\[
{\mathcal S}^{{\mathcal K}(\bld{k})}(L,\Ev^{{\mathcal K}}_{\bld{k}} \Omega^{{\mathcal K}}(E))
\cong {\mathcal S}^{{\mathcal K}(\bld{k})}(L, \Map_{{\mathrm{\textrm{Sp}^{\Sigma}}}}(H(\bld{k}), E)) \cong
{\mathrm{\textrm{Sp}^{\Sigma}}}(H(\bld{k})\wedge_{{\mathcal K}(\bld{k})} L_+, E).\] The first isomorphism is a consequence of the second.
\end{proof}
Given a space $K$, a morphism $\alpha\colon\bld{k}\to\bld{l}$ in ${\mathcal K}$ induces a map of ${\mathcal K}$-spaces $ \alpha^* \colon F^{{\mathcal K}}_{\bld{l}}(K) \to
F^{{\mathcal K}}_{\bld{k}}(K)$.
\begin{lemma}\label{lem:ms_of_maps_of_FK}
Via the first isomorphism of Lemma \ref{lem:mS_of_FK}, the induced map
${\mathbb S}^{{\mathcal K}}[\alpha^*]$ corresponds to
\[
H(\alpha)\wedge K_+\colon H(\bld{l})\wedge K_+ \to H(\bld{k})\wedge K_+.
\eqno\qed
\]
\end{lemma}
As for the category of ${\mathcal K}$-spaces, there are several useful model structures on the category of symmetric spectra. Most importantly, we have the (positive) projective model structures from \cite{HSS} and \cite{MMSS} and the (positive) flat model structures from \cite{HSS} and
\cite{Shipley-convenient} (where they are called $S$-model structures); see also \cite{Schwede-SymSp}. In the next proposition we assume that
${\mathrm{\textrm{Sp}^{\Sigma}}}$ is equipped with an ${\mathcal S}$-model structure whose weak equivalences are the stable equivalences and we provide necessary and sufficient conditions on the functor $H$ for the $({\mathbb S}^{{\mathcal K}}[-], \Omega^{{\mathcal K}})$ adjunction to be a Quillen adjunction.
\begin{proposition}\label{prop:K-spaces-Spsym-adjunction}
Let $({\mathcal K},{\mathcal A})$ be a well-structured relative index category and let ${\mathcal S}^{{\mathcal K}}$ be equipped with the ${\mathcal A}$-relative ${\mathcal K}$-model structure. Let ${\mathrm{\textrm{Sp}^{\Sigma}}}$ be equipped with an ${\mathcal S}$-model structure whose weak equivalences are the stable equivalences. Then the $({\mathbb S}^{{\mathcal K}}[-], \Omega^{{\mathcal K}})$ adjunction \eqref{eq:k-space-spsym-adj} is a Quillen adjunction if and only if
\begin{enumerate}[(i)]
\item
$H(\bld k)/K$ is cofibrant in the given model structure on ${\mathrm{\textrm{Sp}^{\Sigma}}}$ for all objects $\bld k$ in ${\mathcal A}$ and all subgroups $K\subseteq {\mathcal A}(\bld k)$,
\item
the canonical map $H(\bld k)_{hK}\to H(\bld k)/K$ is a stable equivalence for all objects $\bld k$ in ${\mathcal A}$ and all subgroups $K\subseteq {\mathcal A}(\bld k)$, and
\item
the induced map $H(\alpha)\colon H(\bld l)\to H(\bld k)$ is a stable equivalence for all morphisms $\alpha\colon \bld k\to \bld l$ in ${\mathcal K}_{{\mathcal A}}$.
\end{enumerate}
\end{proposition}
\begin{proof}
We show that the assumptions in the proposition are sufficient for the adjunction to be a Quillen adjunction. Similar arguments show that the assumptions are also necessary. Using the criterion in \cite[Lemma 2.1.20]{Hovey-model}, it suffices to prove that $\mathbb S^{{\mathcal K}}[-]$ takes the generating cofibrations for ${\mathcal S}^{{\mathcal K}}$ to cofibrations in ${\mathrm{\textrm{Sp}^{\Sigma}}}$ and the generating acyclic cofibrations for
${\mathcal S}^{{\mathcal K}}$ to stable equivalences. In the following $\bld k$ denotes an object in ${\mathcal A}$, $K$ is a subgroup of ${\mathcal A}(\bld k)$, and $i$ is a generating cofibration for ${\mathcal S}$. By Lemma \ref{lem:mS_of_FK}, $\mathbb S^{{\mathcal K}}[-]$ takes a generating cofibration of the form
$G_{\bld k}^{{\mathcal K}}({\mathcal K}(\bld k)/K\times i)$ to the map $H(\bld k)/K\wedge i_+$, up to isomorphism. The latter is a cofibration in ${\mathrm{\textrm{Sp}^{\Sigma}}}$ by assumption (i) and the requirement that the model structure on ${\mathrm{\textrm{Sp}^{\Sigma}}}$ be an ${\mathcal S}$-model structure. If we replace $i$ with a generating acyclic cofibration for ${\mathcal S}$, a similar argument gives that $\mathbb S^{{\mathcal K}}[-]$ takes the resulting generating acyclic cofibration to a stable equivalence. Consider then a generating acyclic cofibration $G_{\bld k}^{{\mathcal K}}(j_K\Box i)$ where $j_K$ is as in
\eqref{eq:pi_H-factorization}. The functors $\mathbb S^{{\mathcal K}}[-]$ and $G_{\bld k}^{{\mathcal K}}$ both preserve tensors and colimits so $\mathbb S^{{\mathcal K}}[G_{\bld k}^{{\mathcal K}}(j_K\Box i)]$ can be identified with the pushout-product
$\mathbb S^{{\mathcal K}}[G_{\bld k}^{{\mathcal K}}(j_K)]\Box i$ in ${\mathrm{\textrm{Sp}^{\Sigma}}}$. As in Section \ref{subs:G-space-model}, let $\pi_K$ denote the projection
${\mathcal K}(\bld k)\times_K EK\to {\mathcal K}(\bld k)/K$, and observe that by
Lemma~\ref{lem:mS_of_FK} we can identify $\mathbb S^{{\mathcal K}}[G_{\bld k}^{{\mathcal K}}(\pi_K)]$ with the canonical map $H(\bld k)_{hK}\to H(\bld k)/K$ which is a stable equivalence by assumption (ii). Since $\mathbb S^{{\mathcal K}}[-]$ and
$G_{\bld k}^{{\mathcal K}}$ preserve mapping cylinders, this implies that
$\mathbb S^{{\mathcal K}}[G_{\bld k}^{{\mathcal K}}(j_K)]$ is an acyclic cofibration which by the ${\mathcal S}$-model structure on ${\mathrm{\textrm{Sp}^{\Sigma}}}$ shows that
$\mathbb S^{{\mathcal K}}[G_{\bld k}^{{\mathcal K}}(j_K)]\Box i$ is a stable equivalence.
Finally, consider a generating acyclic cofibration
$G_{\bld k}^{{\mathcal K}}(j_{\alpha}\Box i)$ where $\alpha$ is a morphism in
${\mathcal K}_{{\mathcal A}}$ and $j_{\alpha}$ is as in \eqref{eq:alpha-star-fact}. Arguing as above, it suffices to show that $\mathbb S^{{\mathcal K}}[G_{\bld k}^{{\mathcal K}}(j_{\alpha})]$ is a stable equivalence and using Lemma \ref{lem:ms_of_maps_of_FK} this can easily deduced from assumption (iii).
\end{proof}
\begin{remark}
Let ${\mathcal D}$ be an operad in ${\mathcal S}$. It is a formal consequence of the above proposition that whenever the model structures under consideration can be lifted to the categories of $\mathbb D$-algebras in ${\mathcal S}^{{\mathcal K}}$ and ${\mathrm{\textrm{Sp}^{\Sigma}}}$, then the stated assumptions on $H$ imply that the induced adjunction of
$\mathbb D$-algebras in Proposition \ref{prop:structured-adjunction-SK-Spsym} is a Quillen adjunction.
\end{remark}
|
\section{Introduction}
It has long been speculated that the boundary plays a very significant role in establishing the physical properties of a quantum field theory. This idea has been very fruitful in clarifying the physics of the fractional quantum Hall effect, and is also the origin of the holographic principle in black hole physics. An explicit manifestation of this fact is the so-called area law. The area law states that for ground (thermal) states of lattice systems with short-range interactions, the entropy (quantum mutual information) of the reduced density operator $\rho_A$, corresponding to a region $A$, is proportional to the surface of that region, rather than to the volume, at least for gapped systems \cite{Srendnicki,Wilczek2,Hastings,WVHC}. Criticality may reflect itself by the appearance of multiplicative and/or linear logarithmic corrections to the area law \cite{calabrese,vidallatorre03}.
Apart from the deep physical significance of this law, it has important implications regarding the possibility of simulating many-body quantum systems using tensor network (TN) states \cite{white92,schollwoeck04,murgadvances,ciracreview}.
For instance, it has been shown \cite{verstraetecirac05} that any state of a quantum spin system fulfilling the area law in one spatial dimension (including logarithmic violations) can be efficiently represented by a matrix product state (MPS) \cite{fannes92,perezgarcia06}, the simplest version of a TN.
Very recently, another remarkable discovery has been made with relation to the area law \cite{entspectrum}. It has been shown that for certain models in two
spatial dimensions, the reduced density matrix of a region $A$ has a very peculiar spectrum, which is called the "entanglement spectrum": by taking the logarithm of the eigenvalues of $\rho_A$, one obtains a spectrum that resembles very much the one of a 1-dimensional critical theory (i.e. as prescribed by conformal field theory).
This has been established for different systems as diverse as gapped
fractional quantum Hall states~\cite{entspectrum} or spin-1/2 quantum
magnets~\cite{poilblanc2010}. Interestingly, the correlation length in the bulk of the ground state can be naturally interpreted as a thermal length in
one dimension~\cite{poilblanc2010}.
This is all very suggestive for the fact that the reduced density matrix is the thermal state of a 1-dimensional theory. However, there is a clear mismatch in dimensions: the Hilbert space associated to $\rho_A$ has two spatial dimensions, while the 1-dimensional theory obviously has only 1. Intuitively, this is clear as all relevant degrees of freedom of $\rho_A$ should be located around the boundary of region $A$. The main question addressed in this paper is to explicitly identify the degrees of freedom on which this 1-dimensional Hamiltonian acts.
We show that projected entangled-pair states (PEPS) \cite{verstraetecirac04}
give a very natural answer to that question. The degrees of freedom of the 1-dimensional theory correspond to the virtual
particles which appear in the valence bond description of PEPS, and that
"live" at the boundary of region $A$
\cite{verstraetecirac04b,verstraetecirac04}. More specifically, PEPS are
built by considering a set of virtual particles at each node of the
lattice, which are then projected out to obtain the state of the physical
spins. As we show, the boundary Hamiltonian can be thought of as acting on
the virtual particles that live at the boundary of region $A$.
Furthermore, we will present evidence that, for gapped systems, such a boundary Hamiltonian is quasi--local (i.e. contains only
short-range interactions) in terms of those (localized) virtual particles. As a quantum phase transition is approached, the range of
the interactions increases. Finally, we will show that the interactions lose their local character for the case of quantum systems exhibiting topological order. We will also show how operators in the bulk can be mapped to operators on the boundary.
The fact that the boundary Hamiltonian is quasi--local has important
implications for the theory of PEPS which go well beyond those of the area
law. While PEPS are expected to accurately represent well the low energy
sector of local Hamiltonians in arbitrary dimensions \cite{hastings06}, it has not been proven that one can use them to
determine expectation values in an efficient and accurate way. For that,
one has to contract a set of tensors, a task which could in principle require
exponential time in the size of the lattice. In order to circumvent this
problem, a method was introduced \cite{verstraetecirac04} which successively approximates the boundary of a growing region by a matrix product density operator, which is exactly the density matrix of local virtual particles discussed before. It is not clear a
priori to which extent that density matrix can be approximated by a MPS; more specifically,
the bond dimension of that MPS could in principle grow exponentially with the size of the
system if a prescribed accuracy is to be reached, which would lead to an exponential scaling of the computational effort. However, that MPS does nothing but approximate the boundary density operator $\rho_A$ for different regions $A$. In case such an operator can
be written as a thermal state of a quasilocal Hamiltonian, it immediately
follows that in order to approximate it by a MPS one just needs a bond
dimension that scales polynomially with the lattice
size~\cite{hastings06}, and thus that expectation values of PEPS can be
efficiently determined.
\section{PEPS and boundary theories}
\subsection{Model}
We consider a PEPS, $|\Psi\rangle$, of an $N_v\times N_h$ spin lattice in two spatial dimensions. Note that one can always find a finite-range interaction Hamiltonian for which $|\Psi\rangle$ is a ground state \cite{verstraetewolf06}. We will assume that we have open (periodic) boundary conditions in the horizontal (vertical) direction: the spins are regularly placed on a cylinder and the state $|\Psi\rangle$ is translationally invariant along the vertical direction [see Fig. (\ref{Fig1})]. All spins have total spin $S$, except perhaps at the boundaries where we may choose a different spin in order to lift degeneracies related to the open boundary conditions. We will be interested in the reduced density operator, $\rho_\ell$, corresponding to the spins lying in the first $\ell$ columns; that is, when we trace all the spins from column $\ell+1$ to $N_h$.
More specifically, the effective Hamiltonian, $H_\ell$, corresponding to those spins, is defined through $\rho_\ell=\exp(-H_\ell)/Z_\ell$, with $Z_\ell$ a normalization constant. We will be interested not only in the entanglement spectrum \cite{entspectrum}, but also in the specific form of $H_\ell$ and its interaction length, as we will define below.
In order to simplify the notation, it is convenient to label the spin indices of each column with a single vector. We define $I_n=(i_{1,n},i_{2,n},\ldots,i_{N_v,n})$, where $i_{k,n}=-S/2,-S/2+1,\ldots,S/2$ for $n=2,\ldots,N_h-1$ (for $n=1$ or $n=N_h$ we may have different spin $S$). Thus, we can write
\begin{equation}
|\Psi\rangle = \sum_{I} c_{I}|I_1,I_2,\ldots,I_{N_h}\rangle.
\end{equation}
For a PEPS we can write
\begin{equation}
c_{I}= \sum_\Lambda L_{\Lambda_1}^{I_1} B_{\Lambda_1,\Lambda_2}^{I_2}
\ldots B_{\Lambda_{N_h-2},\Lambda_{N_h-1}}^{I_{N_{h-1}}}
R_{\Lambda_{N_h-1}}^{I_{N_h}} .
\end{equation}
Here $\Lambda_n=(\alpha_{1,n},\alpha_{2,n},\ldots,\alpha_{N_v,n})$, where $\alpha_{k,n}=1,2,\ldots,D$ with $D$ the so-called bond dimension. Each of the $B^I$'s can be expressed in terms of a single tensor, $\hat A^i$,
\begin{equation}
B_{\Lambda_{n-1},\Lambda_n}^{I_n} = \mathrm{tr}\left[\prod_{k=1}^{N_v} \hat A^{i_{k,n}}_{\alpha_{k,n-1},\alpha_{k,n}} \right],
\end{equation}
where for each value of $i,\alpha,\alpha'$, $\hat A^i_{\alpha,\alpha'}$ is a $D\times D$ matrix, with elements $A^i_{\alpha,\alpha';\beta,\beta'}$ (the indices $\alpha$ and $\beta$ correspond to the virtual particles entangled along the horizontal and vertical directions, respectively \cite{verstraetecirac04}; see Fig. \ref{Fig1}). For the first (left) and last (right) column we define $L^I$ and $R^I$ similarly in terms of the $D\times D$ matrices $\hat l^i_{\alpha}$, and $\hat r^i_{\alpha'}$:
\begin{eqnarray}
L_{\Lambda_1}^{I_1} &=& \mathrm{tr}\left[\prod_{k=1}^{N_v} \hat l^{i_{k,1}}_{\alpha_{k,1}} \right], \\
R_{\Lambda_{N_h-1}}^{I_{N_h}} &=& \mathrm{tr}\left[\prod_{k=1}^{N_v} \hat r^{i_{k,N_h}}_{\alpha_{k,N_h-1}} \right].
\end{eqnarray}
Thus, the tensors $\hat A$, $\hat l$, and $\hat r$ (for which explicit expressions will be given later on) completely characterize the state $|\Psi\rangle$, which is obtained by "tiling" them on the surface of the cylinder. The first has rank 5, whereas the other two have rank 4. Here we have taken all the tensors $A$ equal, but they can be chosen to be different if the appropriate symmetries are not present.
\begin{figure}
\includegraphics[width=\columnwidth]{Fig1a}
\caption{Top: We consider an $N_v\times N_h$ spin lattice in a cylindrical geometry. The PEPS is obtained by replacing each lattice site with a tensor A, and contracting the virtual indices $\alpha$ and $\beta$ along the horizontal and vertical directions. Bottom: We cut the lattice into two pieces, left and right. The virtual indices $\alpha$ of the tensors $A$ along the cut are shown. The state $|\Psi_L\rangle$ acts on the spins ($i_{k,n}$) as well as on the virtual spins along the cut.}
\label{Fig1}
\end{figure}
\subsection{Boundary density operator}
We now want to express the reduced density operator $\rho_\ell$ in terms of the original tensors. In order to do that, we block all the spins that are in the first $\ell$ columns, and those in the last $N_h-\ell$, and define
\begin{equation}
\hat L^{I_a} = L^{I_1} B^{I_2} \ldots B^{I_\ell}, \quad \hat R^{I_b} = B^{I_{\ell+1}} \ldots B^{I_{N_h-1}} R^{I_{N_h}},
\end{equation}
where we have collected all the indices $I_1,\ldots,I_\ell$ in $I_a$ and the rest in $I_b$. With this notation, the state $|\Psi\rangle$ can be considered as a two-leg ladder, ie $\hat N_h=2$, and $\hat\ell=1$, where $\rho_{\hat\ell}$ is the density operator corresponding to a single leg. Thus, we have
\begin{equation}
\label{2Leg}
|\Psi\rangle = \sum_{I_a,I_b}\sum_{\Lambda} \hat L^{I_a}_\Lambda \hat R^{I_b}_\Lambda |I_a,I_b\rangle.
\end{equation}
It is convenient to consider the space where the vectors $L^I$ and $R^I$ act as a Hilbert space, and use the bra/ket notation there as well. That space, that we call {\it virtual space}, is the one corresponding to the ancillas that build the PEPS in the valence bond construction \cite{verstraetecirac04}. They are associated to the boundary between the $\ell$--th and the $\ell+1$st columns of the original spins. The dimension is thus $D^{N_v}$ (see Fig. \ref{Fig1}). In order to avoid confusion with the space of the spins, we have used $|v)$ to denote vectors on that space. We can define the (unnormalized) joint state for the first $\ell$ columns and the virtual space, $|\Psi_L\rangle$, and similarly for the last columns, $|\Psi_R\rangle$, as
\begin{equation}
|\Psi_L\rangle = \sum_{I_a} |\hat L^{I_a})|I_a\rangle,\quad
|\Psi_R\rangle = \sum_{I_a} |\hat R^{I_b})|I_b\rangle
\end{equation}
with
\begin{equation}
|\hat L^{I_a})=\sum_\Lambda \hat L^{I_a}_\Lambda |\Lambda),\quad
|\hat R^{I_b})=\sum_\Lambda \hat R^{I_b}_\Lambda |\Lambda),
\end{equation}
and $|\Lambda)$ the canonical orthonormal basis in the corresponding
virtual spaces. The state $|\Psi\rangle$ can then be straightforwardly
defined in terms of those two states. The corresponding reduced density
operators for both virtual spaces are
\begin{equation}
\label{sigmaLR}
\sigma_L = \sum_{I_a} |\hat L^{I_a})(\hat L^{I_a}|\ ,\quad \sigma_R =
\sum_{I_b} |\hat R^{I_b})(\hat R^{I_b}|\ .
\end{equation}
In terms of those operators, it is very simple to show that
\begin{equation}
\rho_\ell = \sum_{\Gamma,\Gamma'} \vert \chi_\Gamma\rangle\langle
\chi_{\Gamma'}| \; (\Gamma|\sqrt{\sigma_L^T} \sigma_R
\sqrt{\sigma_L^T}|\Gamma')
\end{equation}
where $\vert\Gamma)$ is an orthonormal basis of the range of
$\sigma_L$, $\sigma_L^T$ is the transpose of $\sigma_L$ in the basis $|\Lambda)$,
and where we have defined an orthonormal set (in the spin space)
\begin{equation}
|\chi_\Gamma\rangle = \sum_I (\Gamma|\frac{1}{\sqrt{\sigma_L}}|\hat L^I)
\;|I\rangle\ .
\end{equation}
Now, defining an isometric operator that transforms the virtual onto the spin space
${\cal U}=\sum_\Gamma |\chi_\Gamma\rangle(\Gamma|$, we have
\begin{equation}
\rho_\ell=U \sqrt{\sigma_L^T} \sigma_R \sqrt{\sigma_L^T} U^\dagger\ .
\end{equation}
The isometry $U$ can also be used to map any operator acting on the bulk onto the virtual spin space; note that this map is an isometry and hence not injective, i.e. a boundary operators might correspond to many different bulk operators. This is of course a necessity, as $U$ is responsible for mapping a 2-dimensional theory to a 1-dimensional one.
\subsection{Boundary Hamiltonian}
The previous equation shows that $\rho_\ell$ is directly related to the
density operators corresponding to the virtual space of the ancillary
spins that build the PEPS. In particular, if we have
$\sigma_L^T=\sigma_R=:\sigma_b$ (eg., when we have the appropriate symmetries
as in the specific cases analyzed below), then $\rho_\ell=U \sigma_b^2 U^\dagger$. The reduced density operator $\rho_\ell$ is thus directly related to that of the virtual spins along the boundary. Since $U$ is isometric it conserves the spectrum and thus the entanglement spectrum of $\rho_\ell$ will coincide with that of $\sigma_b^2$. By writing $\sigma_b^2=\exp(-H_b)$, we obtain an effective one-dimensional Hamiltonian for the virtual spins at the
boundary of the two regions whose spectrum coincides with the
entanglement spectrum of $\rho_\ell$.
We will be interested to see to what extent $H_b$ is a local Hamiltonian
for the boundary (virtual) space. We can always write $H_b$ as a sum of
terms where different spin operators. For instance, for $D=2$, we can take the Pauli operators $\sigma_\alpha$ ($\alpha=x,y,z$) acting on different spins, and the identity operator on the rest. We group those terms into sums $h_n$, where each $h_n$ contains all terms with interaction range $n$, i.e.,
for which the longest contiguous block of identity
operators has length $N_v-n$. For instance, $h_0$ contains only one term,
which is a constant; $h_1$ contains all terms where only one Pauli
operator appears; and $h_{N_v}$ contains all terms where no identity
operator appears. We define
\begin{equation}
d_n=\mathrm{tr}(h_n^2)/2^{N_v}\, ,
\label{Eq:dn}
\end{equation}
which expresses the strength of all the terms in the Hamiltonian with interaction length equal to $n$. A fast decrease of $d_n$ with $n$ indicates that the effective Hamiltonian describing the virtual boundary is quasi-local. In the examples we examine below this is the case as long as we do not have a quantum phase transition. In such a case, the length of the effective Hamiltonian interaction increases.
\subsection{Implications for PEPS}
In case $\sigma_b$ can be written in terms of a local boundary Hamiltonian one can draw important consequences for the theory of PEPS. In particular, it implies that the PEPS can be efficiently contracted, and correlation functions can be efficiently determined. The reason can be understood as follows. Let us consider again the cylindrical geometry (Fig.\ref{Fig1}), and let us assume that we want to determine any correlation function along the vertical direction, eg at the lattice points $(\ell,1)$ and $(\ell,x)$. It is very easy to show that such a quantity can be expressed in terms of $\sigma_L$ and $\sigma_R$. If we are able to write these two operators as Matrix Product Operators (MPO), ie as
\begin{equation}
\sum_{i_n,j_n,=1}^D \mathrm{tr}\left[ M^{i_1,j_1} \ldots M^{i_{N_v},j_{N_v}}\right] |i_1,\ldots,i_{N_v}\rangle\langle j_1,\ldots,j_{N_v}|,
\end{equation}
where the $M'$ are $D'\times D'$ matrices, then the correlation function
can be determined with an effort that scales as $N_v (D')^6$. It has been
shown by Hastings \cite{hastings06} that if an operator can be written as
$\exp(-H_b/2)$, where $H_b$ is quasilocal, then it can be efficiently
represented by an MPO; that is, the bond dimension $D'$ only scales
polynomially with $N_v$. Thus, we have that the time required to
determine correlation functions only scales polynomially with $N_v$.
Later on, when we examine various examples, we will use MPO to represent
$\sigma_b$. In that case, we can directly check if we obtain a good
approximation by using a MPO just by simply observing how much errors
increase when we decrease the bond dimension $D'$. We will see that the
error increases when we approach a quantum phase transition. Furthermore,
whenever $\sigma_b$ can be well approximated by a MPO, we can use the
knowledge gained in the context of MPS \cite{fannes92,perezgarcia06} to
observe the appearance of a quantum phase transition in the original PEPS.
For that, we just have to recall that the correlation length, $\xi$, is
related to the two largest (in magnitude) eigenvalues, $\lambda_{1,2}$, of
the matrix $\sum_i M^{i,i}$; $\xi=1/\log(|\lambda_1/\lambda_2|)$. For
$|\lambda_1|=|\lambda_2|$, the correlation length diverges indicating the
presence of a quantum phase transition.
\subsection{Qualitative discussion}
In order to better understand the structure of $\sigma_b$, let us first consider a 1D spin chain. Even though the boundary of the chain, when cut into two parts, has zero dimensions, it will help us to understand the 2D systems. We take $N_v=1$ so that the PEPS reduces to a MPS. We can use the theory of MPS \cite{fannes92,perezgarcia06} to analyze the properties of the completely positive map (CPM) ${\cal E}$ (the matrices $A^i$ of the MPS are the Kraus operators of the CPM). In the limit $N_h\to\infty$, $\sigma_b$ is nothing but the fixed point of such a CPM. For gapped systems, ${\cal E}$ has a unique fixed point, and thus $\sigma_b$ is unique. For gapless systems, ${\cal E}$ becomes block diagonal (and thus there are several fixed points), the correlation length diverges, and we can write
\begin{equation}
\label{gapless}
\sigma_b=\oplus_{n=1}^B p_n \sigma_b^{(n)},
\end{equation}
where $B$ is the number blocks which coincides with the degeneracy of the eigenvalue of ${\cal E}$ corresponding to the maximum magnitude. In such case, the weights $p_n$ depend on the tensors $l$ and $r$ which are chosen at the boundaries. For critical systems, one typically finds that $D$ increases as a polynomial in $N_v$ such that one obtains logarithmic corrections to the area law \cite{Peschel,vidallatorre03}.
The 2D geometry considered here reduces to the 1D case if we take the
limit $N_h\to\infty$ by keeping $N_v$ finite. According to the discussion
above, we expect to have a unique $\sigma_b$ if we deal with a gapped system. As we will illustrate below with some specific examples, this operator can be
written in terms of a local Hamiltonian $H_b$ of the boundary virtual
space which is quasilocal. As we approach a phase transition, the gap
closes and the correlation length diverges. In some cases, the boundary density operator can be written as a direct sum (\ref{gapless}), eventually leading to the loss of locality in the boundary Hamiltonian.
\section{Numerical Methods}
In order to determine $\sigma_b$ we make heavy use of the fact that $|\Psi\rangle$ is a PEPS.
We have followed three different complementary numerical approaches that we briefly describe here.
\subsection{Iterative procedure}
First of all, for sufficiently small values of $N_v$ (typically $N_v\le 12$) we can perform exact numerical calculations and determine $\sigma_{L,R}$ according to (\ref{sigmaLR}). The main idea is to start from the left and find first $\sigma_L$ for $\ell=1$ by contracting the tensors $l^i$ appropriately. Then, we can proceed for $\ell=2$ by contracting the tensors $A^i$ corresponding to the second column. In this vain, and as long as $N_v$ is sufficiently small we can determine $\sigma_{L,R}$ for all values of $\ell$ and $N_h$.
\subsection{Exact contractions and finite size scaling}
The second (exact) method is a variant applicable to larger values of $N_v$ (typically up to $N_v=20$) but restricted to a finite width in the
horizontal direction.
It consists in exactly contracting the internal indices of two adjacent blocks of size $N_v/2\times N_h$. These two blocks are then contracted together in a second step.
Although limited by the size $2^{N_v+2N_h}$ of the half-block (which has to fit in the computer RAM), this approach can still handle systems of size $20\times 2$ or $16\times 8$
and be supplemented by a finite size scaling analysis.
\subsection{Truncation method}
Finally, to take the $N_h\rightarrow \infty$ limit we can use the methods introduced in \cite{verstraetecirac04} to approximate the column operators. The main idea is to represent those operators by tensor networks with the structure of a MPS. We contract one column after each other, finding the optimal MPS after each contraction variationally. In particular, since we will consider translationally invariant states, we can choose the matrices of the corresponding MPS all equal, which simplifies the procedure. We can even approach the limit $N_v,N_h\to \infty$ as follows (see also \cite{murgadvances,vidal07}): (i) we start out with $\ell=1$, and contract the second column, obtaining another tensor network with the same MPS structure, but with increased bond dimensions. (ii) We continue adding columns, up to some $\ell=r$, where we start running out of resources. At that point, we have a tensor network with the MPS structure representing $\sigma_L$. Let us denote by $C^{n}_{\alpha,\beta}$ the basic tensor of that network, where $n=1,\ldots,D^2$ and $\alpha,\beta=1,\ldots,D^{2r}$ ($n$ denotes the index in the horizontal direction). (iii) When the bond indices $\alpha,\beta$ grow larger than some predetermined value, say $D_c\le D^{2r}$ we start approximating the tensor network by one with bond dimension $D_c$ as follows. We first construct the tensor $K_{\alpha,\alpha';\beta,\beta'}=\sum_n C^{n}_{\alpha,\beta}
\bar C^{n}_{\alpha',\beta'}$. Later on we will always deal with the case in which $K$ is hermitean (when considered as a matrix); if this is not the case, one can always choose a gauge where it is symmetric \cite{perezgarcia06}. We determine the eigenvector, $X_{\beta,\beta'}$, corresponding to the maximum eigenvalue of $K$, diagonalize $X$, consider the $D_c$ largest eigenvalues and build a projector onto the corresponding eigenspace. We then truncate the indices $\alpha$ and $\beta$ by projecting onto that subspace. (iv) We continue in the same vein until the truncated tensor structure converges, which corresponds to the limit $N_h\to\infty$. (v) We can do the same with $\sigma_R$ by going from right to left. For the examples studied below, $\sigma_L=\sigma_R=:\sigma_b=\sigma_b^T$, and thus we just have to carry out this procedure once.
\section{Numerical results for AKLT models}
We now investigate some particular cases. We concentrate on the AKLT model \cite{AKLT1,AKLT2}, whose ground state, $|\Psi\rangle$, can be exactly described by a PEPS with bond dimension $D=2$, as shown in Figs.~\ref{Fig:ladders} and and \ref{Fig:4legs}. The spin in the first and last column have $S=3/2$, whereas the rest have $S=2$. The AKLT
Hamiltonian is given by a sum of projectors onto the subspace of maximum
total spin across each nearest neighbor pair of spins,
\begin{equation}
\label{AKLT}
H_{\rm AKLT}=\sum_{<n,m>} P_{n,m}^{(s)}\ ,
\end{equation}
where $P_{n,m}^{(s)}$ is the projector onto the symmetric subspace of spins $n$ and $m$. This Hamiltonian is $su(2)$ and translationally invariant. This invariance is inherited by the virtual ancillas, and thus $\sigma_b$ and $H_b$ will also be. These symmetries can be used in the numerical procedures. Note that if $H_b$ has this symmetries and has short-range interactions, then since the ancillas have spin 1/2 (as $D=2$), it will be generically critical.
\begin{figure}[h]\begin{center}
\includegraphics[width=0.9\columnwidth]{aklt}\end{center}
\caption{(Color online)
(a) Ribbon made of two ($N_h=2$) coupled periodic S=1/2 Heisenberg chains (2-leg ladder).
(b) Groundstate of a 2-leg S=3/2 AKLT ladder. Each site is split into
three spins-1/2 (red dots). Nearest neighbor spins-1/2 are paired up into
singlet valence bonds. (c) PEPS representation for S=3/2 and S=2 sites
of AKLT wavefunctions in the valence bond (singlet) picture (for
connection to the "maximally entangled picture" see text). Open squares
stand for the $r^m_{\alpha_1,\alpha_2,\alpha_3}$ and
$A^m_{\alpha_1,\alpha_2,\alpha_3,\alpha_4}$ tensors defined in the text
and open circles correspond the to $2\times 2$ matrix $[0,1;-1,0]$.
}
\label{Fig:ladders}
\end{figure}
The lattice is bipartite. It is convenient to apply the operator
$\exp(i\pi S_y/2)$ to every spin on the B sublattice: this unitary
operator does not change the properties of $\rho_\ell$ but slightly
simplifies the description of the PEPS. Thus, we can write the AKLT
Hamiltonian as in (\ref{AKLT}) but with $P_{n,m}^{(s)}\to \tilde P_{n,m}
:= \exp(i\pi (f_n S_{y,n}+f_m S_{y,m}) P_{n,m}^{(s)}\exp(i\pi (f_n
S_{y,n}+f_mS_{y,m})$, with $f_n=0,1/2$ if the spin $n$ is in the A or B
sublattice, resp.
We will study finite $N_h$-legs ladders, as well as infinite square lattices.
We will start out in the next subsection with the simplest case of $N_h=2$. Note that for this particular case the subsystem we consider when we trace one of the legs is a spin chain itself, so that density operator $\rho_{\ell=1}$ already describes a 1-dimensional system and thus the physical spins already represent the boundary. In such a case, we do not need to resort to the PEPS formalism but we can also study other model Hamiltonians besides the AKLT one. For example, we will consider the $su(2)$-symmetric Heisenberg ladder Hamiltonian of $S=1/2$ [Fig. \ref{Fig:ladders}(a)]
\begin{equation}
\label{Heis}
H_{\rm Heis}=\sum_{<n,m>} J_{n,m} {\bf S}_n\cdot {\bf S}_m\, ,
\end{equation}
where the exchange couplings $J_{n,m}$ are parametrized by some angle
$\theta$, i.e.\ $J_{\rm leg}=\cos{\theta}$ ($J_{\rm rung}=\sin{\theta}$) for
nearest-neighbor sites $n$ and $m$ on the legs (rungs) of the ladder.
Although the ground state has no simple PEPS representation, it can be
obtained numerically by standard Lanczos exact diagonalization techniques
on finite clusters of up to $14\times 2$ sites ~\cite{poilblanc2010}. Similarly to the AKLT 2--leg ladder [Fig. \ref{Fig:ladders}(b)], it possesses a finite magnetic correlation length $\xi$ which diverges when $\theta\rightarrow 0$ (decoupled chain limit). The opposite limit $\theta=\pi/2$ ($\theta=-\pi/2$) corresponds to decoupled singlet (triplet) rungs (strictly speaking, with zero correlation length).
For infinite systems, we will also be interested in the behavior of $H_b$ along a quantum phase transition. To this aim, we will also consider a distorted version of the AKLT model, and define a family of Hamiltonians
\begin{equation}
H(\Delta)=\sum_{<n,m>} Q_n(\Delta)Q_m(\Delta) \tilde P_{n,m} Q_n(\Delta)Q_m(\Delta),
\end{equation}
where $Q_n(\Delta)=e^{-8\Delta S_{z,n}^2}$. Note that the Hamiltonian is translationally invariant and has $u(1)$ symmetry. As $\Delta$ increases, it penalizes (nematic) states with $S_z=0$, and thus the spins tend to
take their maximum value of $S_z^2$. As we will show, there exists a critical value of $\Delta$ where a quantum phase transition occurs.
\begin{figure}[h]
\begin{center}
\includegraphics[width=0.9\columnwidth]{rho}
\end{center}
\caption{(Color online)
(a) 4-leg ($N_h=4$) AKLT ladder on a cylinder partitioned (dotted green line) into two halves.
(b) Schematic representation of the density matrix $\sigma_b^2$ of a 4-leg ($N_h=4$, $\ell=2$) AKLT ladder. After being "cut" the two halves are "glued" together (physical indices are contracted).
}
\label{Fig:4legs}
\end{figure}
\subsection{2--leg ladders~:~comparison between AKLT and Heisenberg models}
Let us start out with the $su(2)$-symmetric $\Delta=0$ AKLT model in a
two-leg ladder configuration, where $\rho_\ell$ corresponds to state of
one of the legs; that is, we take $N_h=2$, $\ell=1$, and all spins have
$S=3/2$ as shown in Fig.~\ref{Fig:ladders}(b). The
Hamiltonian is gapped~\cite{AKLT1,AKLT2}, and the ground state is a PEPS with
bond dimension $D=2$. The tensors corresponding to the two legs, $l$ and
$r$, coincide and are given by $r^m_{\alpha_1,\alpha_2,\alpha_3}= \langle
s_m|\alpha_1,\alpha_2,\alpha_3\rangle$, where $\alpha_i=\pm 1/2$, and
$|s_m\rangle$ is the state in the symmetric subspace of the three spin 1/2
with $S_z|s_m\rangle=m|s_m\rangle$, $m=-3/2,-1/2,1/2,3/2$.
We first examine the entanglement spectrum of $H_b$ computed on a
$16\times 2$ ladder. It is shown on Fig.~\ref{Fig:ES}(b) as a function of
the momentum along the legs, making use of translation symmetry (the
vertical direction is periodic) enabling to block-diagonalize the reduced
density matrix in each momentum sector $K$. Note that it is also easy to implement the conservation of the $z$-component $S_z$ of the total spin so that each eigenstate can also be labelled according to its total spin $S$. The
low-energy part of the spectrum clearly reveals zero-energy modes at $K=0$
and $K=\pi$ consistent with conformal field theory of central charge
$c=1$.
\begin{figure}[h]
\begin{center}
\includegraphics[width=0.98\columnwidth]{ES_vsK_Sz}
\end{center}
\caption{(Color on line)
Entanglement spectra of $H_b$ (w.r.t. the groundstate energy $\xi_0$) versus total momenta $K$ in the chain (vertical) direction.
(a) 2--leg ($14\times 2$)
quantum Heisenberg ladder, (b) 2--leg ($16\times 2$) AKLT ladder and (c) 8--leg ($16\times 8$) AKLT ladder.
The eigenvalues are labelled according to their total spin quantum number using different symbols (according to
the legend on the graph).
}
\label{Fig:ES}
\end{figure}
It is of interest to compare the 2--leg AKLT results to the ones of the 2-leg S=1/2 Heisenberg ladder (\ref{Heis}) sketched in Fig.~\ref{Fig:ladders}(a) and investigated in Ref.~\onlinecite{poilblanc2010}.
Fig.~\ref{Fig:ES}(a) obtained on a $14\times 2$ ladder for a typical
parameter $\theta=\pi/3$ shows the entanglement spectrum of $\rho_\ell$
which, again, is very similar to that of a single nearest-neighbor
Heisenberg chain. As mentioned in Ref.~\onlinecite{poilblanc2010}, in
first approximation, varying the parameter $\theta$ (and hence the ladder
spin-correlation length) only changes the overall scale of the energy
spectrum. Hence, it has been suggested~\cite{poilblanc2010} to connect
this characteristic energy scale to an effective inverse temperature
$\beta_{\rm eff}$.
\begin{figure}[h]
\begin{center}
\includegraphics[width=0.9\columnwidth]{JvsTheta}
\end{center}
\caption{(Color on line)
(a) Amplitudes $A_r$ of the (isotropic) spin-spin couplings up to distance $r=7$ of the effective boundary Hamiltonian
of a quantum Heisenberg 2-leg ladder in the Haldane and rung singlet phases vs $\theta$.
(b) Ratio of the same amplitudes normalized to the nearest-neighbor coupling ($r=1$).
Computations are carried out on $12\times 2$ (open symbols) and $14\times 2$ (closed symbols) systems.
Note that when $\theta\rightarrow \pi/2$ (decoupled rung singlets), $A_r/A_0 \rightarrow 0$ for $r\ge 1$ and all the
weights of the reduced density matrix becomes equal to $2^{-N_v}$ ($A_0=\ln{2}$).
}
\label{Fig:JvsTheta}
\end{figure}
The above results strongly suggest that $H_b$ is "close" to a one-dimensional nearest-neighbor Heisenberg Hamiltonian. To refine this statement and make it more precise, we perform an expansion in terms of $su(2)$-symmetric
extended-range exchange interactions,
\begin{equation}
H_b=A_0 N_v +\sum_{r,k} A_r \, {\bf S}_k\cdot {\bf S}_{k+r} + R {\hat X} \, ,
\label{Eq:Hb}
\end{equation}
where $R{\hat X}$ stands for the "rest", i.e.\ (small) multi-spin
interactions. The amplitudes $A_r$ can be computed from simple trace
formulas,
\begin{equation}
A_r= \frac{4}{N_v}{\rm tr}\{H_b\sum_k \sigma^z_k \sigma^z_{k+r}\} / 2^{N_v} \, ,
\end{equation}
requiring the full knowledge of the eigenvectors of $H_b$ (i.e.\ of
$\sigma_b$). $A_0$ is fixed by some normalization condition, e.g. ${\rm
tr}\,\sigma_b=1$. Assuming $\hat X$ is normalized as an extensive
operator in $N_v$, i.e.\ $\frac{1}{N_v}{\rm tr}\{{\hat X}^2 \}= 2^{N_v}$,
the amplitude $R$ is given by:
\begin{equation}
R^2= \frac{1}{N_v}{\rm tr}\{H_b^2\}/2^{N_v}-N_v A_0^2-\frac{3}{16}\sum_{r=1}^{N_v/2} A_r^2 \, .
\end{equation}
The coefficients $A_r$ and R of 2--leg Heisenberg ladders are plotted in
Fig.~\ref{Fig:JvsTheta}(a) as a function of the parameter $\theta$, both
in the Haldane ($J_{\rm rung}<0$ i.e. ferromagnetic) and rung singlet
phases ($J_{\rm rung}>0$ i.e. antiferromagnetic).
Generically, we find that $H_b$ is {\it not frustrated}, i.e.\ all
couplings at odd (even) distances are antiferromagnetic (ferromagnetic),
$A_r >0$ ($A_r <0$). Clearly, the largest coupling is the
nearest-neighbor one ($r=1$). Fig.~\ref{Fig:JvsTheta}(b) shows the
relative magnitudes of the couplings at distance $r>1$ w.r.t.\ $A_1$.
These data suggest that the effective boundary Hamiltonian $H_b$ is short
range, especially in the strong rung coupling limit
($\theta\rightarrow\pi/2$) where $|A_{r^\prime}/A_r|\rightarrow 0$ for
$r^\prime > r$. The amplitude $A_1$ of the nearest-neighbor interaction
can be identified to the effective inverse temperature $\beta_{\rm eff}$
which, therefore, vanishes (diverges) in the strong (vanishing) rung
coupling limit.
\begin{figure}[h]
\begin{center}
\includegraphics[width=0.9\columnwidth]{JvsDist}
\end{center}
\caption{(Color on line)
2-leg quantum Heisenberg ladders -- Ratio of the amplitudes $|A_r|$ by the
nearest-neighbor amplitude $A_1$ plotted using a logarithmic scale as a
function of $r$ for different values of $\theta$. (a) Antiferromagnetic
and (b) ferromagnetic leg couplings (the rung couplings are
antiferromagnetic in both cases).}
\label{Fig:JvsDist}
\end{figure}
Next, we investigate the functional form of the decay of the amplitudes
$|A_r|$ with distance. The ratio $|A_r|/A_1$ versus $r$ are plotted
(using semi-log scales) in Figs.~\ref{Fig:JvsDist}(a,b) for $12\times 2$
and $14\times 2$ Heisenberg ladders with different values of $\theta$.
Similar data for a $20\times 2$ AKLT ladder is shown in
Fig.~\ref{Fig:JvsDist2}(a), providing clear evidence of {\it exponential}
decay of the amplitudes with distance, i.e. $|A_r|\sim\exp{(-r/\xi_b)}$. The
Heisenberg ladder data are also consistent with such a behavior (even
though finite size corrections are stronger than for the AKLT case,
especially when $\theta\rightarrow 0$ or $\pi$). It is not clear however
how deep the connection between the emerging length scale $\xi_b$ and the
2--leg ladder spin correlation length $\xi$ is. Note that the latter can
be related~\cite{poilblanc2010} to some effective {\it thermal} length
associated to the inverse temperature $\beta_{\rm eff}\propto A_1$.
Thanks to the PEPS representation of their ground state, AKLT ladders can be
(exactly) handled up to larger sizes than their Heisenberg counterparts
(typically up to $N_v=20$) enabling a careful finite size scaling analysis
of the boundary Hamiltonian (\ref{Eq:Hb}). As shown in
Fig.~\ref{Fig:JvsW}(a), we observe a very fast (exponential) convergence
of the coefficients $A_r$ with the ladder length $N_v$. Hence, one gets at
least 7 (3) digits of accuracy for all distances up to $r=5$ ($r=7$).
\begin{figure}[h]
\begin{center}
\includegraphics[width=0.9\columnwidth]{JvsDist2}
\end{center}
\caption{(Color on line)
AKLT ladders -- (a) Ratio $|A_r|/A_1$ plotted using a logarithmic scale
as a function of $r$. Results are approximation-free for finite $N_h$ while the $N_h\rightarrow\infty$ limit
is obtained by finite size scaling (see Fig.~\protect\ref{Fig:JvsW}(b)). (b) Comparison with $\sqrt{d_{r+1}/d_2}$ (full symbols) computed (see text) on 2--leg and infinitely long ($N_h=\infty$)
cylinder.}
\label{Fig:JvsDist2}
\end{figure}
In fact, as pointed out previously, the boundary Hamiltonian $H_b$ should not contain only two-body spin interactions.
However, the total magnitude of all left-over (multi-body) contributions, $R$, is remarkably small in the
AKLT 2--leg ladder~: as shown in Fig.~\ref{Fig:JvsW}(a), $R<A_4$.
In fact, the full magnitude of {\it all} many-body terms extending on $r+1$ sites is
given by $\sqrt{d_{r+1}}$ and can be compared directly to $|A_r|$ (after proper normalization).
Fig.~\ref{Fig:JvsDist2}(b) shows that $\sqrt{d_{r+1}/d_2}$ and $|A_r|/A_1$ are quite close, even at large distance.
Note however that multi-body interactions are significantly larger in the boundary Hamiltonian of
the Heisenberg ladder, as shown in Fig~\ref{Fig:JvsTheta} (although no
accurate finite size scaling analysis can be done in that case).
\subsection{$N_h$--leg AKLT ladder}
Now we consider the AKLT model on an $N_h$--leg ladder configuration; we
take $\ell=N_h/2$. The spins in the first and last legs have $S=3/2$, and the corresponding tensors coincide with the ones given above. The rest of the
spins have $S = 2$, and the corresponding tensor is $A^m_{\alpha_1,\alpha_2,\alpha_3,\alpha_4}= \langle
s_m|\alpha_1,\alpha_2,\alpha_3,\alpha_4\rangle$,
where $\alpha_i=\pm 1/2$, and $|s_m\rangle$ is the state in the symmetric
subspace of the four spin 1/2 with $S_z|s_m\rangle=m|s_m\rangle$,
$m=-2,-1,0,1,2$ (see Fig.~\ref{Fig:ladders}(c)). An example of a 4--leg ladder and of a schematic representation of $\rho_\ell$ is shown in
Fig.~\ref{Fig:4legs}.
\begin{figure}[h]
\begin{center}
\includegraphics[width=0.9\columnwidth]{JvsW}
\end{center}
\caption{(Color on line)
(a) Finite size scaling of the amplitudes $A_r$ for a 2-leg AKLT ladder vs $1/N_v$ ($N_v$=14, 16, 18, 20).
(b) Finite size scaling of the amplitudes $A_r$ for $N_h$-leg AKLT ladders vs $1/N_h$ at fixed
$N_v=16$ (open symbols) or $N_v=18$ (filled and $+$ symbols).
(c) VN entropy per unit length (normalized by $\ln(2)$) vs $1/N_h$ at fixed $N_v=16$.
}
\label{Fig:JvsW}
\end{figure}
Let us now follow the same analysis (\ref{Eq:Hb}) of the boundary
Hamiltonian as we did for the case of 2 legs. The decay with distance of
the coefficients $A_r$ are reported in Fig.~\ref{Fig:JvsDist2}(a) for
4--leg, 6--leg and 8--leg AKLT ladders. Clearly, the decay is still
exponential with distance for all values of $N_h$ studied but the
characteristic length scale associated to this decay (directly given by
the inverse of the slope of the curve in such a semi-log plot) smoothly
increases with $N_h$. A careful finite size scaling is performed in
Fig.~\ref{Fig:JvsW}(b) to extract the $N_h\rightarrow\infty$ limit of all
$A_r$ (accurate up to $r=7$). The extrapolated values are reported in
Fig.~\ref{Fig:JvsDist2}(a) showing that $A_r$ also decays exponentially
fast with $r$ in an infinitely long cylinder ($N_h=\infty$). The
characteristic emerging length scale is estimated to be still very short
around $1$.
Lastly, we compute the Von Neumann entanglement entropy defined by $S_{\rm VN}(\rho_\ell)=-{\rm tr} \{\rho_\ell \ln{\rho_\ell}\}$
with the normalization ${\rm tr}\,\rho_\ell=1$.
$S_{\rm VN}$ scales like $N_v$ ("area" law) and is bounded by $N_v \ln{2}$. Fig.~\ref{Fig:JvsW}(c) shows that the entropy
converges very quickly with $N_h$ to its thermodynamic value which is very close to the maximum value.
The entanglement of the two halves of the AKLT cylinder is therefore very strong.
\subsection{Thermodynamic limit and phase transitions}
Now we consider the $N_v,N_h\to\infty$ for the deformed AKLT model in order to investigate the phase transition. We will compare some of the results with the $2$--leg ladder as well. The spins in the first and last legs have $S=3/2$, and the rest $S=2$. The corresponding tensors are defined according to
\begin{eqnarray}
l^m_{\alpha_1,\alpha_2,\alpha_3} &=& r^m_{\alpha_1,\alpha_2,\alpha_3}= \langle s_m|Q(\Delta)|\alpha_1,\alpha_2,\alpha_3\rangle,\nonumber\\
A^m_{\alpha_1,\alpha_2,\alpha_3,\alpha_4}&=& \langle
s_m|Q(\Delta)|\alpha_1,\alpha_2,\alpha_3,\alpha_4\rangle,
\end{eqnarray}
where $\alpha_i=\pm 1/2$, and $|s_m\rangle$ is the state in the symmetric
subspace of the three (four) spin 1/2 with $S_z|s_m\rangle=m|s_m\rangle$,
$m=-3/2,-1/2,1/2,3/2$ ($m=-2,-1,0,1,2$), respectively.
We will use the approximate procedure sketched in Section III-C. In particular, for $N_v$ larger than the correlation length the obtained tensors $C^{n}_{\alpha,\beta}$ will be independent of $N_v$. We have considered those tensors (with $D_c=50$ and 100 iterations), and built $\sigma_b$ and $H_b$ out of them. Note that the $su(2)$ symmetry is explicitly broken by a finite $\Delta$ so that it becomes more convenient to use the variable $d_n$ of Eq.~(\ref{Eq:dn}) instead of $A_r$ to probe the spatial extent of $H_b$. We recall that $(d_n)^{1/2}$ is the
mean amplitude of {\it all} interactions acting at distance $r=n-1$. We
have plotted in Fig.~\ref{Fig:dn} all $d_n$, $n\le N_v/2$, for $N_v=16$ as
a function of $\Delta$. As $\Delta$ increases, we see that the
interaction length of the effective Hamiltonian increases and one sees a
long-range interaction appearing. This indicates that we approach a phase
transition. For the case of the ladder, the interaction length remain
practically constant for the same range of variation of $\Delta$.
\begin{figure}[h]
\begin{center}
\includegraphics[width=0.9\columnwidth]{dn_vsDist}
\end{center}
\caption{(Color on line)
AKLT model with finite "nematic" field $\Delta$ -- (a) Relative amplitude $\sqrt{d_{r+1}/d_2}$ in a 2--leg ladder
plotted using a logarithmic scale as a function of $r$; (b) same for an infinitely long
cylinder ($N_h=\infty$). From bottom to top, $\Delta$ is incremented from $0$ to $\Delta_{\rm max}$
by constant steps.
(c) Inset~: effective temperature $\beta_{\rm eff}$ (see Eq.~\protect\ref{Eq:beta}) versus $\Delta$ for the two cases reported in (a) and (b).
All results are obtained for $N_v=16$ ($D_c=50$, and 100 iterations such that the tensors $C$ already converge). }
\label{Fig:dn}
\end{figure}
Similarly to the investigation of the Heisenberg ladder~\cite{poilblanc2010}, it is interesting to define an
effective inverse temperature via the amplitude of the nearest-neighbor interaction,
\begin{equation}
\beta_{\rm eff}=8\sqrt{\frac{d_2}{3}}\, ,
\label{Eq:beta}
\end{equation}
where the pre-factor is introduced conveniently so that $\beta_{\rm eff}=A_1$ in the $su(2)$-symmetric limit $\Delta=0$.
As seen in the inset of Fig.~\ref{Fig:dn}, the inverse temperature of the ladder scales linearly with $\Delta$.
For the infinite cylinder, no singularity of $\beta_{\rm eff}$ is seen at the cross-over between short and long-range
interactions.
\begin{figure}[h]
\begin{center}
\includegraphics[width=0.9\columnwidth]{xi_vs_Delta}
\end{center}
\caption{(Color online)
(a) Inverse correlation length $\xi^{-1}$ vs $\Delta$ for both 2-leg ladder and infinitely long cylinder ($N_h=\infty$).
These data correspond to the infinite circumference limit, i.e. $N_v=\infty$. The arrow marks the phase transition in the infinitely
long cylinder.
Comparison with the inverse of the "emerging" length scale $\xi_b$ obtained by fitting the decay of the coefficients of $H_b$
plotted in Fig.~\ref{Fig:dn}(a) as $\sqrt{d_{r+1}/d_2}\sim \exp{(-r/\xi_b)}$.
(b) Truncation error in the $N_h\rightarrow\infty$ procedure. The results are compared with those obtained with $D_c=150$, and 100 iterations have always been used.
}
\label{Fig:xi}
\end{figure}
Next, we plot the inverse correlation length as a function of $\Delta$
both for one dimension (i.e.\ an infinitely long ladder) and for two
dimensions (i.e.\ $N_v=N_h=\infty$) in Fig.~\ref{Fig:xi}(a), obtained with $D_c=150$ and 100 iterations (no difference are observed by taking $D_c=50$ and 50 iterations). Clearly, the divergence of $\xi$ (i.e.\ $\xi^{-1}=0$) shows the appearance of a phase transition at $\Delta=0.0061$ in two dimensions. In contrast, $\xi^{-1}$ never crosses zero in the case of the ladder (i.e.\ in one dimension). We have compared $\xi$ with the "emerging" length scale $\xi_b$ obtained by fitting the decay of the coefficients of $H_b$ as $\sqrt{d_{r+1}/d_2}\sim \exp{(-r/\xi_b)}$ on $N_v=16$ 2-leg and infinitely long (i.e.\ $N_h=\infty$) cylinders. In the two leg ladder, we see that the
divergence of the correlation length $\xi$ for $\Delta\rightarrow\infty$
results from the interplay between (i) a (moderate) increase of the range
$\xi_b$ of the Hamiltonian $H_b$ and (ii) a linear increase with $\Delta$
of the effective temperature scale $\beta_{\rm eff}$, therefore approaching
the $T_{\rm eff}\rightarrow 0$ limit when $\Delta\rightarrow\infty$. This
contrasts with the case of two dimensions ($N_v=N_h=\infty$) where the
divergence of $\xi$ occurs at {\it finite effective temperature} when
$H_b$ becomes "sufficiently" long-range. It is however hazardous to fit
the decay of the coefficients of $H_b$ to obtain its functional form at
the phase transition. Finally, in Fig.~\ref{Fig:xi}(b) we have plotted
the truncation error made by taking different $D_c$ in the limit
$N_h\rightarrow\infty$, and, again around $\Delta\approx 0.006$ the error increases. This is consistent with the expectation that as $H_b$ contains longer range interaction, the boundary density operator $\sigma_b$ requires a higher bond dimension to be described as a TN state.
\section{Numerical results for Ising PEPS}
We now continue by considering the Ising PEPS introduced in \cite{verstraetewolf06}. They all have bond dimension $D=2$ and
exhibite the $\mathbb Z_2$-symmetry of the transverse Ising chain. They depend on a single parameter, $\theta\in [0,\pi/4]$. For $\theta\sim \pi/4$ one has a state with all the spins pointing in the $x$ direction, whereas for
$\theta\sim 0$ the state is of the GHZ type (a superposition of all spins up and all down). In the thermodynamic limit ($N_v,N_h\to\infty$) for $\theta\approx 0.35$ they feature a phase transition, displaying critical behavior, where the correlation functions decay as a power law. Thus, by changing $\theta$ we can investigate how the boundary Hamiltonian behaves as one approaches the critical point.
\subsection{2--leg ladders}
The tensors corresponding to the two legs, $l$ and $r$, coincide and are given by $r^m_{\alpha_1,\alpha_2,\alpha_3}=
a_m(\alpha_1)a_m(\alpha_2)a_m(\alpha_3)$, where $m=0,1$, $\alpha_i=\pm
1/2$ and $a_m(\alpha)$ are parametrized as
$a_0(-1/2)=a_1(1/2)=\cos{\theta}$ and $a_0(1/2)=a_1(-1/2)=\sin{\theta}$.
\begin{figure}[h]
\begin{center}
\includegraphics[width=0.9\columnwidth]{ES_vsK_Ising}
\end{center}
\caption{
Entanglement spectrum of a $16\times 2$ Ising PEPS ladder versus momentum along
the ladder leg direction. Comparison between $\theta=0.2$ (a) and
$\theta=0.5$ (b) using the same energy scale. (c) Zoom in of the low-energy part of (b). }
\label{Fig:ES_Ising}
\end{figure}
As seen in Fig.~\ref{Fig:ES_Ising} the entanglement spectrum of the 2-leg
ladder is gapped for all $\theta$ values and resemble the one of an Ising
chain (equally spaced levels) with small quantum fluctuations revealed by
the small dispersion of the bands. The effective inverse temperature,
qualitatively given by the gap (or the spacing between the bands),
decreases for increasing $\theta$.
The interaction length of the boundary Hamiltonian for the ladder is displayed in Fig.~\ref{Fig:dnIsing}(a). The strength of the interactions decay exponentially with the distance for all values of $\theta$. As we increase this angle, one only observes a decrease of the interaction length. Note that as opposed to the AKLT models studied in the previous sections, $d_1\ne 0$. Indeed, there always exists a term with a single Pauli operator $\sigma_x$, describing an effective transverse field in $H_b$. Thus, that Hamiltonian is given by a transverse Ising chain in the non--critical region of parameters.
We have also plotted the inverse correlation length $\xi^{-1}$ as a function of $\theta$ in Fig.~\ref{Fig:xi_Ising} (blue empty dots). While the correlation length increases as $\theta$ decreases, it only tends to infinite in the limit $\theta\to 0$, as it must be for a GHZ state. No signature of a phase transition is found otherwise.
\subsection{Thermodynamic limit and phase transitions}
We now move to the case of an infinitely long cylinder. As above to grow the cylinder in the horizontal direction, one
considers rank-5 tensors, which here take the form $A^m_{\alpha_1,\alpha_2,\alpha_3,\alpha_4}= a_m(\alpha_1)a_m(\alpha_2)
a_m(\alpha_3)a_m(\alpha_4)$, and use the same approximation scheme with 100 iterations as before.
The parameters $d_n$ describing the boundary Hamiltonian $H_b$ behave very differently in the ladder and infinite cylinders as shown in Fig.~\ref{Fig:dnIsing}. While for the Ising PEPS ladder
$H_b$ remains short-ranged with exponential decay of $d_n$ vs $n$, the infinite cylinder shows a transition towards long-range interactions suggesting the existence of a phase transition. This is very similar to what occurred in the AKLT distorted model.
\begin{figure}[h]
\begin{center}
\includegraphics[width=0.9\columnwidth]{dn_vsDist_Ising}
\end{center}
\caption{(Color online)
Ising PEPS -- Relative amplitude $\sqrt{d_{n}/d_1}$ in a 2--leg ladder (a)
and in an infinitely long ($N_h=\infty$) cylinder (b) as a function of $n$ for $\theta$ varying from 0.1 to $\sim 0.51$ with constant intervals
(0.1, 0.1585, 0.2171, 0.2756, 0.3342, 0.3927, 0.4512 and 0.5098, from top to bottom).
Logarithmic scales are used on the vertical axis in both (a) and (b).
Inset: Ratio of the effective transverse field $\sqrt{d_1}$ over the effective Ising nearest-neighbor coupling $\sqrt{d_2}$ versus
$\theta$ for $N_h=\infty$ ($D_c=50$ and 100 iterations). All results are obtained for $N_v=12$. }
\label{Fig:dnIsing}
\end{figure}
As long as $H_b$ remains short-range, the density matrix $\rho_\ell$ can be
(qualitatively) mapped onto the thermal density matrix of an effective
quantum Ising chain (including a "family" of transverse-like fields) and,
therefore, no ordering is expected (at finite effective temperature). A
phase transition however can appear when $H_b$ becomes long-ranged as it
is the case for an infinitely long cylinder. This is evidenced by the the
behavior of correlation lengths computed for the 2-leg and infinitely long
($N_h=\infty$) cylinder and reported in Fig.~\ref{Fig:xi_Ising}. These
correlation lengths are compared to the respective "emerging" length
scales $\xi_b$ characterizing the decay of $\sqrt{d_{n}}$ with $n$. In
the 2-leg ladder case, $\xi_b$ increases quite moderately when
$\theta\rightarrow 0$ ($\xi_b\sim 1$) so that the divergence of the
correlation length $\xi$ in this limit is only attributed to a {\it
vanishing} of the effective temperature scale $T_{\rm eff}$. In contrast,
as for the AKLT model, the phase transition in two dimensions occurs at
finite (effective) temperature at the point where
$\xi_b\rightarrow\infty$.
\begin{figure}[h]
\begin{center}
\includegraphics[width=0.9\columnwidth]{xi_vs_Delta_Ising}
\end{center}
\caption{(Color on line)
Ising PEPS --
Inverse correlation length $\xi^{-1}$ vs $\Delta$ for both 2-leg ladder and infinitely long cylinder ($N_h=\infty$, $D_c=150$ and 100 iterations).
These data correspond to the infinite circumference limit, i.e. $N_v=\infty$. The arrow marks the phase transition in the infinitely
long cylinder.
Comparison with the inverse of the "emerging" length scale $\xi_b^{-1}$ obtained by fitting the decay of
the coefficients plotted in Fig.~\protect\ref{Fig:dnIsing} as $\sqrt{d_{n}}\sim \exp{(-n/\xi_b)}$. }
\label{Fig:xi_Ising}
\end{figure}
In summary, these results evidence that whenever we approach a phase transition, the interaction length of the boundary Hamiltonian increases.
\section{Topological Kitaev code}
Let us finally consider systems with topological order. We will focus on
Kitaev's code state~\cite{kitaev:toriccode}: It can be defined on a square
lattice with spin-$\tfrac12$ systems (qubits) on the vertices, with two
types of terms in the Hamiltonian,
\begin{equation}
\label{eq:tcode-terms}
h_X = X^{\otimes 4}\ , \quad
h_Z= Z^{\otimes 4}
\end{equation}
(where $X$ and $Z$ are Pauli matrices), each of which acts on the four
spins adjacent to a plaquette, and where the $h_X$ and $h_Z$ form a
checkerboard pattern (see Fig.~\ref{Fig:kitaev}(a)). The ground state subspace of the code state can be
represented by a PEPS with $D=2$~\cite{verstraetewolf06}; a particularly
convenient representation is obtained by taking $2\times 2$ blocks of
spins across $h_Z$ type plaquettes, and jointly describing the spins in
each block by one tensor of the form~\cite{schuch:peps-sym}
\begin{equation}
\label{eq:tcode-tensor}
A^{i_{1,2},i_{2,3},i_{3,4},i_{4,1}}_{\alpha_1,\alpha_2,\alpha_3,\alpha_4}
= \left\{\begin{array}{ll}
1 &\mathrm{if\ } i_{x,x+1} = \alpha_{x+1} - \alpha_{x}
\ \mathrm{mod}\,2 \ \forall\, x\\
0 & \mathrm{otherwise.}
\end{array}\right.
\end{equation}
Here, $i_{x,x+1}$ denotes the spin located between the bonds $\alpha_x$
and $\alpha_{x+1}$ (numbered clockwise) as shown in Fig.~\ref{Fig:kitaev}(b). It can be checked
straightforwardly that the resulting tensor network is an eigenstate of
the Hamiltonians of Eq.~(\ref{eq:tcode-terms}). Excitations of the model
correspond to violations of $h_X$-terms (charges) or $h_Z$-terms (fluxes),
which always come in
pairs~\cite{kitaev:toriccode}.
We put the code state on a cylinder of $N_h\times N_v$ tensors (i.e.,
$2N_h\times 2N_v$ sites), where we choose boundary conditions
\begin{equation}
\label{eq:kitaev-bnd}
|\chi_\theta) = \cos\tfrac\theta2\,|0)^{\otimes N_v}+
\sin\tfrac\theta2\,|1)^{\otimes N_v}\ .
\end{equation}
This yields a state which is also a ground state of
$h_Z^b=Z^{\otimes 2}$ terms at the boundary, but not of the corresponding
$X^{\otimes 2}$ boundary terms; in other words, charges (Pauli $Z$ errors)
can condense at the boundaries of the cylinder~\cite{dennis}.
The full Hamiltonian---including the $h_Z^b$ terms at the boundary---has a
two-fold degenerate ground state which is topologically protected, and
where the logical $X$ and $Z$ operators are a loop of Pauli $X$'s around
the cylinder and a string of Pauli $Z$'s between its two ends (where they
condense), respectively.
To compute $\rho_\ell$, we start by considering the PEPS on the cylinder
without the boundary conditions (\ref{eq:kitaev-bnd}), i.e., with open
virtual indices at both ends (labelled $B$ and
$B'$). Cutting the cylinder in the middle leaves us with with
$\sigma_{BL}$, the joint reduced density operator for the virtual spaces
at the boundary, $B$ (or $B'$), and the cut, $L$ (or $R$). From
(\ref{eq:tcode-tensor}), one can readily infer that the transfer operator
for a single tensor is $\openone^{\otimes 4}+X^{\otimes 4}$, and thus,
\begin{equation}
\label{eq:tcode-1}
\sigma_{BL}=\sigma_{B'R} \propto
\openone^{\otimes N_v}\otimes\openone^{\otimes N_v} +
X^{\otimes N_v}\otimes X^{\otimes N_v}\ ,
\end{equation}
where the two tensor factors correspond to the $B$ ($B'$) and $L$ ($R$)
boundary, respectively. Imposing the boundary condition
$|\chi_\theta)(\chi_\theta|$, Eq.~(\ref{eq:kitaev-bnd}), at $B$ ($B'$),
we find that (up to normalization)
\[
\rho_\ell \propto
(1+\sin^2\theta)\,\openone^{\otimes N_v} +
(2\sin\theta)\, X^{\otimes N_v}\ ,
\]
which is the thermal state $\rho_\ell\propto \exp[-\beta_\mathrm{eff}
H_\ell]$ of $H_\ell=-\mathrm{sign}(\sin\theta)\,X^{\otimes N_v}$ at
an effective inverse temperature
\[
\beta_\mathrm{eff} = \left\vert
\tanh^{-1}\left[\frac{2\sin\theta}{1+\sin^2\theta}\right]
\right\vert\ .
\]
The fact that $H_\ell$ acts globally is a signature of the topological
order, and comes from the fact that measuring an $X$ loop operator gives a
non-trivial outcome (namely $\sin\theta$). Note that the entropy
$S(\rho_\ell)$ increases by one as $1/\beta_\mathrm{eff}$ goes from zero
to infinity. This can be understood as creating an entangled pair of
charges $\ket{\mathrm{vac}}+f(\beta_\mathrm{eff})\ket{c,c^*}$ across the
cut, thereby additionally entangling the two sides by at most an ebit, and
subsequently condensing the charges at the boundaries.
Instead of considering $\sigma_L$, one can also see the topological order
by looking at $\sigma_{BL}$: It is the zero-temperature state of a
completely non-local Hamiltonian $X^{\otimes N_v}\otimes X^{\otimes N_v}$
which acts simultaneously on both boundaries in a maximally non-local way;
this relates to the fact that the expectation values of any two $X$ loop
operators around the cylinder are correlated.
Let us point out that systems with conventional long-range order behave
quite differently, even though they also exhibit correlations between
distant boundaries. Consider the spin-$\tfrac12$ Ising model
without field, which has a PEPS tensor
\[
A^i_{\alpha_1,\alpha_2,\alpha_3,\alpha_4}=
\delta_{i,\alpha_1}\delta_{\alpha_1,\alpha_2}
\delta_{\alpha_2,\alpha_3}\delta_{\alpha_3,\alpha_4}\ .
\]
The resulting local transfer operator is $|0)(0|^{\otimes 4}+|1)(1|^{\otimes
4}$, and thus,
\[
\sigma_{BL}=
|0)(0|^{\otimes N_v} + |1)(1|^{\otimes N_v}\ .
\]
By imposing boundary conditions at $B$, one
arrives at
\[
\rho_\ell =
\sin\theta\,|0)(0|^{\otimes N_v}
+ \cos\theta\,|1)(1|^{\otimes N_v}\ ,
\]
which is the thermal state of of the classical Ising Hamiltonian
\[
H(\beta) = -\sum_{i} Z_i Z_{i+1} -
\frac{\log\tan\theta}{2\beta N_v} \sum_i Z_i
\]
for $\beta\rightarrow\infty$. Thus, for the Ising model, $\rho_\ell$
is described by a local Ising Hamiltonian, rather than a completely
non-local interaction as for Kitaev's code state. The same holds true for
$\sigma_{BL}$, which is the ground state of a classical Ising model
without field: while it has correlations between the two boundaries, they
arise from a local (i.e., few-body) interaction coupling the two
boundaries, rather than from terms acting on \emph{all} sites on both
boundaries together. Correspondingly, the long-range correlations in the
Ising model can be already detected by measuring local observables, instead of
topologically nontrivial loop operators as for Kitaev's code state.
\begin{figure}[h]
\begin{center}
\includegraphics[width=0.9\columnwidth]{kitaev}
\end{center}
\caption{(Color on line)
(a) Checkerboard decomposition in the Kitaev code. Spin-$\tfrac12$ are represented by (red) dots
at the vertices of the square lattice. $X$ and $Z$ operators act on the 4 spins of each type of (shaded and non-shaded) plaquettes.
(b) PEPS representation of the Kitaev code (see text).}
\label{Fig:kitaev}
\end{figure}
\section{Conclusions and outlook}
In this paper, we have introduced a framework which allows to associate
the bulk of a system with its boundary in the spirit of the holographic
principle. To this end, we have employed the framework of Projected
Entangled Pair States (PEPS) which provide a natural mapping between the
bulk and the boundary, where the latter is given by the virtual degrees of
freedom of the PEPS. This framework allows to map the state of any region
to a Hamiltonian on its boundary, in such a way that the properties of the
bulk system, such as entanglement spectrum or correlation length, are
reflected in the properties of the
Hamiltonian. Since our framework also identifies observables in the bulk
with observables on the boundary, it establishes a general holographic
principle for quantum lattice systems based on PEPS.
In order to elucidate the connection between the bulk system and the
boundary Hamiltonian, we have numerically studied the AKLT
model and the Ising PEPS. We found that the Hamiltonian is
local for systems in a gapped phase with local order, whereas a diverging
interaction length of the Hamiltonian is observed when the system
approaches a phase transition, and topological order is reflected in
a Hamiltonian with fully non-local interactions; thus, the quantum phase
of the bulk can be read off the properties of the boundary
model.
Our holographic mapping between the bulk and the boundary in the PEPS
formalism has further implications.
In particular, the contraction of PEPS in
numerical simulations requires to approximate the boundary operator by one
with a smaller bond dimension, which can be done efficiently if the boundary
describes the thermal state of a local Hamiltonian, i.e., for non-critical
systems. Also, since renormalization in the PEPS formalism requires to
discard the degrees of freedom in the bond space with the least
weight~\cite{levinwen},
the duality allows to understand real space renormalization in the bulk
as Hamiltonian renormalization on the boundary.
Our techniques can also be applied to systems in higher dimensions, and in
fact to arbitrary graphs, to relate the boundary of a system with its bulk
properties. The mapping applies to arbitrary regions in the lattice, such
as simply connected (e.g., square) regions used for instance for the
computation of topological entropies. Also, relating the bulk to the
boundary using the PEPS description can be generalized beyond spin systems
by considering fermionic or anyonic PEPS~\cite{fpeps}, as well as
continous PEPS in the case of field theories~\cite{cmps1,cmps2}.
Finally, when studying edge modes, the
one-dimensional system which describes the physical
boundary is given by a Matrix Product Operator acting on the virtual
boundary state, and thus, the relation between bulk properties and the
virtual boundary implies a relation between the properties of the bulk and
its edge modes physics.
\section*{Acknowledgements}
We acknowledge the hospitality of the Kavli Institute for Theoretical Physics (UC Santa Barbara, USA) where this work was initiated.
DP acknowledges support by the French Research Council (Agence Nationale de la Recherche) under grant No.~ANR~2010~BLANC~0406-01 and thanks IDRIS (Orsay, France) and
CALMIP (Toulouse, France) for the use
of NEC-SX8 and Altix SGI supercomputers, respectively.
NS acknowledges support by the Gordon and Betty Moore Foundation through
Caltech's Center for the Physics of Information, the NSF Grant
No.~\mbox{PHY-0803371}, and the ARO Grant No.~W911NF-09-1-0442.
FV acknowledges funding from the SFB projects Vicom and Foqus and the EC projects QUERG and Quevadis. JIC acknowledges the EC project Quevadis, the DFG Forschergruppe 635, and Caixa Manresa.
|
\section{\bf Introduction}
The \emph{\.{I}n\"on\"u-Wigner contraction} (IW) \cite{Inonu:1953sp} (see \cite{WeimarWoods} for the generalizations) plays a very important r\^ole in physics.
To see the most basic example of the contraction, start with the $SO(3)$ algebra of $3d$ rotations. The algebra
consists of three operators $J_i$ satisfying the commutator relations:
\begin{equation}
[J_2, J_3] = i J_1 \, , [J_2, J_3] = i J_1 \quad \textrm{and} \quad [J_1, J_2] = i J_3 \, .
\end{equation}
Rescaling the operators $J_1$ and $J_2$:
\begin{equation} \label{WeylRescaling}
J_{i=1,2} = \tilde{J}_{i=1,2}/\sigma \quad J_{3} = \tilde{J}_{3}
\end{equation}
and taking the $\sigma \to 0$ limit one ends up with a new algebra:
\begin{equation}
\label{tildeJJJ}
[\tilde{J}_2, \tilde{J}_3] = i \tilde{J}_1 \, , [\tilde{J}_3, \tilde{J}_1] = i \tilde{J}_2
\quad \textrm{and} \quad [\tilde{J}_1, \tilde{J}_2] = 0 \, ,
\end{equation}
which is the algebra of $ISO(2)$, the group of the $\mathbb{R}^2$ isometries, where $\tilde{J}_{1,2}$
and $\tilde{J}_{3}$ are the translations and the rotation respectively. One can think of this contraction
as of taking to infinity the radius of the 2-sphere, while zooming near a point on it.
Probably the most famous physical application of the IW contraction
is the \emph{non-relativistic} limit of the Poincar\'e algebra. As we will briefly review
in the paper, the contraction parameter is the speed of light $c$ and the rescaled operators are
the time translation (the Hamiltonian) and the boosts. By the construction, however, the $c \to \infty$
limit leads to the Galilean algebra, which has exactly the same number of generators
as the original Poincar\'e algebra. For physical applications, though, it is necessarily to \emph{centrally
extend} the non-relativistic algebra by including the mass parameter $M$. The $M \neq 0$ version of the
Galilean algebra is commonly known as the Bargmann algebra \cite{Bargmann:1954gh}.
In fact, the Poincar\'e algebra is itself a contraction of the de Sitter algebra $SO(4,1)$,
where the contraction parameter now is
the cosmological constant $\Lambda$.
As was shown \cite{Bacry:1968zf} by Bacry and L\'evy-Leblond in 1967, under some physical assumptions (like parity,
time-reversal, \emph{etc.}) the full list of possible kinematic groups consists of the (anti) de Sitter groups and their
IW contractions. For example, one of the possibilities is the $(c, \Lambda) \to (\infty,0)$ contraction
of the de Sitter algebra with the parameter $\omega = c \sqrt{\Lambda}$ kept fixed.
The contraction produces the so-called Newton-Hooke (NH) algebra,
which further reduces to the Galilean algebra for $\omega=0$.
One can also take the \emph{ultra-relativistic} ($c \to 0$) limit of the Poincar\'e algebra arriving at the Carrollian algebra \cite{Levy-Leblond1965,SenGupta1966}, the much less studied counterpart of the Galilean algebra.
In quantum mechanics we are interested in general in \emph{projective} representations of groups.
On the other hand, a projective representation of a group $G$ is essentially equivalent to
a regular representation of the central extension of $G$. For this reason
central extensions are of special importance in quantum physics. The canonical (and the simplest)
example is the Weyl algebra:
\begin{equation}
\left[q, p\right] = i \hbar \, .
\end{equation}
Unlike the Galilean group, both the Poincar\'e algebra has no central extensions
except for the two dimensional case.\footnote{
The Galilean and the NH algebra have an additional, the so called \emph{exotic},
central extension in $2+1$ dimensions \cite{LevyLeblond:1971} as we will describe in the paper.}
This can be seen from the group cohomology arguments (see \cite{deAzcarraga:1995jw} for a detailed review).
These groups, although, still have plenty of \emph{non-central}
extensions, some of which are interesting from the physical point of view, \emph{e.g.}
the Maxwell algebra \cite{Bacry:1970ye,Schrader:1972zd}, a specific non-central extension of the Poincar\'e algebra.
The main goal of this paper is to show that starting from an IW contraction one can straightforwardly
find an extension of the contracted algebra.
Although the proposed procedure explicitly yields an infinite extension, it can be easily truncated at any level.
It works both for central and non-central extensions, and we find a criterion for an extension to be central.
Our first example is a IW contraction of $SO(3)$
given by the rescaling:
\begin{equation}\
\label{SO3scaling}
J_{1,2} = \tilde{J}_{1,2}/\sigma^2 \quad J_{3} = \tilde{J}_{3}/\sigma^3 \, .
\end{equation}
For $\sigma \to 0$ we find an abelian algebra (all the commutators vanish), but as we demonstrate in the paper
the first level extension dictated by this contraction reproduces the central extension of the Weyl algebra.
The higher level extensions, however, are not anymore central.
The organization of the paper is as follows. In the next section we present our method for constructing
algebra extensions from its \.{I}n\"on\"u-Wigner contraction. We then describe various applications,
like the non-central extension of the Poincar\'e algebra from the de Sitter algebra contraction,
the Galilean and the Carrollian contractions of the Poincar\'e algebra and the exotic extension of the Galilean and the NH algebra.
We end up with a short list of open questions.
\section{\bf The method}
\subsection{Lie algebra bundles}
For our construction we will need the notion of bundles of Lie algebras \cite{LieAlgebraBundle} (see also \cite{Weinstein} for a pedagogical introduction).
A \emph{Lie algebra bundle} (Lie bundle for short) is a vector bundle for which each fiber
has a smoothly varying Lie algebra structure.\footnote{
We will deal exclusively with Lie algebras and not with Lie groups.
Let us only notice here that every bundle of Lie groups defines a bundle of Lie algebras, and
every bundle of algebras can be integrated to a bundle of groups.
The group bundle, however, is not necessarily Hausdorf (see \cite{Weinstein}).}
More explicitly, the vector bundle $(\mathcal{E}, \pi, \mathcal{S})$ should be equipped with a morphism
$\theta : \, \mathcal{E} \otimes \mathcal{E} \to \mathcal{E}$, which induces a Lie algebra structure on
each fiber $\mathcal{E}_\sigma$. Here $\mathcal{E}$ is the total space, $\pi$ is the projection map and $\sigma$ is
a point on the base $\mathcal{S}$. This definition is sometimes called in the literature a \emph{weak}
Lie algebra bundle, in contrast to the \emph{strong} one which requires also
local triviality of the Lie structure.
It means that for any $\sigma \in \mathcal{S}$ there exists a neighborhood $U_\sigma$ of $\sigma$, a Lie algebra $g$
and a morphism $\phi : \, g \times U_\sigma \to \pi^{-1} (U_\sigma)$ such that
$\left. \phi \right\vert_{\sigma^\prime} : \, g \to \pi^{-1} (\sigma^\prime)$
is a Lie algebra isomorphism for each $\sigma^\prime \in U_\sigma$.
We will refer to $\phi$ as a Lie bundle trivialization.
If $\phi$ extends to the entire base, the Lie algebra bundle will be called \emph{trivial}.
Next, let us restrict our attention to weak Lie bundles, where the base $\mathcal{S}$ is just an affine
line\footnote{Our discussion can be easily generalized to complex numbers.} $\mathbb{R}^1$.
In this case any Lie algebra bundle will be trivial as a vector bundle, but not
necessarily so from the Lie algebra point of view. The Lie algebra $g$ at any fiber is the same
as a vector space, but its brackets are in general $\sigma$-dependent.
If $a^i$ are the generators of $g$ and $a_i^\sigma=(a_i,\sigma)$ denotes a point on the Lie bundle,
then the most general form of the commutators
is:
\begin{equation}
\left[ a_i^\sigma, a_j^\sigma \right] = f^k_{ij} (\sigma) a_k^\sigma \, ,
\end{equation}
where $f^k_{ij} (\sigma)$ are some smooth functions.
Let us consider the following example.
Assume that $i$ runs from $1$ to $3$ and:
\begin{equation}
\label{f-ijk}
f^k_{ij} (\sigma) = \frac{1}{2} \epsilon^k_{ij} \cdot f(\sigma) \, .
\end{equation}
If $f(\sigma) \neq 0$ for any $\sigma \in \mathbb{R}^1$, then the Lie bundle is trivial
(namely strong) with the trivialization $\phi$ given by:
\begin{equation}
\label{fsigma}
\phi(a_i,\sigma)= \frac{a_i^\sigma}{f(\sigma)} \,
\end{equation}
and $a_i$'s identified with the $SO(3)$ generators $J_i$'s.
On the other hand, the Lie bundle will be non-trivial\footnote{If the base space is $\mathbb{R}^1$ local triviality implies also global triviality.} (weak but not strong) if $f(\sigma_0)= 0$ for some
$\sigma_0 \in \mathbb{R}^1$, since in this case $\phi$ is ill-defined for $\sigma=\sigma_0$.
We still, however, can claim that the Lie bundle is trivial everywhere \emph{except} at
$\sigma=\sigma_0$.
Similarly we can see the rescaling (\ref{WeylRescaling}) as a trivialization map for the Lie algebra bundle given by:
\begin{equation}
\label{tildeJJJ2}
[\tilde{J}_2, \tilde{J}_3] = i \tilde{J}_1 \, , [\tilde{J}_3, \tilde{J}_1] = i \tilde{J}_2
\quad \textrm{and} \quad [\tilde{J}_1, \tilde{J}_2] = \sigma^2 \tilde{J}_3 \, .
\end{equation}
At $\sigma=0$ the trivialization map is singular and the Lie algebra structure reduces to (\ref{tildeJJJ}), but at any other point the algebra is isomorphic to $SO(3)$.
\subsection{The \.{I}n\"on\"u-Wigner bundle}
We are now in a position to define the main ingredient of our construction.
{\bf Definition.}
\emph{
The \.{I}n\"on\"u-Wigner (IW) bundle of a Lie algebra $g$ is a (weak) Lie algebra bundle over an affine line $\mathbb{R}^1$,
such that its restriction to $\mathbb{R}^1 \! \setminus \! \{0\}$ is a trivial bundle $g \times \mathbb{R}^1 \! \setminus \! \{0\}$.
}
From the above discussion the Lie bundle defined by (\ref{tildeJJJ2}) and (\ref{f-ijk}) are examples of IW bundles. The latter is non-trivial if and only if $f(0)=0$.
As yet another example consider the following IW Lie algebra bundle\footnote{
The form of the commutators in (\ref{Weyl1}) suggests an analogy with \emph{loop algebras}. The similarity is, however, illusive,
since the parameter $\sigma$ takes values in $\mathbb{R}^1$ and not in $\mathbb{S}^1$, and, even more importantly, for the loop algebra construction, the Lie algebra necessarily remains the same all over the loop, while in our case, a new algebra emerges at
$\sigma=0$.
}:
\begin{equation}
\label{Weyl1}
[J_2^\sigma, J_3^\sigma] = i \sigma^3 J_1^\sigma , \, [J_3^\sigma, J^\sigma_1] = i \sigma^3 J^\sigma_2 ,
\, [J_1^\sigma, J_2^\sigma] = i \sigma J_3^\sigma \, .
\end{equation}
Upon using the trivialization map:
\begin{equation}
\label{Weyl2}
\phi(J_{1,2}, \sigma) =\frac{J_{1,2}^\sigma}{\sigma^2} \quad \textrm{and} \quad \phi(J_{3}, \sigma) =\frac{J_{3}^\sigma}{\sigma^3} \, .
\end{equation}
we find that for $\sigma \neq 0$ the algebra is isomorphic to $SO(3)$. We see that
the trivialization morphism $\phi$ is actually the rescaling of the generators in (\ref{SO3scaling}).
The fact that $\phi$ fails to be an isomorphism at $\sigma=0$ indicates that
we find a new algebra at this point.
\subsection{The extensions}
Our goal in this subsection is to explore the $\sigma \to 0$ limit.
Let $\widetilde{\mathfrak{g}}$ stand for a vector space of
\emph{smooth} sections of an IW bundle of an algebra $g$. Again, as a vector space $g$
is the same all over the base (and so we drop here the $\sigma$-index),
but its commutators are not. The (infinite) vector space of smooth sections $\widetilde{\mathfrak{g}}$
is spanned
by the vectors $\sigma^n a_i$, where $a_i \in g$ and $n \geqslant 0$.
Moreover, this vector space is a
Lie algebra by its own right, since we can multiply the sections pointwise.
Sections vanishing at $\sigma=0$ is an ideal of this algebra. This ideal is $\sigma \widetilde{\mathfrak{g}}$.
The \emph{classical IW algebra} is the quotient:
\begin{equation}
\mathfrak{g}_0 \equiv \widetilde{\mathfrak{g}}/ {\sigma \widetilde{\mathfrak{g}}} \, .
\end{equation}
To arrive at this algebra it is enough to directly take the $\sigma \to 0$ limit.
Notice that $\mathfrak{g}_0$ and $g$ are identical as vector spaces, but not as Lie algebras.
For (\ref{Weyl1})
the algebra $\mathfrak{g}_0$ is an abelian algebra of $J_1, J_2$ and $J_3$.
Similarly, the scaling (\ref{tildeJJJ}) leads to the $ISO(2)$ algebra.
Up to this point, the $\sigma \to 0$ contraction is identical to the original \.{I}n\"on\"u and Wigner prescription.
We, however, don't want to stop here. Notice that $\sigma^n \widetilde{\mathfrak{g}}$ is an ideal of $\widetilde{\mathfrak{g}}$
for any $n \geqslant 1$. The \emph{semi-classical IW algebra} is defined by:
\begin{equation}
\mathfrak{g}_1 \equiv \widetilde{\mathfrak{g}}/ {\sigma^2 \widetilde{\mathfrak{g}}} \, .
\end{equation}
The algebra $\mathfrak{g}_1$ has \emph{twice} more generators than $\mathfrak{g}_0$, but it might happen that some
of these new generators form an ideal and can be quotiented out.
For the IW bundle defined by
(\ref{Weyl1}) let us introduce $J_i^{(0)} \equiv J_i$ and $J_i^{(1)} \equiv \sigma J_i$.
We find that the semi-classical algebra $\mathfrak{g}_1$ has only one non-trivial commutator:
\begin{equation}
\label{WeylJ}
\left[ J_1^{(0)}, J_2^{(0)} \right] = i J_3^{(1)} \, .
\end{equation}
As was announced in Introduction this is the central extension of the Weyl algebra.
To be more precise, in order to identify $\mathfrak{g}_1$ with the Weyl algebra
we have to quotient the algebra by its abelian ideal $\left\{ J_3^{(0)},J_1^{(1)},J_2^{(1)} \right\}$.
We can easily repeat this procedure for any $n$. Clearly, the new algebra will have $n$ times more generators than the original algebra. For the IW bundle in (\ref{Weyl1}) the
$n$ \emph{level} algebra $\mathfrak{g}_n \equiv \widetilde{\mathfrak{g}}/ {\sigma^{n+1} \widetilde{\mathfrak{g}}}$ is
(here $J_i^{(n)}$ stands for $\sigma^n J_i$ and $J_i^{(n)}=0$ for negative $n$):
\begin{eqnarray}
&& \left[ J_2^{(n)}, J_3^{(m)} \right] = i J_1^{(n+m-3)}, \, \left[ J_3^{(n)}, J_1^{(m)} \right] = i J_2^{(n+m-3)} , \nonumber \\
&& \qquad \left[ J_1^{(n)}, J_2^{(m)} \right] = i J_3^{(n+m-1)}
\end{eqnarray}
with all the other commutators vanishing. One may argue that the output algebra is not particularly interesting.
The situation changes drastically, however, for a more sophisticated trivialization (rescaling) $\phi$
than the one in (\ref{Weyl2}).
Before concluding this section, let us discuss how our semi-classical extension $\mathfrak{g}_1$
may lead to a \emph{central} extension of the $\mathfrak{g}_0$ algebra. To this end it is worth recalling the
rigorous definition of a (not necessarily central) extension. An algebra $\mathfrak{a}^\prime$ is called an extension of
$\mathfrak{a}$ by an ideal $\mathfrak{n}$, if $\mathfrak{a}$ is isomorphic to the quotient $\mathfrak{a}^\prime/\mathfrak{n}$.
This definition can be shortly summarized with the following \emph{short exact sequence}:
\begin{equation}
0 \to \mathfrak{n} \to \mathfrak{a}^\prime\to \mathfrak{a} \to 0 \, .
\end{equation}
If $\mathfrak{n}$ is also inside the center of $\mathfrak{a}^\prime$, then the extension is called \emph{central}.
Consider the quotient $\mathfrak{n} \equiv \sigma \widetilde{\mathfrak{g}} / \sigma^2 \widetilde{\mathfrak{g}} = \sigma \mathfrak{g}_1$.
Obviously, $\mathfrak{n}$
is an ideal of $\mathfrak{g}_1$. Moreover, $\mathfrak{n}$ is an abelian ideal and is isomorphic\footnote{A Lie algebra $g_{\rm abelian}$ is isomorphic to the algebra $g$ as a vector space but all its commutators are vanishing.}
to $\left( \mathfrak{g}_0 \right)_{\rm abelian}$. Our first level extension is actually given by the following exact sequence:
\begin{equation}
\label{exact}
0 \rightarrow \mathfrak{n} \rightarrow \mathfrak{g}_1 \xrightarrow{\pi} \mathfrak{g}_0 \rightarrow 0 \, .
\end{equation}
Since $\mathfrak{n}$ is by construction abelian, our $\mathfrak{g}_1$ extension is \emph{always} abelian.
The ideal $\mathfrak{n}$ is, however, not necessarily inside the center of $\mathfrak{g}_1$.
In fact, $\mathfrak{n}$ consists of elements of the form
$\sigma a_i$ and becomes central if and only if $\mathfrak{g}_0$ is abelian. This was exactly the case for the IW bundle defined by
(\ref{SO3scaling}), which led to the Weyl algebra (\ref{WeylJ}).
It seems, therefore, that our construction produces central extensions only upon very restricting conditions.
In particular, it rules out the Bargmann (non-zero mass) extension of the Galilean algebra.
Luckily, we can slightly modify the extension described by (\ref{exact}).
Assume that $\mathfrak{m}_0$ is an ideal of $\mathfrak{g}_0$. It automatically implies that $\pi^{-1} \left( \mathfrak{m}_0 \right)$ is an ideal of $\mathfrak{g}_1$, where $\pi$ is the projection map from (\ref{exact}). Furthermore,
$\mathfrak{m}_1 \equiv \sigma \pi^{-1} \left( \mathfrak{m}_0 \right)$ is an abelian ideal of $\mathfrak{g}_1$ and the quotient
$\mathfrak{n}/\mathfrak{m}_1 = \sigma \mathfrak{g}_1/\mathfrak{m}_1$
is isomorphic to $\left( \mathfrak{g}_0/\mathfrak{m}_0 \right)_\textrm{abelian}$.
We get the following \emph{abelian} extension of $\mathfrak{g}_0$:
\begin{equation}
\label{exact2}
0 \rightarrow \mathfrak{n}/\mathfrak{m}_1
\rightarrow \mathfrak{g}_1/\mathfrak{m}_1 \xrightarrow{\pi} \mathfrak{g}_0 \rightarrow 0 \, .
\end{equation}
The above extension is central if and only if $\mathfrak{g}_0/\mathfrak{m}_0$ is abelian.
Indeed, the latter requirement is equivalent to the statement that
$[ \mathfrak{g}_0, \mathfrak{g}_0] \subset \mathfrak{m}_0$. But that means that
$[ \mathfrak{g}_1, \mathfrak{g}_1] \subset \pi^{-1} \left( \mathfrak{m}_0 \right)$
and so $[ \sigma \mathfrak{g}_1, \mathfrak{g}_1] \subset \mathfrak{m}_1$. The latter immediately implies
that $\mathfrak{n}/\mathfrak{m}_1$ is central.
This way the number of the new generators in the extended algebra will be smaller than total number of generators in $g$.
We will see in the next section that (\ref{exact2}) naturally leads to the central (mass) extension of the Galilean algebra.
\section{\bf Examples}
We have already seen how the central extension of the Weyl algebra emerges from the $SO(3)$ contraction.
In the rest of the paper we will see other applications of the proposed extension method.
\subsection{\bf The non-central extensions of the Poincar\'e algebra from the de Sitter
algebra contraction}
In $d$ space-time dimensions the commutators of the de Sitter and the anti de Sitter algebras ($SO(d,1)$ and $SO(d-1,2)$) are:
\begin{eqnarray}
\label{deSitter}
\left[ M_{\mu \nu}, M_{\lambda \rho} \right] &=& \eta_{\mu \lambda} M_{\nu \rho} + \eta_{\nu \rho} M_{\mu \lambda} -
\left( \lambda \leftrightarrow \rho \right) \, , \nonumber \\
\left[ M_{\lambda \rho}, P_{\mu} \right] &=& \eta_{\mu \lambda} P_{\rho} + \eta_{\mu \rho} P_{\lambda} \, , \nonumber \\
\left[ P_{\mu}, P_{\nu} \right] &=& \varepsilon M_{\mu \nu} \, .
\end{eqnarray}
Here $\eta = \textrm{diag} \left(-1,1,\ldots,1 \right)$ and $\varepsilon=1,-1$ for the de Sitter
and the anti de Sitter algebra respectively.
We are interested in the zero cosmological constant limit, $\Lambda \to 0$, of the (anti) de Sitter algebra.
As such the algebra (\ref{deSitter}) is $\Lambda$-independent, so we have to properly rescale the operators
introducing $\Lambda$ into the commutators. The new $\Lambda$-dependent algebra will be isomorphic to the original one
for any $\Lambda$ except for $\Lambda=0$, where it should reduce to the Poincar\'e algebra.
The right rescaling of the generators is, in fact, well known in the literature.
The metric on the (anti) de Sitter space can be written as:
\begin{equation}
\label{metric}
\textrm{d} s^2_{\textrm{(a)dS}}
= R^2 \cdot \frac{\eta_{\mu \nu} \textrm{d} y^{\mu} \textrm{d} y^{\nu}}{1 + \varepsilon y^2/4 } \, ,
\end{equation}
where $y^2 \equiv \eta_{\mu \nu} y^{\mu} y^{\nu}$ and $R={\Lambda}^{-1/2}$ is the (anti) de Sitter radius.
In terms of these coordinates the isometry generators are:
\begin{align}
\label{MP}
& M_{\mu \nu} =
\eta_{\nu \lambda} y^\lambda \frac{\partial}{\partial y^\mu} - \eta_{\mu \lambda} y^\lambda \frac{\partial}{\partial y^\nu} \, , \nonumber \\
& P_{\mu} = \left( \varepsilon - \frac{y^2}{4} \right) \frac{\partial}{\partial y^\mu} +
\frac{1}{2} \eta_{\mu \lambda} y^\lambda y^\rho \frac{\partial}{\partial y^\rho} \, .
\end{align}
In the limit $\left( R, y^\mu \right) \to \left( \infty, 0 \right)$ with $x^\mu = R \cdot y^\mu$
kept fixed, the metric (\ref{metric}) reduces to the Minkowski metric, while the leading scaling behavior
of the generators (\ref{JBPH}) is:
\begin{equation}
\label{PoincareRescaling}
M_{\mu \nu} \to M_{\mu \nu} \, , \,\, P_{\mu} \to R \cdot P_{\mu} \, .
\end{equation}
It may be also interesting to see the limit in terms of the $\left( \xi^\mu, \xi^d \right)$ coordinates
used for the (anti) de Sitter space embedding in $\mathbb{R}^{d,1}$:
\begin{equation}
\xi^\mu \xi_\mu + \varepsilon \left( \xi^d \right)^2 = \varepsilon R^2 \, .
\end{equation}
These coordinates are related to $y^\mu$'s by:
\begin{equation}
\frac{\xi^\mu}{R} = \frac{y^\mu}{1 + \varepsilon y^2/4} \, , \quad \frac{\xi^d}{R} = \frac{1 - \varepsilon y^2/4}{1 + \varepsilon y^2/4} \, .
\end{equation}
In the Poincar\'e limit this reduces to $\left( \xi^\mu, \xi^d \right)= \left( x^\mu, R \right)$ and so, analogously to the $SO(3)$ example
in the first paragraph of Introduction, the limit corresponds to the zooming around the $\xi^d = R$
point of the (anti) de Sitter space.
With this rescaling we find the following IW bundle:
\begin{eqnarray}
\left[ P^\sigma_{\mu}, P^\sigma_{\nu} \right] &=& \sigma \cdot \varepsilon M^\sigma_{\mu \nu} \, ,
\end{eqnarray}
where the affine parameter $\sigma$ is just the cosmological constant
$\sigma = R^{-2} = \Lambda$ and we left out the commutators from the first two lines in
(\ref{deSitter}) since they remain completely $\sigma$-independent.
As expected, the ``classical'' $\mathfrak{g}_0$ algebra is precisely the Poincar\'e algebra:
\begin{eqnarray}
\label{Poincare}
\left[ M^\textrm{\tiny{(0)}}_{\mu \nu}, M^\textrm{\tiny{(0)}}_{\lambda \rho} \right] &=& \eta_{\mu \lambda} M^\textrm{\tiny{(0)}}_{\nu \rho} \quad + \textrm{permutations} \, , \nonumber \\
\left[ M^\textrm{\tiny{(0)}}_{\lambda \rho}, P^\textrm{\tiny{(0)}}_{\mu} \right] &=& \eta_{\mu \lambda} P^\textrm{\tiny{(0)}}_{\rho} + \eta_{\mu \rho} P^\textrm{\tiny{(0)}}_{\lambda} \, , \nonumber \\
\left[ P^\textrm{\tiny{(0)}}_{\mu}, P^\textrm{\tiny{(0)}}_{\nu} \right] &=& 0 \, .
\end{eqnarray}
The next level ``semi-classical'' algebra $\mathfrak{g}_1$ includes also the operators $M_{\mu \nu}^\textrm{\tiny{(1)}}=\sigma M_{\mu \nu}$ and
$P_{\mu}^\textrm{\tiny{(1)}}=\sigma P_{\mu}$. Notice, however, that the translations $P_{\mu}^\textrm{\tiny{(0)}}$ form an ideal of the Poincar\'e algebra
$\mathfrak{g}_0$. We therefore can use the modified exact sequence (\ref{exact2}) with $\mathfrak{m}_0$ being the
subalgebra of translations $P_{\mu}^\textrm{\tiny{(0)}}$, so that the extension will include
only the $M_{\mu \nu}^\textrm{\tiny{(1)}}$ generators and not $P_{\mu}^\textrm{\tiny{(1)}}$'s. Defining $Z_{\mu \nu} = \varepsilon M_{\mu \nu}^\textrm{\tiny{(1)}}$ we arrive at the Maxwell algebra:
\begin{eqnarray}
\label{Maxwell}
\left[ M^\textrm{\tiny{(0)}}_{\mu \nu}, Z_{\lambda \rho} \right] &=& \eta_{\mu \lambda} Z_{\nu \rho} + \eta_{\nu \rho} Z_{\mu \lambda} -
\eta_{\mu \rho} Z_{\nu \lambda} - \eta_{\nu \lambda} Z_{\mu \rho} \, , \nonumber \\
\left[ P^\textrm{\tiny{(0)}}_{\mu}, P^\textrm{\tiny{(0)}}_{\nu} \right] &=& Z_{\mu \nu} \, ,
\end{eqnarray}
with the first two commutators in (\ref{Poincare}) remaining intact. The extension is non-central since $Z_{\mu \nu}$
does not commute with the Lorentz generators. It becomes a scalar only for $d=2$.
Following the recipe from the previous section we can extend the algebra to higher levels.
The commutators between $Z_{\mu \nu}$'s become non-trivial already at the $\mathfrak{g}_2$ algebra:
\begin{equation}
\label{ZY1}
\left[ Z_{\mu \nu}, Z_{\lambda \rho} \right] = \eta_{\mu \lambda} Y_{\nu \rho} + \eta_{\nu \rho} Y_{\mu \lambda} -
\eta_{\mu \rho} Y_{\nu \lambda} - \eta_{\nu \lambda} Y_{\mu \rho} \, ,
\end{equation}
where $Y_{\mu \nu} \equiv M^\textrm{\tiny{(2)}}_{\mu \nu}=\sigma^2 M_{\mu\nu}$ and:
\begin{eqnarray}
\label{ZY2}
\left[ P^\textrm{\tiny{(0)}}_{\mu}, P^\textrm{\tiny{(1)}}_{\nu} \right] &=& Y_{\mu \nu} \, .
\end{eqnarray}
For similar extensions of the Maxwell algebra recently studied in the literature
see \cite{Soroka:2006aj}, \cite{Bonanos:2008kr}, \cite{Bonanos:2008ez} and \cite{Gomis:2009dm}.
Before ending this section let us remark that the de Sitter algebra (\ref{deSitter}) permits an
additional, infinite cosmological constant limit \cite{Aldrovandi:1998ux}. For fixed $x^\mu = R \cdot y^\mu$ but with
$\left( R, y^\mu \right) \to \left( 0,\infty\right)$ the momenta scale like $P_{\mu} \to P_{\mu}/R$. The contracted algebra
is equivalent to the Poincar\'e but its physical meaning is different, since the $P^\textrm{\tiny{(0)}}_{\mu}$'s now are special conformal transformations
and not translations.
\subsection{\bf The Galilean contraction of the Poincar\'e algebra}
Our next example is the non-relativistic contraction of the Poincar\'e algebra.
In $d$ space-time dimensions the Poincar\'e algebra consists of\footnote{
We will use bold font to indicate $3$-dimensional space vectors.}
$(d-1)(d-2)/2$ space-space rotations $\mathbf{J}$,
$d-1$ boosts $\mathbf{B}$, $d-1$ space translations $\mathbf{P}$ and the time translation $H$.
In all of the remaining examples but the last one we will explicitly consider the $d=4$ case,
with a straightforward generalization for other dimensions.
The commutators are\footnote{Here $\left[ \mathbf{A}, \mathbf{B} \right] = \mathbf{C}$ is a shorthand notation for
$\left[ A_i, B_j \right] = \epsilon_{ijk} C_k$ and $\left[ \mathbf{A}, \mathbf{B} \right] = C$ stands for
$\left[ A_i, B_j \right] = \delta_{ij} C_k$. }:
\begin{equation}
\label{PoincareJBPH1}
\left[ \mathbf{J}, H \right] = 0, \,\,\, \left[ \mathbf{J}, \mathbf{J} \right] = \mathbf{J}, \,\,\,
\left[ \mathbf{J}, \mathbf{B} \right] = \mathbf{B}, \,\,\, \left[ \mathbf{J}, \mathbf{P} \right] = \mathbf{P}
\end{equation}
and:
\begin{equation}
\label{PoincareJBPH2}
\left[ H, \mathbf{B} \right] = \mathbf{P}, \quad \left[ \mathbf{B}, \mathbf{B} \right] = -\mathbf{J},
\quad \left[ \mathbf{P}, \mathbf{B} \right] = H \, ,
\end{equation}
with all other commutators vanishing.
The first set of the commutators
(\ref{PoincareJBPH1}) simply implies that $\mathbf{B}$, $\mathbf{P}$ and $\mathbf{J}$ transform as vectors under space rotations,
while $H$ is a scalar.
The differential representation of the operators in terms of the Minkowskian coordinates is:
\begin{align}
\label{JBPH}
& J_{i} = - \epsilon_{ij}^{\,\,\,\,k} x^j \frac{\partial}{\partial x^k} \, , \quad
B_{i} = \delta_{ij} x^j \frac{\partial}{\partial x^0} + x^0 \frac{\partial}{\partial x^i} \, , \nonumber \\
& \qquad P_{i} = \frac{\partial}{\partial x^i} \, , \qquad H = \frac{\partial}{\partial x^0} \, .
\end{align}
The non-relativistic limit corresponds to $c, x^0 \to \infty$ with $t=x^0/c$ held fixed.
The Poincar\'e algebra generators scale in this limit as:
\begin{equation}
\label{GalileanRescaling}
\mathbf{J} \to \mathbf{J} \, , \,\, \mathbf{B} \to c \mathbf{B} \, , \,\,
\mathbf{P} \to \mathbf{P} \, , \,\, H \to \frac{H}{c} \, .
\end{equation}
This rescaling leads to the following IW bundle:
\begin{equation}
\label{GalileiIW}
\left[ H, \mathbf{B} \right] = \mathbf{P}, \quad \left[ \mathbf{B}, \mathbf{B} \right] = - \sigma \mathbf{J},
\quad \left[ \mathbf{P}, \mathbf{B} \right] = \sigma H \, ,
\end{equation}
where $\sigma = c^{-2}$ and for simplicity we dropped the $\sigma$-indices. We also left out the $\sigma$-independent
part of the algebra (\ref{PoincareJBPH1}).
Putting $\sigma=0$ in (\ref{GalileiIW}) we of course find the Galilean algebra $\mathfrak{g}_0$:
\begin{equation}
\label{Galilei}
\left[ H^\textrm{\tiny{(0)}}, \mathbf{B}^\textrm{\tiny{(0)}} \right] = \mathbf{P}^\textrm{\tiny{(0)}}, \,\, \left[ \mathbf{B}^\textrm{\tiny{(0)}}, \mathbf{B}^\textrm{\tiny{(0)}} \right] = 0,
\,\, \left[ \mathbf{P}^\textrm{\tiny{(0)}}, \mathbf{B}^\textrm{\tiny{(0)}} \right] = 0
\end{equation}
together with the regular transformations of $H^\textrm{\tiny{(0)}}$, $\mathbf{B}^\textrm{\tiny{(0)}}$ and $\mathbf{P}^\textrm{\tiny{(0)}}$ under the $\mathbf{J}^\textrm{\tiny{(0)}}$ rotations.
As we have already emphasized before the $\mathfrak{g}_1$ extension has by construction twice more operators than $\mathfrak{g}_0$.
The new operators are $\mathbf{J}^\textrm{\tiny{(1)}}$, $\mathbf{B}^\textrm{\tiny{(1)}}$, $\mathbf{P}^\textrm{\tiny{(1)}}$ and $H^\textrm{\tiny{(1)}}$.
In order to get the central non-zero mass extension of (\ref{Galilei}) we have to use (\ref{exact2}), where the ideal
$\mathfrak{m}_0$ is the subalgebra of all Galilean rotations, boosts and space translations, namely $\mathbf{J}^\textrm{\tiny{(0)}}$,
$\mathbf{B}^\textrm{\tiny{(0)}}$ and $\mathbf{P}^\textrm{\tiny{(0)}}$. This allows us to mode out $\mathbf{J}^\textrm{\tiny{(1)}}$,
$\mathbf{B}^\textrm{\tiny{(1)}}$ and $\mathbf{P}^\textrm{\tiny{(1)}}$ from the extended algebra. Denoting $M=H^\textrm{\tiny{(1)}}$ we find that $\mathfrak{g}_1/\mathfrak{m}_1$
is exactly the Bargmann algebra:
\begin{equation}
\label{Bargmann}
\left[ H^\textrm{\tiny{(0)}}, \mathbf{B}^\textrm{\tiny{(0)}} \right] = \mathbf{P}^\textrm{\tiny{(0)}}, \,\, \left[ \mathbf{B}^\textrm{\tiny{(0)}}, \mathbf{B}^\textrm{\tiny{(0)}} \right] = 0,
\,\, \left[ \mathbf{P}^\textrm{\tiny{(0)}}, \mathbf{B}^\textrm{\tiny{(0)}} \right] = M \, .
\end{equation}
It is noteworthy that one can formally derive the Bargmann algebra by plugging $H=M/\sigma+H^\textrm{\tiny{(0)}}$ into (\ref{GalileiIW}) and matching the
$\sigma$-expansion from the both sides. This approach is, however, not rigorously defined and unlike our method cannot be used for the next
order extension.
\subsection{\bf The Carrollian contraction of the Poincar\'e algebra}
The Carrollian contraction of the Poincar\'e algebra (\ref{PoincareJBPH1}, \ref{PoincareJBPH2})
is given by:
\begin{equation}
\mathbf{J} \to \mathbf{J} \, , \,\, \mathbf{B} \to \frac{\mathbf{B}}{c} \, , \,\,
\mathbf{P} \to \mathbf{P} \, , \,\, H \to \frac{H}{c} \, .
\end{equation}
It naturally follows from the $c \to 0$ limit of (\ref{JBPH}). The IW bundle is:
\begin{equation}
\label{CarrollIW}
\left[ H, \mathbf{B} \right] = \sigma \mathbf{P}, \quad \left[ \mathbf{B}, \mathbf{B} \right] = - \sigma \mathbf{J},
\quad \left[ \mathbf{P}, \mathbf{B} \right] = H \, ,
\end{equation}
where this time $\sigma=c^2$. For $\sigma=0$ we find the Carrollian algebra:
\begin{equation}
\label{Carroll}
\left[ H^\textrm{\tiny{(0)}}, \mathbf{B}^\textrm{\tiny{(0)}} \right] = 0, \quad \left[ \mathbf{B}^\textrm{\tiny{(0)}}, \mathbf{B}^\textrm{\tiny{(0)}} \right] = 0,
\quad \left[ \mathbf{P}^\textrm{\tiny{(0)}}, \mathbf{B}^\textrm{\tiny{(0)}} \right] = H^\textrm{\tiny{(0)}} \, .
\end{equation}
By analogy with the Galilean case we can mode out some operators from the
next level $\mathfrak{g}_1$ extension. Notice that , $\mathbf{J}^\textrm{\tiny{(0)}}$, $\mathbf{B}^\textrm{\tiny{(0)}}$ and $H^\textrm{\tiny{(0)}}$ form an ideal $\mathfrak{m}_0$
and so we can exclude $\mathbf{J}^\textrm{\tiny{(1)}}$, $\mathbf{B}^\textrm{\tiny{(1)}}$ and $H^\textrm{\tiny{(1)}}$ from the $\mathfrak{g}_1$ extension.
With $\mathbf{Z} = \mathbf{P}^\textrm{\tiny{(1)}}$ the $\mathfrak{g}_1/\mathfrak{m}_1$ algebra takes the following form:
\begin{equation}
\left[ H^\textrm{\tiny{(0)}}, \mathbf{B}^\textrm{\tiny{(0)}} \right] = \mathbf{Z}, \quad \left[ \mathbf{B}^\textrm{\tiny{(0)}}, \mathbf{B}^\textrm{\tiny{(0)}} \right] = 0,
\quad \left[ \mathbf{P}^\textrm{\tiny{(0)}}, \mathbf{B}^\textrm{\tiny{(0)}} \right] = H^\textrm{\tiny{(0)}} \, .
\end{equation}
This extension is central only for $d=2$ when $\mathbf{Z}$ becomes a scalar.
\subsection{\bf The exotic Newton-Hooke algebra}
In our discussion of the Galilean algebra extension we excluded the operator $\mathbf{J}^\textrm{\tiny{(1)}}$ from the ``semi-classical'' algebra.
We did so because this operator is not central in the extended $4d$ Galilean algebra, for it does not commute with the $\mathbf{J}^\textrm{\tiny{(0)}}$ rotations.
The situation changes for $d=3$, since in this case $\mathbf{J}$ is a scalar. The $d=3$ Galilean algebra with
the $J$ central extension is known in the literature as the exotic Galilean algebra.\footnote{One may also consider a similar
exotic central extension of the $3d$ Carrolllian algebra.}
Instead of focusing our attention on this algebra let us make the discussion a bit more general by starting with the Newton-Hooke (NH) contraction.
The NH algebra is the non-relativistic limit of the de Sitter algebra.
To be more precisely, one sends both $c$ and $R$ to infinity with the parameter $\omega=c/R$ held fixed.
For $d=3$ the de Sitter algebra is:
\begin{equation}
\label{d3deSitter1}
\left[ J, H \right] = 0 \, ,
\quad \left[ J, B_i \right] = \epsilon_{ij} B_j \, ,
\quad \left[ J, P_i \right] = \epsilon_{ij} P_j
\end{equation}
and:
\begin{align}
& \left[ H, B_i \right] = P_i \, ,
\quad \left[ B_i, B_j \right] = -\epsilon_{ij} J\, ,
\quad \left[ P_i, B_j \right] = \delta_{ij} H \, ,\nonumber \\
& \qquad \left[ P_i, H \right] = B_i \, ,
\quad \left[ P_i, P_j \right] = \epsilon_{ij} J \, ,
\end{align}
where $(H,P_i)=P_\mu$, $B_{i}=M_{0i}$ and $J=M_{12}$.
Combining the rescalings (\ref{PoincareRescaling}) and (\ref{GalileanRescaling}):
\begin{equation}
J \to J \, , \,\, B_i \to c B_i \, , \,\,
P_i \to \frac{P_i}{R} \, , \,\, H \to \frac{H}{c R}
\end{equation}
we obtain a new IW bundle with the $\sigma=c^{-2}$ affine parameter:
\begin{align}
& \left[ H, B_i \right] = P_i \, ,
\,\, \left[ B_i, B_j \right] = - \sigma \cdot\epsilon_{ij} J\, ,
\,\, \left[ P_i, B_j \right] = \sigma \cdot \delta_{ij} H \, ,\nonumber \\
& \qquad \left[ P_i, H \right] = \omega^2 B_i \, ,
\,\, \left[ P_i, P_j \right] = \sigma \cdot \epsilon_{ij} \omega^2 J \, ,
\end{align}
where we omitted again all the commutators with $J$, because they are $\sigma$-independent.
Obviously the boosts $B^\textrm{\tiny{(0)}}_{i}$ and the momenta $P^\textrm{\tiny{(0)}}_{i}$ will form an ideal of $\mathfrak{g}_0$. Using this ideal for
(\ref{exact2}) we arrive at the \emph{exotic} $3d$ Newton-Hooke algebra:
\begin{align}
\label{eNH}
& \left[ H^\textrm{\tiny{(0)}}, B^\textrm{\tiny{(0)}}_i \right] = P_i \, ,
\,\, \left[ B^\textrm{\tiny{(0)}}_i, B^\textrm{\tiny{(0)}}_j \right] = - \epsilon_{ij} \widetilde{Z} \, , \nonumber \\
& \qquad \qquad \left[ P^\textrm{\tiny{(0)}}_i, B^\textrm{\tiny{(0)}}_j \right] = \delta_{ij} M \, ,\nonumber \\
& \left[ P^\textrm{\tiny{(0)}}_i, H^\textrm{\tiny{(0)}} \right] = \omega^2 B^\textrm{\tiny{(0)}}_i \, ,
\,\, \left[ P^\textrm{\tiny{(0)}}_i, P^\textrm{\tiny{(0)}}_j \right] = \epsilon_{ij} \omega^2 \widetilde{Z} \, ,
\end{align}
where $\widetilde{Z} \equiv J^\textrm{\tiny{(1)}}$ and $M \equiv H^\textrm{\tiny{(1)}}$. For $\omega=0$ (\ref{eNH}) transforms into
the exotic Galilean algebra.
\section{\bf Open questions}
In our paper we have shown that given an \.{I}n\"on\"u-Wigner contraction of one algebra to another, one can easily find an infinite
extension of the latter. The method works for both central and non-central extensions and the extension may be truncated
at any level.
There are plenty of open problems to be explored. Let us list some of them:
\begin{itemize}
\item It is will be interesting to establish a connection between our extension approach and the expansion method presented in
\cite{deAzcarraga:2002xi}. This method is somewhat similar in spirit to ours.
It is based on the Maurer-Cartan one-forms expansion in powers of a real parameter which
is related to the rescaling of the Lie group coordinates.
Our results should also be compared to those of \cite{Bonanos:2008ez}.
\item It is not clear what is the physical meaning of the new (in)finite algebras. In some cases the answer is already known. For example,
the Maxwell algebra (\ref{Maxwell}) corresponds to a charged relativistic particle moving in a constant electromagnetic field.
The meaning of the second level extension (\ref{ZY1},\ref{ZY2}) is yet to be understood.
\newline
\item Finding irreducible representations of the infinitely extended algebras and their truncated versions might be a challenging problem.
\end{itemize}
\begin{acknowledgments}
SK would like to thank Frank Ferrari for collaboration on the early stages of this long term project.
We are grateful to Jarah Evslin, Carlo Maccaferri, Chethan Krishnan, Glenn Barnich, Laurent Houart,
Ricardo Argurio, Marc Henneaux, Jacob Sonnenschein, Anatoly Dymarsky, Ayan Mukhopadhyay and especially Daniel and Melvin
Persson for fruitful discussions.
The work of SK is supported by the European Commission Marie Curie Fellowship under the
contract IEF-2008-237488.
\end{acknowledgments}
\begingroup\raggedright |
\section{Introduction}
\indent In quantum informatic and quantum control settings, trapped ions have demonstrated their versatility for high-fidelity studies of entanglement~\cite{Turchette98, Sackett2000, Blatt8Ion, Reichle2006, Blatt2008, Monroe09}, coherent control~\cite{Bible, BiercukQIC2009, Monroe_Fast2010}, quantum logic~\cite{Monroe95, BlattCZ, Leibfried2003, Monroe_Transverse09}, precision spectroscopy~\cite{Schmidt2005}, and quantum simulation~\cite{QSim98, Leibfried2002, BlattDirac, Monroe_QSim}. In addition to the manipulation of internal atomic states, these experiments rely on the coherent excitation of normal motional modes of these charged particles in approximately harmonic (pseudo-)potential wells~\cite{Bible}. The Coulomb interaction between neighboring ions allows common motional modes to serve as a quantum coherent bus for sharing information between particles, and has functioned as a fundamental mechanism for the experimental realization of two-qubit gates~\cite{Leibfried2003} and multipartite entanglement~\cite{Turchette98}. Recently, exploitation of transverse motional modes in linear ion crystals has been proposed as a means by which large-scale quantum information processing may be realized using trapped ions~\cite{CiracPorras_SpinBoson, MonroeTransverse}.
\\
\indent Implementing a multi-qubit quantum logic operation requires the accumulation of a conditional phase between constituent particles. In trapped-ion experiments, extant techniques rely on the ability to apply a state-dependent force to a Coulomb crystal of harmonically bound trapped ions. Such a force may be realized using an optical dipole excitation created by a pair of off-resonant excitation laser beams. In short, the application of a state-dependent ac Stark shift to trapped ions in a spatially graded light field results in the ions experiencing differential forces dependent upon their internal quantum states (e.g. for a qubit $F_{\uparrow}$ and $F_{\downarrow}$). This provides a means to conditionally excite normal modes of motion. In the interaction frame the excited harmonic oscillator coherently acquires a state-dependent phase. Applying this interaction for an appropriate duration leads to a maximal entanglement of the internal states of multiple ions, and can be used to implement a multipartite quantum logic operation~\cite{Leibfried2003}.
\\
\indent Exploiting ion motional modes in this way requires precise knowledge of motional frequencies, mode spacings, absolute and relative force strengths, and motional coherence. Such characterization is generally accomplished via measurement of time-averaged ion fluorescence under continuous or pulsed excitation of motional modes, and the observation of motional sidebands. This technique is highly efficient for small ion crystals~\cite{King98} in which only a few motional modes exist, and where the optical-dipole interaction is weakly sensitive to the alignment of the excitation-beam interference wavefronts relative to a particular crystallographic direction. However, such detection becomes inefficient and complex in large crystals~\cite{BlattCrystals, mitchell98, Bollinger2000} with densely packed mode structures, and in which the Lamb-Dicke parameter makes observation of sidebands difficult.
\\
\indent We are thus motivated to develop methods by which ion motional excitation may be observed directly, and individual motional excitations may be resolved in a densely packed mode structure. The experiments we present in this manuscript describe the detection of optical dipole excitation of trapped-ion motional modes via laser Doppler velocimetry~\cite{mitchellDoppler, Berkeland1998}. We utilize phase-coherent detection of the transverse center-of-mass (COM) motional mode of a two-dimensional crystal of $\sim$100 $^{9}$Be$^{+}$ ions in a Penning trap~\cite{Biercuk_yN_2010}, excited using state-dependent optical dipole forces. Experimental measurements of the temporal modulation of photon arrival times agree well with theoretical predictions for a driven harmonic oscillator and provide both amplitude and phase information about the excited motion.
\\
\indent The remainder of this paper is organized as follows. In Sec.~\ref{Sec:Expt} we briefly introduce our experimental system, followed by a detailed description of phase-coherent motional excitation and detection via Doppler velocimetry in Sec.~\ref{Sec:Technique}. Sec.~\ref{Sec:OD} describes our excitation-laser system for generation of optical dipole forces and presents phase-coherent Doppler detection measurements. We move on to a study of the benefits of maintaining phase information in the discrimination of nearly degenerate motional modes in Sec.~\ref{Sec:Phase}, and conclude with a discussion and a future outlook.
\section{\label{Sec:Expt}Experimental system}
\indent The technique we describe is general and may be applied in RF Paul or Penning traps; here we focus on experimental implementation in a macroscopic Penning trap. This is a device allowing for charged-particle confinement in three dimensions using static electric and magnetic fields in an ultra-high-vacuum envelope. The trap consists of a stack of axially aligned cylindrical electrodes in the presence of a $\sim$4.5 T, axially oriented external magnetic field. Axial confinement for charged particles is achieved by the application of a static electric potential difference between endcap and center electrodes (here $\sim$1,200 V, see Fig.~\ref{Fig:Raman}c). Radial confinement is realized by ion motion in the axially oriented magnetic field producing a centripetal Lorentz force~\cite{Bollinger1984,brel88}.
\\
\indent Ions are Doppler laser-cooled to an axial temperature $\sim$0.5 mK \cite{jenm04, jenm05} using $\sim$313-nm UV laser light red-detuned from an atomic transition between the $2s$ $^{2}S_{1/2}$ and $2p$ $^{2}P_{3/2}$ manifolds of $^{9}$Be$^{+}$. When cooled, the ions form arrays with well defined crystal structure \cite{itaw98, mitchell98, jenm04, jenm05} and ion spacing $\sim$15 $\mu$m due to Coulomb repulsion. The ion array rotates rigidly at a frequency $\sim$50 kHz due to the geometry of the confining electric and magnetic fields, and the rotation frequency is precisely controlled by an external rotating dipole potential \cite{huap98a, huap98b}. Adjusting the rotation rate of the array allows controllable manipulation of the crystal aspect ratio and dimensionality (planar vs. 3D). For these experiments we focus on crystals with $\sim$100 ions in a two-dimensional planar array having a diameter $\sim$300 $\mu$m.
\\
\indent Quartz windows allow optical access to the trap along the axial direction (parallel to $\vec{B}$) and also transverse to the applied magnetic field (the ``side-view''). An $f/5$ imaging system connected to a CCD camera or phototube (selectable via an electrically controlled flipper mirror) allows for direct side-view imaging of resonantly scattered photons from the ion crystal or detection of the total resonant fluorescence. Pulsed control over all laser beams is provided by switching the rf-drive applied to acousto-optic modulators (AOMs). The time-domain control system is based on a programmable logic device under computer control.
\section{\label{Sec:Technique}Phase-coherent detection of ion motional modes}
\indent In an approximately harmonic confining potential, ions experience shared normal-modes of motion, and $n$ ions exhibit $3n$ normal modes. In our experiments we will aim to excite a normal mode of motion using an optical dipole force, and detect this excitation directly. A variety of three-dimensional motional modes in a one-component $^{9}$Be$^{+}$ plasma have been studied previously~\cite{mitchellDoppler} in this system. In this work we focus exclusively on the center-of-mass (COM) motional mode transverse to the crystal plane, where ions undergo uniform axial motion parallel to the magnetic field. For our operating parameters the COM-mode resonance frequency is $\sim$867 kHz, although over a timescale of hours it drifts downward as heavy-mass ions (e.g. BeH$^{+}$) accumulate in the trap due to background-gas collisions.
\\
\indent In previous work we demonstrated phase-coherent detection of ion motion in which the electrically excited motional response was synchronized to the external driving force. It was shown that this technique was capable of discriminating motional responses to forces at the scale of yoctonewtons. A detailed characterization of the method, its sensitivity, performance limits, and scaling with external parameters is presented in~\cite{Biercuk_yN_2010} and the associated supplementary material. Here we present only a review of the salient elements of the technique.
\\
\indent Motional detection is accomplished using laser Doppler velocimetry. In this technique, ion motion in a direction parallel to the propagation direction of a laser beam modulates the intensity of resonant fluorescence due to Doppler shifts. In these experiments we employ the axially directed laser used for Doppler cooling near 313 nm as a probe of ion motion in the confining potential of the Penning trap. The laser is red-detuned from the Doppler cooling cycling transition by $\sim$12 MHz such that its frequency sits near the steepest section of the atomic profile mentioned above, where Doppler cooling efficiency and detection sensitivity are maximized~\cite{mitchellDoppler} (Fig. 1a). Under oscillatory ion motion, the intensity of ion fluorescence will exhibit periodicity with the same frequency as the ion motion.
\\
\indent A pulsed force is applied for a fixed duration, after which the Doppler detection light is switched on and resonantly scattered photons are detected using a phototube (~\ref{Fig:Schematic}). The driving force, $F_{d}$, is applied at a fixed frequency, and a voltage discriminator is used to produce a chain of ``start'' pulses (here, from a photodiode interference signal) synchronized to the drive that are then sent to a time-to-amplitude converter (TAC). The TAC is gated such that it accepts only the final start pulse prior to the end of the driving force pulse. A ``stop'' pulse is generated by the detection of a resonantly scattered photon on the phototube, and is fed into the TAC. The voltage output of the TAC scales linearly with the difference between start and stop pulse arrival times, and is read using a triggered data acquisition board. The experiment is repeated a fixed number of times, and statistics of photon arrival times synchronized to the drive are collected.
\begin{figure}[tb]
\centering\includegraphics[width=10cm]{ODSchematic}
\caption{\label{Fig:Schematic}Schematic of excitation and detection scheme. (a) Details of pulse-timing used in phase-coherent excitation and detection of motional drive. (b) Schematic of experimental measurement apparatus. Dashed lines represent optical signals, solid lines represent electrical signals. Switches controlled via TTL signals shown schematically in panel (a).}
\end{figure}
\\
\indent In order to understand the expected signal in a Doppler velocimetry experiment we begin by presenting a quantitative analysis of the driven harmonic oscillator. Considering uniform ion motion in a quadratic potential we may write the equation of motion for the axial COM mode in terms of the equation of motion for a single ion as
\begin{equation}
m(\ddot{z}+\omega_{z}^{2}z)=F_{d}\sin(\omega_{d} t + \psi),
\end{equation}
where $m$ is the ion mass, $z$ is the axial coordinate, $\omega_{z}$ is the resonant oscillation frequency, $F_{d}$ is the magnitude of the spatially uniform driving force, and $\omega_{d}$ and $\psi$ are the frequency and the phase of the drive, respectively. Without loss of generality we may set $\psi\equiv0$, and assuming $(\omega_{z}-\omega_{d})/\omega_{z}\ll1$ and $z(0)=\dot{z}(0)=0$, we find
\begin{equation}
z(t)\approx\frac{2F_{d}}{m(\omega^{2}_{z}-\omega^{2}_{d})}\cdot\sin\left[\frac{1}{2}\left(\omega_{d}-\omega_{z}\right)t\right]\cdot\cos\left[\omega_{z}t+\frac{1}{2}\left(\omega_{d}-\omega_{z}\right)t\right].
\end{equation}
After application of a driving force at frequency $\omega_{d}$ of duration $t_{d}$ an ion will undergo a steady-state sinusoidal oscillation with velocity
\begin{equation}
\label{Eq:velocity1}
\dot{z}(t)=v\sin\left[\omega_{z}t+\phi\right],
\end{equation}
\noindent where the amplitude $v$ is defined by
\begin{equation}
\label{Eq:velocity2}
v=\frac{2F_{d}\omega_{d}}{m(\omega_{z}^{2}-\omega_{d}^{2})}\sin\left[\frac{(\omega_{z}-\omega_{d})t_{d}}{2}\right],
\end{equation}
\noindent and the oscillation phase $\phi$ is
\begin{equation}
\label{Eq:velocity3}
\phi=\frac{(\omega_{d}-\omega_{z})t_{d}}{2}.
\end{equation}
\noindent Doppler detection directly probes the ion velocity, as described in equations~\ref{Eq:velocity1}-~\ref{Eq:velocity3}. The measurement protocol we have developed preserves the phase information of the driven harmonic oscillator (Eq.~\ref{Eq:velocity3}) rather than simply averaging measurements over random oscillation phases.
\section{\label{Sec:OD}Direct observation of the optical dipole force}
\indent Optical dipole forces are applied using a two-color laser setup~\cite{Bible} that produces a traveling-wave interference pattern. By appropriately tuning the laser frequency and polarization (relative to the level structure of the trapped ions), the laser can produce ac Stark shifts of varying magnitude and sign. The presence of a spatial gradient in the light field then causes the ions to experience a force as they seek regions of low or high electric field intensity, depending on the sign of the Stark shift.
\\
\indent Our experimental setup involves splitting a 313 nm laser beam into two beamlines, each passing through an independently controlled AOM (Fig.~\ref{Fig:Raman}a). One beam is frequency shifted by 210 MHz and the other by 210 MHz + $\Delta\omega$. The beams enter the trap stacked vertically, and nearly perpendicular to the applied magnetic field at a shallow angle of $\theta=\pm0.75^{\circ}$, intersecting at the location of the planar ion crystal (Fig.~\ref{Fig:Raman}c). Before entering the trap each beam passes through a $f=$ 75 cm cylindrical lens ($f$ indicates focal length) and a $f=$ 75-cm spherical lens to produce an oblate beam with a $\sim$10:1 aspect ratio, and a minimum waist parallel to the trap axis of $\sim$80 $\mu$m at the location of the ion crystal (Fig.~\ref{Fig:Raman}b). Each beam is coarsely aligned using a large cloud of bright ions illuminated by the cooling beam while the excitation-laser beams are tuned to optically pump the ions to a dark state. Adjusting the positions of the independently controlled 75-cm spherical lenses allows for maximization of the beam overlap with the ion crystal. Care is taken to ensure that the beams enter the trap in the same vertical plane.
\begin{figure}[tb]
\centering\includegraphics[width=10cm]{Figure1v4OPX}
\caption{\label{Fig:Raman}Laser excitation setup. (a) Schematic diagram of laser beamline. (b) Schematic of lens system used for creation of oblate beams with independent control of wavefront tilts. Double-headed arrows indicate translation stages. In experimental setup beams are stacked vertically, and converge with crossing angle of 1.5$^{\circ}$. (c) Schematic side-view of laser excitation beamline showing trap electrodes, ion cloud, and orientation of axial magnetic field. (d) Side-view image of fluorescence from the excitation lasers when tuned to be degenerate ($\Delta\omega=0$) and on-resonance with cycling transition used for Doppler detection. The stationary interference pattern arising from lasers used to excite the optical dipole force is clearly visible. Here, the ions are heated above the liquid-solid phase transition and do now show crystalline order.}
\end{figure}
\indent When $\Delta\omega=0$ a standing-wave interference pattern is established where the beams intersect. We verify that the beams are striking the ions and that they are overlapped by direct imaging; the beams are tuned near the Doppler-cooling cycling transition and we image the standing-wave interference pattern as alternating bright and dark bands on an ion cloud above the liquid-solid phase transition~\cite{Bollinger1984, Bollinger1994}. For the small angle at which the beams cross and the $\sim$313 nm wavelength we observe interference fringes with spacing $\sim$10 $\mu$m (Fig.~\ref{Fig:Raman}d). The phase wavefronts of the interference pattern are aligned with the trap axes through beam steering and the addition of a mirror mounted on a linear translation stage which can laterally displace the lower beam relative to the upper beam before each passes through the 75-cm lenses (Fig.~\ref{Fig:Raman}b). A combination of translation of the mirror with translation of the spherical lens allows for adjustment of the wavefront angles relative to the ion crystal.
\\
\indent Optical-dipole excitation of the COM mode is realized when $\Delta\omega\equiv\omega_{d}\approx\omega_{Z}$. In this situation, the overlapped beams produce a ``running wave'' interference pattern, where the AC electric field amplitude at a fixed location oscillates temporally at $\Delta\omega$~\cite{Bible}. Equation 1 is valid if the axial extent of a crystallized plane of ions is small relative to the wavelength of the interference pattern (the Lamb-Dicke criterion). Here, the Lamb-Dicke parameter for a single ion at $T\approx 1$ mK is
\begin{equation}
\eta=\sqrt{\frac{\hbar}{2m_{Be}\omega_{Z}}}2k\sin(\theta)\sqrt{\overline{n}+1}\approx0.07
\end{equation}
where $k$ is the wavevector of the composite two-color laser system, and the average thermal phonon occupation, $\overline{n}=23$.
\indent We set the laser frequency $\sim$3 GHz red of the cooling/detection cycling transition, laser polarization $\sigma^{+}$, and $\Delta\omega\approx\omega_{Z}$. A beat-note is generated for the detection system by picking off $\sim$5$\%$ of the excitation beam power immediately before the beams enter the trap, and interfering the beams on a photodiode. This beat-note is used to synchronize photon arrival times with the drive excitation. In this configuration the beat-note is stable on a timescale of up to a few seconds before undergoing $\sim$$\pi$ phase-slips due to interferometric instability between the two excitation laser beam paths.
\\
\indent We measure the arrival time of the first scattered photon, synchronized to the excitation beat-note, as a function of $\Delta\omega$. In the data presented in Fig.~\ref{Fig:Doppler} the colorscale represents the residual, in the presence of the dipole excitation, of an exponential fit to photon arrival times used to account for photon arrival statistics in the absence of optical-dipole excitation~\cite{Biercuk_yN_2010}. Modulation of the Doppler velocimetry signal due to ion motion is manifested as positive or negative deviations in the detected photon number at a particular time from the exponential background. Periodic modulations are thus represented by alternating dark and light bands as a function of delay time from the beat trigger.
\\
\indent Experimental data match well with theoretical calculations based on the equations presented above (Fig.~\ref{Fig:Doppler}), showing a clear resonance feature near $\omega_{Z}/2\pi=867$ kHz. Doppler velocimetry indicates a linear phase shift in the oscillator as $\omega_{d}$ is tuned through resonance, and the linewidth of the resonance scales as the inverse of $t_{d}$. The excitation of the oscillator is zero when $|\omega_{d}-\omega_{z}|=2\pi/t_{d}\times n$, for $n$ an integer, producing oscillation sidebands. At excitation nulls, the external force and system response desynchronize and resynchronize~\cite{Leibfried2003}, yielding a closed loop in position-momentum phase-space. We see sidelobes of the main resonance; however, these are in general difficult to observe, possibly due to timing jitter induced by interferometric instability in producing a beat-note.
\\
\indent In the experimental data we also observe a damping of the oscillation strength as a function of the delay time due to radiation pressure from the detection laser. This effect is not accounted for in the calculations shown in the second column of Fig.~\ref{Fig:Doppler}. In these calculations the strength of the applied force is used as a model parameter, and is extracted from theoretical fits to the oscillation amplitude as a function of drive frequency, as in previous experiments. Because the strength of the drive force is extracted from the average oscillation amplitude over the whole oscillation period, the failure to account for laser-induced damping provides a slight underestimate of the optical dipole force strength. However, we clearly observe motional coherence beyond 10 $\mu$s.
\\
\indent Calibration of the Doppler velocimetry signal relative to a given applied force is accomplished using quasistatic electric fields as described in detail in~\cite{Biercuk_yN_2010}. The maximum observed force as deduced from theoretical fits to the measured data and comparison against the calibration described above is approximately 100 yN for $\sim$3 GHz detuning from resonance and power of $\sim$5 mW/beam. This measured force strength matches well to theoretical calculations of the achievable force given the excitation beam detuning, configuration, polarization, and power~\cite{Bible}, within a factor of order unity (300 yN for these laser conditions). The strength of the optical dipole force scales inversely with the detuning of the excitation laser frequency from the relevant atomic transition~\cite{Bible}, and we have verified this scaling qualitatively by choosing operating detunings from $\sim1-12$ GHz.
\begin{figure}
\centering\includegraphics[width=8cm]{Figure3v4}
\caption{\label{Fig:Doppler}Direct detection of the optical dipole force in a two-dimensional ion crystal. (a,c,e) Experimental measurements of phase-coherent Doppler velocimetry as a function of excitation-laser difference frequency (vertical) and time delay for arrival of first detected photon. Times shorter than $\sim$4.5 $\mu$s are discarded due to hardware delays. Color scale indicates deviation from a fit to a simple exponential accounting for photon arrival statistics. (b,d,f) Theoretical calculations using equations 4 and 5, and a fit to the magnitude of the oscillation amplitude. Theory replicates salient experimental features including the resonance-linewidth scaling with drive time, and phase evolution with drive frequency. (g) Optical dipole force excitation of a mode nearly degenerate with the COM. In this measurement the phase evolution across resonance changes abruptly indicating the presence of multiple motional resonances. (h) Integrated root-mean-squared velocity of the resonance, using coherent excitation and detection, but averaging over phase information.}
\end{figure}
\indent Using this detection technique we are able to maximize the strength of the measured COM excitation by precisely aligning the excitation laser beams relative to the planar ion crystal through an iterative search in the beam positions. The magnitude of the observed Doppler signal is sensitive not only to misalignment of the center of each beam's waist relative to the ion cloud (which reduces the average AC-Stark shift experienced by the ions), but also to misalignment of the interference wavefronts relative to the orientation of the crystal plane. In this setup, a lateral shift of the lower beam's mirror translator (coupled with a shift of the associated 75 cm spherical lens) relative to the upper beam by $\sim$200 $\mu$m tilts the interference wavefronts by 0.5$^{\circ}$ relative to the ion plane, resulting in a $\sim\pi/2$ phase shift of the force across the diameter of a crystal with $\sim$100 ions. Spatially inhomogeneous force application results in reduced contrast in Doppler detection measurements due to spatial averaging arising from the ion rotation in the trap. Thus, we are able to optimize the measured dipole force strength through micron-scale translations in the relative beam positions, aligning the interference wavefronts in two dimensions with precision much better than 0.5$^{\circ}$.
\\
\section{\label{Sec:Phase}Benefits of maintaining phase information}
\indent Typical measurements of motional modes use incoherent probes of, e.g. microwave energy absorption, to reveal resonant absorption features. These approaches are limited by the fact that it is often difficult to discriminate between a single absorption resonance at frequency $\omega_{0}$ with Fourier-linewidth $t_{d}^{-1}$, and a pair of closely spaced resonances at frequencies $\omega_{0}\pm \Delta/2$, with $\Delta$ the mode separation. Resolving closely spaced resonance features using standard techniques requires increasing the interrogation time to produce a more favorable ratio of $\Delta/t_{d}^{-1}$, which may be undesirable for technical or practical reasons.
\\
\indent By contrast, coherent excitation and detection of closely spaced motional modes provides a means to improve discrimination by extracting information from the phase degree of freedom. We have experimentally performed this kind of discrimination, identifying features in the phase degree of freedom that are obscured in averaged experiments. In Fig.~\ref{Fig:Doppler}g we present measurements of Doppler velocimetry as a function of drive frequency that reveal a resonant feature with internal structure. Instead of a simple monotonic evolution of the oscillation phase as the drive frequency is swept across resonance, we see a more complex pattern in which the phase evolution abruptly changes direction twice, producing a zig-zag pattern.
\\
\indent We associate the observed complex phase evolution in the motional resonance with the crossing of a higher-order asymmetric mode with the symmetric COM mode. Plasma-mode calculations show that for a planar ion-crystal asymmetric, drum-head, or tilt modes can arise in which motion in the axial direction is not uniform across all ions. Asymmetric modes are Doppler shifted by the rotation of the planar ion crystal and, for certain rotation frequencies, produce apparent mode crossings in the lab frame.
\\
\indent This particular mode-crossing evolved in time from an independent spectral feature to an apparent crossing, possibly due to the gradual accumulation of heavy-mass ions that centrifugally separate from $^{9}$Be$^{+}$. A time-varying fraction of heavy-mass ions produces a Doppler-shifted mode whose resonant frequency shifts in the lab frame. Further, this feature appeared in a circumstance where, prior to the accumulation of a significant fraction of heavy-mass ions, the Doppler velocimetry signal was weak, indicating poor alignment of the excitation-laser interference wavefronts with the ion crystal.
\\
\indent In order to further investigate the benefits of this kind of phase-sensitive detection in discriminating nearly degenerate motional modes, we calculate the velocimetry response for an excitation of duration $t_{d}$ at frequency $\omega_{d}$, in the presence of two modes at $\omega_{0}\pm \Delta/2$, with $\Delta$ the mode separation (Fig. 4). The first row of this figure compares the integrated RMS velocimetry signal using phase-coherent excitation and detection against an incoherent excitation in which total energy absorption is measured. For large $\Delta$, the individual resonances are easily resolved by either technique. However, as $\Delta\rightarrow^{+}t_{d}^{-1}$, energy absorption measurements produce a ``double-hump'' feature that eventually becomes a single peak with a non-Lorenzian lineshape for $\Delta<t_{d}^{-1}$. In some regimes both techniques produce a double-hump, but phase-coherent excitation and detection produces a deeper trough due to interference between the closely spaced modes (Fig. 4d).
\\
\indent Additional information is provided by directly examining the phase degree of freedom in a coherent excitation and detection, either as a function of time (second row, Fig.4) or only at $t=0$ (third row). Maintaining this information provides an additional means to distinguish between closely spaced resonance features. For sufficiently small $\Delta$, distinguishability is provided by interference between the sidebands (Fig. 4g) producing a pattern distinct from the single-peak feature (Fig. 4f). Alternatively, by adding a temporal offset between the conclusion of the drive period and the beginning of the velocimetry measurement (bottom row, Fig. 4), one may allow the closely spaced resonances to acquire phase $\sim\pi$, producing destructive interference. The ability to employ these approaches to extracting additional information from phase-coherent doppler velocimetry are predominantly limited by the $Q$ of the mechanical resonance, and clear observation of sidebands. Additionally, as the required time delay increases as $\Delta^{-1}$, the time required to resolve closely spaced resonances must be traded-off against the benefits of using an increased drive time, reducing the Fourier linewidth of the resonances.
\begin{figure}[htbp]
\centering\includegraphics[width=12cm]{Figure4}
\caption{\label{Fig:Phase}Comparison of calculated phase-coherent Doppler velocimetry to energy-absorption measurments in the presence of closely spaced motional modes. (a-e) Integrated RMS velocity and absorbed energy for motional modes with spacings, $\Delta$, designated by the column headers, and parameterized by $t_{d}^{-1}$. (f-j) Two-dimensional doppler velocimetry signal as a function of drive detuning (horizontal) and photon arrival time (vertical). Phase information in both the central resonances and sidebands due to interference between closely spaced motional modes permits distinguishability relative to energy absorption approaches. (k-o) One-dimensional phase-trace demonstrating information contained in the phase-degree of freedom for $t=0$. (p-s) Demonstration of interference between closely spaced motional modes due to the inclusion of a delay ("Offset") between conclusion of the drive and commencement of Doppler detection, presented in units of the inverse peak spacing $\Delta^{-1}$. For a fixed value of $\Delta$, accumulation of a time-dependent phase shift between motional modes improves resolving power.}
\end{figure}
\section{Discussion and conclusion}
\indent Optical dipole forces are widely employed in atomic physics, and have become a fundamental tool in the realization of quantum gates for trapped-ion quantum information. In these experiments, optical dipole forces are tuned to provide a force which is conditioned on the internal state of a trapped atomic ion, and as such produce a means by which information may be coherently transferred from an internal to an external degree of freedom~\cite{Bible}. Precise characterization of motional modes is therefore required in order to realize a robust quantum operation. In the case of ion pairs or small ion crystals only a small number of motional modes are present and the system is not particularly sensitive to the alignment of the excitation lasers relative to the crystallographic axes. In such settings a simple measurement of ion fluorescence as a function of excitation frequency is sufficient to characterize the optical dipole excitation of a given motional mode. However, as ion trap quantum information experiments move towards more complex and larger ion crystals, mode densities increase and it becomes increasingly difficult to distinguish particular resonant frequencies using these techniques. Further, alignment of the excitation lasers with the crystal becomes more important as the size of a one- or two-dimensional crystal increases. With these considerations in mind, it becomes clear that a diagnostic technique providing more information about the optical dipole force becomes important as we look to scale-up ion trap experiments.
\\
\indent In this manuscript we have detailed the phase-coherent excitation and detection of motional modes in a $\sim$100-ion crystal via optical dipole forces. This technique has provided a direct means to observe the optical dipole force using laser Doppler velocimetry, and provided quantitative information about the magnitude of the driving force, the phase and frequency of the excited motional mode, and the effect of radiation damping.
\\
\indent Our results demonstrate that it is possible to exploit the phase information in our detection technique to discriminate nearly degenerate motional modes spaced by less than the Fourier-limited linewidth of a particular resonance. This capability is especially important in large crystals where the mode density is high and it may be difficult to distinguish between narrowly spaced modes using time-averaged measurements in which phase-information is discarded. In addition we have shown that it is possible to observe asymmetric motional excitations and mode-crossings using this technique, and that it may be employed as a sensitive probe of beam alignment relative to an ion crystal.
\\
\indent In these experiments we have focused on excitation of the COM mode of a two-dimensional ion crystal using a single detection beam oriented perpendicular to the ion plane. In this configuration we are limited to the detection of motion along the detection beam. However, it is in principle possible to expand this system by performing sequential Doppler velocimetry measurements with a series of detection beams oriented along orthogonal directions. Similarly, the addition of a spatially resolving top-view imaging system allows time-resolved detection of ion fluorescence from individual ions, and hence the ability to directly detect general nonsymmetric drumhead modes. We believe that this technique will serve as a useful diagnostic tool for laser-beam alignment and motional-mode characterization in complex experiments using large multidimensional ion crystals.
\\
\\
\noindent This manuscript is a contribution of NIST, a US government agency, and is not subject to copyright. This work was supported by the DARPA OLE program and by IARPA. M.J.B. acknowledges partial fellowship support from IARPA and GTRI. The authors wish to thank C. Ospelkaus for technical assistance, James Chou and Brian Sawyer for comments on the manuscript, and T. Rosenband for valuable discussions.
\end{document}
|
\section{Introduction}
Measurements of the expansion history, $H(z)$, from Type Ia supernovae (SNe~Ia)
provide crucial, empirical constraints to help guide the emerging cosmological
model. While high-redshift SNe~Ia reveal that the Universe is now accelerating
\citep{riess98,perlmutter99}, nearby ones provide the most precise measurements
of the present expansion rate, $H_0$.
Recently, high-redshift measurements from the cosmic microwave background
radiation (CMB), baryon acoustic oscillations (BAO), and SNe~Ia have been used
in concert with an assumed cosmological model to {\it predict} the value of
$H_0$ \citep[e.g.,][]{komatsu11}. They are not, however, a substitute for its
{\it measurement} in the local Universe. Such forecasts of $H_0$ from the
high-redshift Universe also make specific assumptions about unsettled questions:
the nature of dark energy, the global geometry of space, and the basic
properties of neutrinos (number and mass). Instead, we can gain insights into
these unknowns from a precise, local measurement of $H_0$. The most precise
measurements of $H_0$ have come from distance ladders which calibrate the
luminosities of nearby SNe~Ia through {\it Hubble Space Telescope (HST)}
observations of Cepheids in their host galaxies (see \citealt {freedman10} for a
review).
In the early Cycles of {\it HST}, the SN~Ia {\it HST} Calibration Program
\citep[hereafter SST]{sandage06} and the {\it HST} Key Project \citep[hereafter
KP]{freedman01} each calibrated $H_0$ via Cepheids and SNe~Ia using the Wide
Field/Planetary Camera 2 (WFPC2) and the Large Magellanic Cloud (LMC) as the
first rung on their distance ladder. Unfortunately, the LMC was not an ideal
anchor for the cosmic ladder because its distance was constrained to only
5--10\% \citep{gibson00a}, its Cepheids (observed from the ground) are of
shorter mean period ($\Delta <P> \approx 25$~d), and lower metallicity ($\Delta
{\rm [O/H]} = 0.4$) than those of the spiral galaxies hosting nearby
SNe~Ia. These differences and uncertainties between ground-based and space-based
photometric zeropoints introduced a 7\% systematic error in the determinations
of $H_0$ obtained by those teams (see \S 4). Additional uncertainty arose from
the unreliability of the measurements from several of the SNe~Ia selected by
SST, which were {\it photographically} observed, highly reddened, atypical, or
discovered after peak brightness. Only three SNe~Ia (SNe 1990N, 1981B, and
1998aq) from the SST sample lacked these shortcomings, defining only a small set
of nearby SNe suitable to calibrate $H_0$. Despite careful work, the teams'
estimates of $H_0$, each with an uncertainty of $\sim 10$\%, differed from each
other's by 20\%, due to the aforementioned systematic errors. Additional
progress required rebuilding the distance ladder to address these systematic
errors.
The installation of the Advanced Camera for Surveys (ACS) extended the
range of {\it HST} for observing Cepheids, reduced their crowding with
finer pixel sampling, and increased their rate of discovery by
doubling the field of view. In Cycle 11, members of our team began
using ACS to measure Cepheids at optical wavelengths in the hosts of
more modern SNe~Ia (SN 1994ae by \citealt{riess05}; SN 1995al and SN
2002fk by \citealt{riess09b}) and in a more ideal anchor galaxy
\citep[NGC\,4258 by][]{macri06}.
In {\it HST} Cycle 15, we began the ``Supernovae and $H_0$ for the
Equation of State'' (SH0ES) project to measure $H_0$ to better than
5\% precision by addressing the largest remaining sources of
systematic error. The SH0ES program constructed a refurbished
distance ladder from high-quality light curves of SNe~Ia, a geometric
distance to NGC\,4258 determined through radio (very long baseline
interferometry; VLBI) observations of megamasers, and Cepheid
variables observed with {\it HST} in NGC\,4258 and in the hosts of
recent SNe~Ia. The reduction in systematic errors came from additional observations of NGC\,4258 and from our use of purely
{\it differential} measurements of the fluxes of Cepheids with similar
metallicities and periods throughout all galaxies in our sample. The
latter rendered our distance scale insensitive to possible changes in
Cepheid luminosities as a function of metallicity or to putative
changes in the slope of the period-luminosity relations from galaxy to
galaxy. We measured $H_0$ to 4.7\% precision \citep[hereafter
R09]{riess09a}, a factor of two better than previous measurements
with {\it HST} and WFPC2. An alternate analysis using the
\citet{benedict07} parallax measurements of Milky Way Cepheids in lieu
of the megamaser distance to NGC\,4258 showed good agreement, with comparable
5.5\% precision.
This result formed a triumvirate of constraints in the
Wilkinson Microwave Anisotropy Probe (WMAP) 7-year analysis (i.e., BAO
+ $H_0$ + WMAP-7yr) which were selected as the combination most insensitive to systematic
errors with which to constrain the cosmological parameters
\citep{komatsu11}. Together with the WMAP constraint on $\Omega_M
h^2$, this measurement of $H_0$ provides a constraint on the nature of dark energy,
$w = -1.12 \pm 0.12$ \citep[R09 and][]{komatsu11},
which is comparable to but independent of the use of high-redshift SNe~Ia.
It also improves constraints on the
properties of the elusive neutrinos, such as the sum of their masses and the
number of species \citep{komatsu11}.
In {\it HST} Cycle 17 we used the newly installed Wide Field Camera 3
(WFC3) to increase the sample sizes of both the Cepheids and the SN~Ia
calibrators along the ladder used by SH0ES to determine $H_0$. The
near-infrared (IR) channel of WFC3 provides an order of magnitude
improvement in efficiency for follow-up observations of Cepheids over
the Near-Infrared Camera and Multi-Object Spectrograph (NICMOS), while
the finer pixel scale of the visible channel (relative to ACS) is
valuable for reducing the effects of crowding when searching for
Cepheids. We present these new observations in \S2, the
redetermination of $H_0$ in \S3, and an analysis of the error budget
including systematics in \S4. In \S5, we address the use of this new
measurement along with external datasets to constrain properties of
dark energy and neutrinos.
\section{WFC3 Observations of Cepheids in the SH0ES Program}
The SH0ES program was developed to improve upon the calibration of the
luminosity of SNe~Ia in order to better measure the Hubble constant. To ensure
a reliable calibration sample we selected SNe~Ia having the following qualities:
(1) modern photometric data (i.e., photoelectric or CCD), (2) observed before
maximum brightness, (3) low reddening (implying $A_V < 0.5$ mag), (4)
spectroscopically normal, and (5) optical {\it HST}-based observations
of Cepheids in its host galaxy. In addition to providing robust distance
measures, these qualities are crucial for producing a calibration sample which
is a good facsimile of the SN~Ia sample they are used to calibrate --- i.e.,
those defining the modern SN~Ia magnitude-redshift relation at $0.01 < z < 0.1$
\citep[e.g.,][]{hicken09a}.
In {\it HST} Cycles 16 and 17, we used WFC3, ACS, and WFPC2 to discover Cepheids
in two new SN~Ia hosts: NGC\,5584 \citep[host of SN 2007af;][]{macri11a} and
NGC\,4038/9 \citep[``the Antennae,'' host of SN 2007sr;][]{macri11b} whose light
curves were presented in \citet{hicken09a}\footnote{ We augmented the Hicken et
al light curve of SN 2007sr with 3 pre-discovery V-band observations from the
All-Sky Automated Survey (ASAS) extending to pre-maximum; MJD,mag,error
triplets are (4441.85,13.44,0.12), (4448.86,12.65,0.05),(4452.85,12.73,0.08)}.
We also employed the optical channel of WFC3 to {\it reobserve all previous
SN~Ia hosts in the calibration sample and NGC\,4258}. This provided for the
first time a calibration of all Cepheid optical and infrared photometry using
the same zeropoints. In the case of some hosts, the additional epoch (obtained
well after the prior ones) allowed us to discover previously unidentified,
longer period ($P > 60$~d) variables. We also used these observations to search
for additional Cepheids in the hosts which previously had the smallest numbers
of Cepheids: NGC\,3021, NGC\,3982, NGC\,4536, and NGC\,4639
\citep{macri11c}. The new observations, together with those from
\citet{riess09b}, \citet{saha96,saha97,saha01}, \citet{gibson00b},
\citet{stetson01}, and \citet{macri06}, provide the position, period, and phase
of 730 Cepheids in 8 hosts with reliable SN~Ia data as well as NGC\,4258. The
Cepheids in each host were typically imaged on 14 epochs in $F555W$ and 2--5
epochs with $F814W$ (except for NGC\,4258, which has 12 epochs of $F814W$
data). An illustration of the entire dataset used to observe the Cepheids is
shown in Figure 1. Having previously determined the positions, periods, and
optical magnitudes of these Cepheids, it is highly advantageous to observe their
near-IR magnitudes with a single photometric system in order to (1) reduce the
differential extinction by a factor of five over visual bands, (2) reduce the
possible dependence of Cepheid luminosities on chemical composition
\citep{marconi05}, and (3) negate zeropoint errors. This was previously done
with NICMOS on {\it HST} in Cycle 15 by R09 for a subset of these Cepheids.
The near-IR channel on WFC3 provides a tremendous gain over
NICMOS for the study of extragalactic Cepheids. Photometry of
comparable signal-to-noise ratio can be obtained in a quarter of the
exposure time. More significant for Cepheid follow-up observations is
the factor of 40 increase in area of WFC3-IR over NICMOS/Camera~2
(NIC2), the channel which offered the best compromise between area and
uniform pixel sampling. The one advantage of NIC2 over WFC3-IR is
better sampling of the point-spread function (PSF); the $0\farcs13$
pixels of WFC3-IR undersample the {\it HST} PSF by a factor of 1.6 at
1.6~$\mu$m. However, the finer sampling of NIC2 is largely offset by
the numerous photometric anomalies unique to that camera, whose
subsequent correction leads to correlated noise among neighboring
pixels which reduces the independence of the NICMOS pixel sampling.
In contrast, the detector of WFC3-IR is much better behaved and
pixel sampling noise can be mitigated with dithering.
\subsection{WFC3 Data Reduction}
Each host galaxy was observed for 2--7~ks with individual exposures
400--700~s in length, using integer and half-pixel dithering between
exposures to improve sampling of the PSF (see Table 1). The WFC3
images of the two new hosts, NGC\,5584 and NGC\,4038, are shown in
Figures 2 and 3. Figure 4 shows an example of a host previously
observed with NIC2 by R09 and with WFC3-IR in this study.
We developed an automated pipeline to calibrate the raw WFC3 $F160W$
frames. The first step was to pass the data through the
STScI-supported {\it calwf3} pipeline in the {\it STSDAS} suite of
routines in {\it PyRAF} to remove the bias and dark current, reject
cosmic rays through up-the-ramp sampling, and flat-field the data. A
small correction to the standard flat-field frame was used to correct
the WFC3 ``blobs,'' which are 10\% depressions in flux covering $\sim
1$\% of the area due to spotting on the WFC3 Channel Select Mechanism
(CSM). Next, we used {\it multidrizzle} to combine the exposures from
each visit into a master image, resampling onto a finer pixel scale
while correcting for the known geometric distortions in the camera. We
utilized a final pixel scale of $0\farcs08$ per pixel and an
input-to-output fraction of 0.6.
We identified the positions of Cepheids in the WFC3-IR
images by deriving geometric transformations from the $F814W$ images
to those in $F160W$, successively matching fainter sources to improve
the registration. This procedure empirically determined the
difference in plate scale between ACS-WFC, WFPC2, WFC3-UVIS, and
WFC3-IR. We typically identified more than 100 sources in common,
resulting in an uncertainty in the mean position of each Cepheid below
0.03 pixels ($<2.4$ milliarcsec).
We carried out the photometry of Cepheids using the algorithms
developed by R09; they employ PSF fitting to model the crowded regions
around Cepheids, fixing their positions to those derived from optical
data and using artificial-star tests to determine photometric errors
and crowding biases. As an example, we show in Figure 5 the {\it HST}
optical image, near-IR image, model, and residuals for 8 typical
Cepheids spanning a wide range of periods in the SN host NGC\,5584.
We used the same approach to determine zeropoints as in R09 from the Persson et al. (1998) standard star
P330E.
\begin{table}[h]
\tablenum{2}
\begin{small}
\begin{center}
\vspace{0.4cm}
\begin{tabular}{lll}
\multicolumn{3}{c}{Table 1: Hosts Observed with WFC3-IR $F160W$ by GO-11570} \\
\hline
\hline
Host & SN~Ia & Exp. time (s) \\
\hline
NGC\,4536 & SN 1981B & 2564 \\
NGC\,4639 & SN 1990N & 5377 \\
NGC\,3982 & SN 1998aq & 4016 \\
NGC\,3370 & SN 1994ae & 4374 \\
NGC\,3021 & SN 1995al & 4424 \\
NGC\,1309 & SN 2002fk & 6988 \\
NGC\,4038/9 & SN 2007sr & 6794$^a$ \\
NGC\,5584 & SN 2007af & 4926 \\
NGC\,4258 & ------------ & 2011$^b$ \\
\hline
\hline
\multicolumn{3}{l}{$^a$Data in GO-11577} \\
\multicolumn{3}{l}{$^b$Depth per pointing; galaxy covered in 16 pointing mosaic} \\
\end{tabular}
\end{center}
\end{small}
\end{table}
Due to the low amplitudes of their near-IR light curves ($< 0.3$ mag),
Cepheid magnitudes determined at random phases provide nearly the same precision as mean fluxes for
determining the intercept
of their $P$--$L$\ relations \citep{madore91}.\footnote{In R09 we
corrected the measured NICMOS $F160W$ magnitude to the mean-phase
magnitude using the Cepheid phase, period, and amplitude from the
optical data, the dates of the NICMOS observations, and the Fourier
components of \citet{soszynski05} which quantify the relations
between Cepheid light curves in the optical and near-IR. However,
these phase corrections of $\sigma \sim 0.1$ mag were found to be
insignificant in the subsequent analysis, since the dispersion of
the observed $P$--$L$\ relations is $\sigma \sim 0.3$ mag. Here we have not
attempted such corrections because the Cepheid phases at the time of
the WFC3 $F160W$ observations were too poorly constrained to allow
for a significant correction.}
Since we had previously observed with NIC2 many of the Cepheids now
observed with WFC3-IR, we can directly compare their $F160W$
photometry on these two systems. Figure 6 shows the magnitude
differences for the Cepheids utilized in the $P$--$L$\ relations in both R09
and in \S 2.2. The mean difference is $0.036 \pm 0.027$~mag (in the
sense that photometry with WFC3 is brighter), with no apparent
dependence on Cepheid brightness. While the difference in photometry
between instruments may include differences in system zeropoints, the
subsequent determination of $H_0$ via Cepheids observed with a single
instrument in the SN~Ia hosts and in NGC\,4258 will be independent of
instrument zeropoints. Thus, for the determination of $H_0$ it is
more relevant to calculate the change in magnitudes between Cepheids
in NGC\,4258 and the SN hosts between WFC3 and NIC2; the measurement
of this change is $0.019 \pm 0.054$~mag.
Table 2 contains relevant parameters for each Cepheid observed with WFC3
$F160W$. The first 8 columns give the Cepheid's host, position, identification
number (from \citealt{macri06}, \citealt{riess09b}, \citealt{macri11a, macri11b,
macri11c}), period, mean $V-I$ color, WFC3 $F160W$ magnitude, and the
magnitude uncertainty. Column 9 contains the displacement of the flux centroid
in the near-IR data relative to the optical Cepheid position, expressed in units
of pixels (1 pixel $ = 0\farcs08$), a quantity used to refine the determination
of the crowding bias. Column 10 gives the photometric crowding bias determined
using the artificial-star tests for each Cepheid's environment (see \S 2.3 of
R09) and the displacement tabulated in the previous column; this correction has
already been applied to the magnitudes listed in Column 7. Column 11 contains
the root-mean square of the residual image, weighted by the inverse distance
from the Cepheid position, useful for determining the quality of the
crowded-scene fit. Column 12 contains the metallicity parameter, 12 + log\,[O/H]
\citep{zaritsky94} derived from the deprojected galactocentric
radii of each Cepheid and the abundance gradient of its host.\footnote{These
gradients were published in R09 for NGC\,4258 and the previously observed six
SN hosts; the values for the new hosts, following the same convention as Table
12 of that paper, are 12 + log\,[O/H] $= 8.981 - 0.064(x-30\arcsec)/10\arcsec$
for NGC\,5584 and 12 + log\,[O/H] $= 9.129 - 0.043(x-30\arcsec)/10\arcsec$ for
NGC\,4038/9.} Column 13 contains a rejection flag used for the $P$--$L$\ relations.
\subsection{Near-Infrared Cepheid Relations}
The nine individual $P$--$L$\ relations measured with WFC3 $F160W$ and fit with a
common slope are shown in Figure 7. Intercepts relative to NGC\,4258 are given
in Table 3 and compared in Figure 8 to the SN distances. While 636 Cepheids previously identified at optical
wavelengths were measurable\footnote{Cepheids were considered to be measured if
our software reported a possible magnitude for the source with an uncertainty
less than 0.7 mag, a model residual with rms better than $3 \sigma $ from the
distribution of all model residuals, and a crowding correction less than 1.5
magnitudes. While these thresholds are somewhat arbitrary, they are
sufficient to remove catastrophic failures in convergence on the measurement
of the photometry for a source.} in the WFC3-IR $F160W$ images, $\sim 20$\%
appeared as outliers in the IR $P$--$L$\ relations. This is not surprising, as we
expect outliers to occur from (1) a complete blend with a bright, red source
such as a red giant, (2) a poor model reconstruction of a crowded group when
the Cepheid is a small component of the group's flux, (3) objects misidentified
as classical Cepheids in the optical (e.g., blended Type II Cepheids), and (4)
Cepheids with the wrong period (aliasing or incomplete sampling of a single
cycle). As expected, the outlier fraction is greater in WFC3 images than in
NIC2 ones because the former contain a larger fraction of Cepheids from crowded
regions (such as the nucleus) which yield more outliers and were intentionally
avoided in the small, selective NIC2 pointings (see Figure 4).
As in R09, we eliminated outliers $> 0.75$ mag or $> 2.5\sigma$ (following
Chauvenet's criterion) from an initial fit of the $P$--$L$\ relations, refitted the
relations and repeated these tests for outliers until convergence. This
resulted in a reduction of the sample to 484 objects; the next section
considers the effect of this rejection on the determination of $H_0$ and an
alternative method for contending with outliers.
\section{Measuring the Hubble Constant}
The determination of the Hubble constant follows from the relations
given in \S 3 of R09. To summarize, we perform a single, simultaneous
fit to all Cepheid and SN~Ia data to minimize the $\chi^2$ statistic
and measure the parameters of the distance ladder. We express the
{\it j}th Cepheid magnitude in the {\it i}th host as
\bq m_{W,i,j}=(\mu_{0,i}-\mu_{0,4258})+zp_{W,4258}+b_W \ {\rm log} \
P_{i,j}+Z_W \ \Delta {\rm log\,[O/H]}_{i,j}, \eq
\noindent where the ``Wesenheit reddening-free'' mean magnitude
\citep{madore82} is given as
\bq m_{W,i,j} =m_{H,i,j} - R(m_{V,i,j} - m_{I,i,j}), \eq
\noindent and $R \equiv A_H / (A_V - A_I)$. The Cepheid parameters with $i,j$
subscripts are given in Table 2. For a \citet{cardelli89} reddening law, a
Galactic-like value of $R_V = 3.1$, and the $H$ band corresponding to the WFC3
$F160W$ band, we have $R = 0.410$. In the next section we consider the
sensitivity of $H_0$ to the value of $R_V$.
We determine the values of the nuisance parameters $b_W$ and $Z_W$ --- which
define the relation between Cepheid period, metallicity, and luminosity --- by
minimizing the $\chi^2$ for the global fit to all Cepheid data. The
reddening-free distances, $\mu_{0,i}$, for the hosts relative to NGC\,4258 are
given by the fit parameters $\mu_{0,i}-\mu_{0,4258}$, while $zp_{4258}$ is the
intercept of the $P$--$L$\ relation simultaneously fit to the Cepheids of NGC\,4258.
The SN~Ia magnitudes in the SH0ES hosts are simultaneously expressed as
\bq m_{v,i}^0=( \mu_{0,i}-\mu_{0,4258})+m^0_{v,4258}. \eq
\noindent
\noindent where the value $m_{v,i}^0$ is the maximum-light apparent $V$-band
brightness of a SN~Ia in the {\it i}th host at the time of $B$-band peak
corrected to the fiducial color and luminosity. This quantity is determined
for each SN~Ia from its multi-band light curves and a light-curve fitting
algorithm, either from the MLCS2k2 \citep{jha07} or the SALT-II \citep{guy05}
prescription (see \S 4.2 for further discussion).
A minor change from R09 is the inclusion of a recently identified, modest
relationship between host-galaxy mass and the calibrated SN~Ia
magnitude. Several studies \citep{hicken09b,kelly10,lampeitl10,sullivan10} have
shown the existence of a correlation between the corrected SN magnitude and the
mass of its host, with a value of 0.03 mag per dex in $M_{\rm stellar}$, in the
sense that more massive (and metal rich) hosts produce more luminous SNe. This
correlation has been independently detected using both low- and high-redshift
samples of SNe~Ia, as well as with multiple fitting algorithms. The effect on
$H_0$ is quite small, a decrease of 0.75\%, due to the modest difference in
mean masses for the nearby hosts \citep[mean log $M_{\rm stellar} =
10.0$]{neill09} and for those that define the magnitude-redshift relation
\citep[mean log $M_{\rm stellar} = 10.5$]{sullivan10}. We include these
corrections based on host-galaxy mass in our present determination of
$m_{v,i}^0$, given in Table 3, normalizing to a fiducial host mass of log
$M_{\rm stellar} = 10.5$ as appropriate for the objects used to measure the
Hubble flow.
\begin{table}[h]
\tablenum{3}
\begin{small}
\begin{center}
\vspace{0.4cm}
\begin{tabular}{lllllll}
\multicolumn{6}{c}{Table 3: Distance Parameters} \\
\hline
\hline
Host & SN~Ia & Filters & $m_{v,i}^0 +5a_v$ & $\sigma^a$ & $\mu_{0,i}-\mu_{0,4258}$ & $\mu_0$ best \\
\hline
n4536 & SN 1981B & $UBVR$ & 15.147 & 0.145 & 1.567 (0.0404) & 30.91 (0.07) \\
n4639 & SN 1990N & $UBVRI$ & 16.040 & 0.111 & 2.383 (0.0630) & 31.67 (0.08) \\
n3370 & SN 1994ae & $UBVRI$ & 16.545 & 0.101 & 2.835 (0.0284) & 32.13 (0.07) \\
n3982 & SN 1998aq & $UBVRI$ & 15.953 & 0.091 & 2.475 (0.0460) & 31.70 (0.08) \\
n3021 & SN 1995al & $UBVRI$ & 16.699 & 0.113 & 3.138 (0.0870) & 32.27 (0.08) \\
n1309 & SN 2002fk & $BVRI$ & 16.768 & 0.103 & 3.276 (0.0491) & 32.59 (0.09) \\
n5584 & SN 2007af & $BVRI$ & 16.274 & 0.122 & 2.461 (0.0401) & 31.72 (0.07) \\
n4038 & SN 2007sr & $BVRI$ & 15.901 & 0.137 & 2.396 (0.0567) & 31.66 (0.08) \\
Weighted Mean & ----- & ----- & ----- & 0.0417 & ------ (0.0133) & ----- \\
\hline
\hline
\multicolumn{6}{l}{$^a$For MLCS2k2, 0.08 mag added in quadrature to fitting error.} \\
\end{tabular}
\end{center}
\end{small}
\end{table}
The simultaneous fit to all Cepheid and SN~Ia data via Equations (1) and (3)
results in the determination of $m^0_{v,4258}$, which is the expected
reddening-free, fiducial, peak magnitude of a SN~Ia appearing in NGC\,4258.
Lastly, the Hubble constant is determined from
\bq {\rm log} \ H_0={(m_{v,4258}^0-\mu_{0,4258})+5a_v+25 \over 5}. \eq
\noindent where $\mu_{4258,0}$ is the independent, geometric distance estimate
to NGC\,4258 obtained through VLBI observations of water megamasers orbiting
its central supermassive black hole
\citep{herrnstein99,humphreys05,argon07,humphreys08,greenhill09}. The term
$a_v$ is the intercept of the SN~Ia magnitude-redshift relation, approximately
${\rm log} cz - 0.2m_v^0$ but given for an arbitrary expansion history as
\bq a_v = {\rm log} \ cz
\left\{ 1 + {1\over2}\left[1-q_0\right] {z}
-{1\over6}\left[1-q_0-3q_0^2+j_0 \right] z^2
+ O(z^3) \right\} - 0.2m_v^0, \eq
\noindent measured from the set of SN~Ia ($z,m_v^0$) independent of any
absolute (i.e., luminosity or distance) scale. As in R09, we determine $a_v$
from a Hubble diagram for 240 SNe~Ia from \citet{hicken09a} using MLCS2k2
\citep{jha07} or the SALT-II \citep{guy05} prescription to determine $m_v^0$.
Limiting the sample to $0.023 < z < 0.1$ (to avoid the possibility of a local,
coherent flow; $z$ is the redshift in the rest frame of the CMB) leaves 140
SNe~Ia. (In the next section we consider a lower cut of $z>0.01$.) Together
with the present acceleration $q_0=-0.55$ and prior deceleration $j_0=1$
\citep{riess07}, we find $a_v=0.698 \pm 0.00225$. Note that
\citet{ganeshalingam10} recently published light curves of large sample of
SNe~Ia from the Lick Observatory Supernova Search, but there is a large overlap
with those given by \citet{hicken09a}. There are only 13 SNe~Ia at $z > 0.023$
not already included in our sample, and their inclusion would have a negligible
impact on the uncertainty in $a_v$, itself one of the smallest contributors to
the error in $H_0$.
The full statistical error in $H_0$ is the quadrature sum of the uncertainty in
the three {\it independent} terms in Equation (4): $\mu_{4258,0}$,
$m^0_{v,4258}$, and $5a_V$, where $\mu_{4258,0}$ is the geometric distance
estimate to NGC\,4258 by \citet{herrnstein99}, claimed by \citet{greenhill09}
to currently have a 3\% uncertainty.
\citet{hui06} point out that the peculiar velocities of SN~Ia hosts and their
correlations can produce an additional systematic error in the determination of
the SN~Ia $m$--$z$ relation used for cosmography. However, by making use of a
map of the matter density field, it is possible to correct individual SN~Ia
redshifts for these peculiar flows \citep{riess97}. \citet{neill07} made use of
the IRAS PSCz density field \citep{branchini99} to determine the effect of the
density field on the low-redshift SN~Ia $m$--$z$ relation and its impact on the
equation-of-state parameter of dark energy, $w = P/(\rho c^2)$ (where $P$ is
pressure and $\rho c^2$ is energy density). Using their results for a
light-to-matter bias parameter $\beta = 0.5$ and the dipole from \citet{pike05}
results in an increase of the mean velocity of the low-redshift sample and in
the Hubble constant by 0.4\% over the case of uncorrelated velocities at rest
with respect to the CMB. We use a new estimate of this mean peculiar velocity
for the \citet{hicken09a} SN sample which is a slightly larger value of
0.5\%. We account for this and assume an uncertainty of 0.1\% resulting from a
$\pm 0.2$ error in the value of $\beta$.
The result is $H_0=$\honosys, a \uncsnosys\ measurement (top line, Table 4).
It is instructive to deconstruct the individual sources of uncertainty to
improve our insight into the measurement. In principle, the covariance between
the data and parameters does not allow for an exact and independent allocation
of propagated error for each term toward the determination of $H_0$. However,
in our case, the diagonal elements of the covariance matrices provide a very
good approximation to the individual components of error. These are given in
Table 5 and shown in Figure 9 for past and present determinations of $H_0$.
A number of improvements since R09 are evident by comparing Columns 2 and 3 in
Table 5 and as shown in Figure 9. The uncertainty in $H_0$ from all of the terms independent of the
megamaser distance to NGC\,4258 is 2.3\%, 50\% smaller than these same terms in
R09, a result of the increased sample of Cepheids and SN calibrators. This term
includes uncertainties due to the form of the $P$--$L$\ relation, Cepheid metallicity
dependences, photometry bias, and zeropoints --- all of which were important
systematic uncertainties in past determinations of the Hubble constant
\citep[see Column 1, which contains the values from][]{freedman01}. In this
analysis, as in R09, these uncertainties have been reduced by the collection of
samples of Cepheids whose measures (i.e., metallicity, periods, and photometric
systems) are a good match between NGC\,4258 and the SN hosts. Here the
contribution from an unknown dependence of Cepheid luminosity on metallicity
has been furthered reduced by 40\% owing to a better match between the
metallicity of the Cepheid samples in NGC\,4258 and the expanded SN host
sample. In R09, the mean metallicity of the NGC\,4258 Cepheid sample on the
ZHK abundance scale was 12 + log\,[O/H] =8.91, nearly the same as the present
mean of 8.90. However, the mean metallicity of the Cepheid sample in the SN
hosts has risen from 8.81 to 8.85. Some of this change can be attributed to
the inclusion of Cepheids closer to the nuclei of the hosts and some to the
inclusion of two new hosts, NGC\,5584 and NGC\,4038/9, with higher-than-average
metallicities. The reduction in the mean abundance difference between
NGC\,4258 and the SN~Ia hosts from 0.077 to 0.045~dex results in a decrease of
the error propagated into $H_0$ from 1.1\% to 0.6\%. A similar reduction is
seen with the use of Milky Way Cepheids whose mean metallicity of 8.9 is closer
to the mean of the new Cepheid sample in the SN hosts. We consider an
alternative calibration of abundances from \citet{bresolin11} in \S 4.1.
\subsection {Buttressing the First Rung}
In our present determination of $H_0$, the 3\% uncertainty in the distance to
NGC\,4258 claimed by \citet{greenhill09} is now greater than all other sources
combined (in quadrature). The next largest term, the uncertainty in mean
magnitude of the eight nearby SNe~Ia, is 1.9\%. To significantly improve upon
our determination of $H_0$, we would need an {\it independent} calibration of
the first rung of the distance ladder as good as or better than the
megamaser-based measurement to NGC\,4258 in terms of precision and reliability.
Independent calibration of the first rung is also valuable as an alternative to
NGC\,4258, should future analyses reveal previously unidentified systematic
errors affecting its distance measurement.
A powerful alternative has recently become available through high
signal-to-noise ratio measurements of the trigonometric parallaxes of Milky Way
Cepheids using the Fine Guidance Sensor (FGS) on {\it HST}. \citet{benedict07}
reported parallax measurements for 10 Cepheids, with mean individual precision
of 8\% and an error in the mean of the sample of 2.5\%. These were used in R09
as a test of the distance scale provided by NGC 4258, but the improvement in
precision beyond the first rung in the previous section suggests greater value
in their use to enhance the calibration of the first rung.
\citet{vanleeuwen07} reanalyzed {\it Hipparcos} observations and determined
independent parallax measurements for the same 10 Cepheids (albeit with half
the precision of {\it HST}/FGS) and for 3 additional Cepheids (excluding Polaris which is an overtone pulsator
and whose estimated fundamental period is an outlier among the Cepheids pulsing in the fundamental mode). The resulting
sample can be considered an independent anchor with a mean, nominal uncertainty
of just 1.7\%. We use the combined parallaxes tabulated by
\citet{vanleeuwen07} and their $H$-band photometry as an alternative to the
Cepheid sample of NGC\,4258 by replacing Equation (1) for the Cepheids in the
hosts of SNe~Ia with
\bq m_{W,i,j}=\mu_{0,i}+M_{W,1}+b_W \ {\rm log} \ P_{i,j}+Z_W \ \Delta
{\rm log\,[O/H]}_{i,j}, \eq
\noindent where $M_{W,1}$ is the absolute Wesenheit magnitude for a Cepheid
with $P = 1$~d, and simultaneously fitting the Milky Way Cepheids with the
relation
\bq M_{W,i,j}=M_{W,1}+b_W \ {\rm log} \ P_{i,j}+Z_W \ \Delta
{\rm log\,[O/H]}_{i,j} . \eq
\noindent Equation (3) for the SNe~Ia is replaced with
\bq m_{v,i}^0=\mu_{0,i}-M_V^0. \eq
\noindent The determination of $M^0_V$ for SNe~Ia together with the previous
term $a_v$ then determines the Hubble constant,
\bq {\rm log} \ H_0={M_V^0+5a_v+25 \over 5}. \eq
Since the near-IR magnitudes of these Milky Way Cepheids have not been directly
measured with WFC3, the use of these variables requires an additional allowance
for possible differences in their photometry. These may arise from differences
in instrumental zeropoints, crowding, filter transmission functions, and
detector well depth at which the sources are measured together with an
uncertainty in detector linearity. Analysis of the absolute photometry from
WFC3-IR \citep{kalirai09} and the ground system (e.g., 2MASS;
\citealt{skrutskie06}) claim absolute precision of 2\%--3\%. We therefore
assume a {\it systematic} uncertainty in the relative magnitudes between {\it
HST} WFC3 $F160W$ Cepheid photometry and the ground-based measurements of
Milky Way Cepheids on the $H$-band system of \citet{persson98} of 4\%. This
reduces the effective precision of the parallax distance scale from 1.7\% to
2.6\%. The ground-based photometry of these Milky Way Cepheids is tabulated by
\citet{groenewegen99} and R09. This systematic error is included in the global
fit as an additional calibration equation with uncertainty given in the error
correlation matrix.
When using the Milky Way Cepheids, we now include an external constraint on the
slope of the near-IR $P$--$L$\ relation. No such constraint was necessary or even of
significant value in the previous section because the Cepheid periods in
NGC\,4258 (mean log\,$P = 1.51$) are so similar to those in the SN~Ia host
(mean log\,$P = 1.63$). In contrast, the mean period of the Milky Way sample
(mean log\,$P = 1.0$) is substantially lower, giving an unconstrained slope of
the $P$--$L$\ relation a greater and unrealistically large lever arm. Following
analyses of optical and near-IR Cepheid data in the Milky Way \citep{fouque07}
and the LMC \citep{persson04,udalski99}, we adopt a conservative constraint on
the slope of the Wesenheit relation of $-3.3 \pm 0.1$~mag per dex in log\,$P$.
Using the Milky Way Cepheids instead of NGC\,4258 as the first rung of the
distance ladder gives $75.7 \pm 2.6$ km s$^{-1}$ Mpc$^{-1}$, in good agreement with (and even greater
precision than) the NGC\,4258-based value. However, an overall improvement in
precision is realized by the {\it simultaneous} use of both the Milky Way
parallaxes and the megamaser-based distance to NGC\,4258, yielding
$74.5 \pm 2.3$ km s$^{-1}$ Mpc$^{-1}$, a remarkably small uncertainty of \uncbothnosys.
Another opportunity to improve upon the first rung on the distance ladder comes
from the sample of $H$-band observations of Cepheids in the LMC by
\citet{persson04}. Recent studies of detached eclipsing binaries (DEBs) by
different groups provide claims of a reliable and precise distance to the LMC.
\citet{guinan98}, \citet{fitzpatrick02}, and \citet{ribas02} studied three
B-type systems (HV2274, HV982, EROS1044) which lie close to the bar of the LMC
and therefore provide a good match to the Cepheid sample of \citet{persson04}.
The error-weighted mean of these is $49.2 \pm 1.6$ kpc\footnote{A fourth system
\citep[HV5936,]{fitzpatrick03} is located several degrees away from the bar
and yields a distance that is closer by 3$\sigma$. Additional lines of
evidence presented in that paper suggest this system lies above the disk of
the LMC, i.e., closer to the Galaxy.}. \citet{pietrzynski09} analyzed
OGLE-051019.64-685812.3, an eclipsing binary system comprised of two giant
G-type stars also located near the barycenter of the LMC, and found a distance
of $50.2 \pm 1.3$ kpc. The average result, 49.8~kpc, provides a good estimate
of the distance to the LMC\footnote{However, we note the analysis by
\citet{schaefer08}, who suggests a level of agreement in recent distance
estimates to the LMC which is too good to be consistent with statistics.}.
Here we retain the larger of the two previous uncertainties to estimate the
distance modulus as $18.486 \pm 0.065$~mag, or an effective error of $\pm
0.076$~mag when including the aforementioned 0.04~mag uncertainty between the
ground-based and {\it HST}-based near-IR photometric systems. Using this
distance to the LMC and the Cepheid sample of \citet{persson04} yields
$71.3 \pm 3.8$ km s$^{-1}$ Mpc$^{-1}$, as seen in Table 4.
Combining all three first rungs (Milky Way, Large Magellanic Cloud, and
NGC\,4258) provides the most precise measurement of $H_0$: \hoall, a slightly
smaller uncertainty of \uncallnosys. As expected, the use of all three anchors
for the distance ladder instead of just one has the largest impact on the
overall uncertainty, reducing the total contribution of the first rung to the
error from 3.3\% to 1.5\%. However, a substantial penalty is paid for the
mixing of ground-based and space-based photometric systems and the resultant
uncertainties in Wesenheit or dereddened magnitudes, adding a 1.4\% error to
$H_0$ where for NGC\,4258 alone none pertained. Modest increases in error also
result from the larger difference in mean Cepheid metallicity (LMC) and period
(LMC and Milky Way).
Past determinations of the absolute distance scale have had a checkered
history, with revisions common. Thus, it may be prudent to rely on no more
than any two of the three possible anchors of the distance scale in the
determination of $H_0$. The omission of NGC\,4258, Milky Way parallaxes, or the
LMC yields a precision in $H_0$ of 3.3\%, 3.2\%, and 3.0\%, respectively. We
thus adopt as our best determination \hoalle, the measurement from all three
sources of the distance scale, but with the larger error associated from only
two independent origins of the distance scale.
Should future work revise the distance to any one of the absolute distance
scale determinations, we provide the following recalibration: $H_0$ decreases
by 0.25, 0.30, and 0.14 km s$^{-1}$ Mpc$^{-1}$ for each increase of 1\% in the
distance to either NGC\,4258, the Milky Way parallax scale, or the distance to
the LMC.
In the last column of Table 3 we also give the best estimate of the distance to
each host from the global fit to all first rungs, Cepheid and SN data. These
are useful to compare to alternative methods of measuring distances to these
hosts or to place a sample of relative measures of SNe Ia distances onto an
absolute scale. For example, there has been recent dissagreement on the
distance modulus of the Antennae (NGC 4038/9); \citet{saviane08} claim a value
of $\mu_0$=30.62 $\pm 0.17$~mag based on the apparent tip of the red giant
branch (TRGB), while \citet{schweizer08} obtain $\mu_0=31.74 \pm 0.27$~mag from
SN 2007sr and $\mu_0=31.51 \pm 0.16$~mag from a different determination of the
TRGB, in agreement with previous estimates by \citet{whitmore99} and
\citet{tonry00} based on flow-field models. Our result of $\mu_0=31.66 \pm
0.08$~mag (with the uncertainty based on the global fit) strongly favors the
``long'' distance to the Antennae.
Although we have been careful to propagate our statistical errors, as well as
past sources of systematic error such as metallicity dependence, system
zeropoint, and instrumental uncertainties, we now consider a broader range of
systematic uncertainties relating to alternative approaches to the analysis of
the data.
\section{Analysis Systematics}
In the preceding section we presented our preferred approach to analyzing the
Cepheid and SN~Ia data, incorporating uncertainties within the framework used
to model the data. Here we follow the same approach used by R09 to quantify
the systematic uncertainty in the determination of $H_0$, by measuring the
impact of a number of variants in the modeling of the Cepheid and SN~Ia data.
In Table 4 we show 15 variants of the previously described analysis for every
combination of choices of distance anchors (NGC\,4258, Milky Way, or LMC), any
two of the preceding or all three; these amount to a total of 105
combinations. Our primary analysis for any anchor choice is given in the first
row (shown in bold) for which that choice initially appears. Column (1) gives
the value of $\chi^2_{\nu}$, Column (2) the number of Cepheids in the fit,
Column (3) the value and total uncertainty in $H_0$, and Column (4) whether the
near-IR data for Cepheids with periods shorter than the completeness limit from
the optical selection were included. Column (5) gives the SN Ia
magnitude-redshift intercept parameter, Column (6) gives the determination of
$M_V^0$ which is specific to the light-curve fitter employed, Column (7) the
calibration system for the metal abundances Column (8) the value and uncertainty
in the metallicity dependence, and Column (9) the value and uncertainty of the
slope of the Cepheid $P$--$L$\ or $P$--$W$ relation. Column (10) gives the minimum
SN~Ia redshift used to define the $m$--$z$ relation, Column (11) encodes aspects
of the SN fitting routine and assumptions therein addressed below, and Column
(12) is the choice of anchors to set the distance scale. Column (13) gives the
type of $P$--$L$\ relation employed, either Wesenheit ($H,V,I$) or $H$-band
only. Column (14) is the reddening law value used for the Cepheids. Column (15)
lists the filters allowed for fitting the SN~Ia light curves and column (16)
gives the value of $R_V$ used to fit the SN light curves.
\subsection{Cepheid Systematics}
In the preceding analysis of the Cepheid data, differences in the determination
of $H_0$ may result from the following variants in the primary analysis: (1)
retention of Cepheids with periods below the optical incompleteness limit; (2)
not allowing for a metallicity dependence; (3) changing the Cepheid reddening
law from $R_V = 3.1$ to $R_V = 2.5$; (4) using only near-IR magnitudes without
reddening corrections; (5) no rejection of outliers in the $P$--$L$\ relations; and
(6) a change in the calibration of chemical abundances. Each of these changes
was implemented as a variant of the primary analysis with results given in
Table 4. The rationale for the primary analysis over each variant was discussed
in detail in \S 4 of R09, with the exception of (6) which is discussed below.
Taken individually, these variants result in $H_0$ rising or declining by
$\lesssim 1.0$ km s$^{-1}$ Mpc$^{-1}$, which is less than half of the
statistical uncertainty. A variant resulting in a larger change occurs when we
do not reject Cepheids which are outliers on the $P$--$L$\ relation, raising $H_0$ by
1.3 km s$^{-1}$ Mpc$^{-1}$. However, the value of $\chi^2_{\nu}$ also triples,
with a total increase in $\chi^2$ of 6 per rejected outlier. As we expect
outliers {\it a priori} to arise from blending or misidentification of Cepheids
(type or period), resulting in residuals in excess of the typical uncertainty,
we believe it is most sensible to reject them to minimize their impact on the
global solution. The use of higher or lower thresholds for outlier rejection
has even less impact than including all outliers. Lowering the outlier
threshold to $2.25\sigma$ (and its accompanying residual magnitude)
reduces $H_0$ by 0.1 km s$^{-1}$ Mpc$^{-1}$. Raising the threshold to $3.0$
or $4.0\sigma$ reduces $H_0$ by 1.0 or 0.8 km s$^{-1}$ Mpc$^{-1}$,
respectively. Neglecting a reddening correction for the Cepheids also raises
$H_0$ by 1.3 km s$^{-1}$ Mpc$^{-1}$ but we believe this correction is
warranted.
As an alternative to rejecting outliers we also considered the approach of
simultaneously modeling the distribution of Cepheids and the outliers.
Following \citet{kunz07} we allowed for a nuisance population
of sources along the $P$--$L$\ relation characterized by a broader distribution
($\sigma=1$ mag) and an intercept independent from that of classical Cepheids.
The {\it a posteri} likelihood function for the intercepts of the Cepheid hosts
was then compared to that derived from outlier rejection. The mean zeropoint of
the SN hosts is greater by 0.013 $\pm$ 0.012 mag. The mean uncertainty of the
intercepts are a factor of 1.38 greater than those from outlier rejection but
still small compared to the distance precision of each SN. The only difference
of note (i.e., $> 0.03$ mag) was for the intercept of NGC 4536 which was greater
by $0.08\pm 0.05$ mag in the outlier modeling over the use of rejection.
The chemical abundance values for the Cepheids used in R09 and here were
estimated from nebular lines in \ion{H}{2} regions of the Cepheid hosts using
the R$_{23}$ parameter and the transformation to an oxygen abundance following
\citet[][hereafter ZKH]{zaritsky94}. There are several alternative
calibrations of the transformation from R$_{23}$ to $\log\,[O/H]$
\citep[e.g.][]{mcgaugh91,pilyugin05}, but these primarily affect the absolute
normalization of the metallicity scale and do not alter the relative
host-to-host differences in abundance or the determination of $H_0$. Recently,
\citet{bresolin11} has redetermined the abundance gradient of NGC 4258 by
adopting the \citet{pilyugin05} calibration of R$_{23}$. This calibration
yields abundances that are consistent with those determined directly by
\citet{bresolin11} in 4 \ion{H}{2} regions in the outer disk of NGC\,4258,
measuring the electron temperature ($T_e$) via the auroral line [\ion{O}{3}]
$\lambda$ 4363. Given this agreement, \citet{bresolin11} suggests the adoption
of a so-called ``T$_e$ scale'' for the determination of absolute chemical
abundances of extragalactic Cepheids.
The T$_e$ recalibration of nebular oxygen abundances not only reduces the
values of $\log\,[O/H]$ by $\sim$ 0.4 dex at the metal-rich end but also
compresses the abundance scale by a factor of 0.69. Based on this scale and
consistent atomic data, \citet{bresolin11} finds a nebular oxygen abundance for
the LMC of $12+\log\,[O/H]=8.36$, moderately lower than the ``canonical'' value
of 8.5 in the ZKH scale. On the T$_e$ scale, the mean apparent metallicity of
the SN Ia and maser hosts would be $12+log\,[O/H]=8.42$; this is closer to the
LMC Cepheids than to the Milky Way Cepheids and a departure from the ZKH scale.
While the abundances of Milky Way Cepheids are not measured the same way (i.e.,
they are based on stellar absorption lines rather than on nearby ionized gas),
they have been directly measured to be $\sim$ 0.3 dex higher than those of LMC
Cepheids \citep{andrievsky02,romaniello08}. The resulting estimate of 8.66 for
the MW Cepheids on the T$_e$ scale would agree well with recent estimates of
the solar oxygen abundance of 8.69 \citep{asplund09} together with a small
gradient in metallicity away from the solar neighborhood. This LMC to MW
Cepheid abundance difference of 0.3~dex also agrees well with the T$_e$ scale
compression of the 0.4~dex difference on the ZKH scale for which the value for
MW Cepheids was taken (here and in R09) to be 8.9.
We determined the effect on $H_0$ of a change from the ZKH scale to the T$_e$
scale by transforming the values of 12+log$\,[O/H]$ using equation (3) of
\citet{bresolin11} and assigning values of 8.36 and 8.66 to LMC and MW
Cepheids, respectively. As seen in Table 4, the value of $H_0$ increases by
0.4 km s$^{-1}$ Mpc$^{-1}$ when using all 3 calibrators and increases by less
than 1.0 km s$^{-1}$ Mpc$^{-1}$ for any combination of 2 calibrators. The
biggest change, an increase of 2.0 km s$^{-1}$ Mpc$^{-1}$, occurs when only the
MW is used to calibrate the first rung, a direct consequence of the increase of
the metallicity difference between the SN Ia host and MW Cepheids on the T$_e$
scale. In the presence of uncertainties concerning the appropriate values of
Cepheid abundances, the determination of $H_0$ based on {\it infrared}
observations of Cepheids should be significantly less sensitive to metallicity
differences than optical Cepheid data \citep{marconi05}. Indeed, the
metallicity correction empirically determined here, $-0.10\pm 0.09$ mag/dex
(using all 3 calibrations) is less than half the value of $\sim -0.25$ mag/dex
measured at optical wavelengths \citep{kennicutt98,sakai04} and its absolute
value is not significant. A better determination of the difference in
metallicity between MW and extragalactic Cepheids may not occur until the
launch of JWST.
\subsection{ SN Systematics}
Here we consider the following variants in the analysis of the SN~Ia data: (1)
minimum range of SN~Ia $m$--$z$ relation lowered from $z = 0.023$ to $z = 0.01$,
(2) discarding $U$-band SN~Ia light-curve data (fit 61), (3) SN~Ia reddening
parameter $R_V=1.5,2.0,3.1$ (fits 29, 28, and 20), (4) use of a SN~Ia
luminosity-color correction with no prior (i.e., as in the $\beta$ parameter of
SALT II instead of an extinction parameter, $R_V$, in MLCS2k2) (fit 26), (5) a
host-galaxy extinction likelihood prior from galaxy simulations (fit 27), and
(6) use of the SALT-II light-curve fitter (fit 42). The motivation for these
variants is described in greater detail in R09.
As seen in Table 4, none of these variants taken individually alters the value
of $H_0$ by more than $\sim 1.5$ km s$^{-1}$ Mpc$^{-1}$ from the preferred
solution, less than half the statistical uncertainty. One of the more
noteworthy variants is the use of the SALT-II light-curve fitter \citep{guy05}
{\it in lieu} of MLCS2k2, since the result of this change can be substantial
for high-redshift data \citep{kessler09}. Observations of high-redshift SNe~Ia
typically have lower signal-to-noise ratios, and thus place greater reliance on
fitters and on the assumptions they include (e.g., the relation between SN~Ia
color and distance). In contrast, the determination of $H_0$ is quite {\it
insensitive} to the fitter; the use of SALT-II results in an increase in
$H_0$ of 1 km s$^{-1}$ Mpc$^{-1}$.
The dispersion of the 15 different determinations of $H_0$ is 0.7 or 0.8 km
s$^{-1}$ Mpc$^{-1}$ for any selected pair of sources of the absolute distance
scale. Adding this measure of analysis systematics to the previous yields
\hofin, a \uncfin\ uncertainty, our best determination.
\begin{table}[h]
\tablenum{5}
\begin{small}
\begin{center}
\vspace{0.4cm}
\begin{tabular}{llllll}
\multicolumn{6}{c}{Table 5: $H_0$ Error Budget for Cepheid and SN~Ia Distance Ladders$^*$} \\
\hline
\hline
Term & Description & Previous & R09 & Here & Here \\
& & LMC & N4258 & N4258 & All 3 \\
\hline
$\sigma_{\rm anchor}$ & Anchor distance & 5\% & 3\% & 3\% & 1.3\% \\
$\sigma_{{\rm anchor}-PL}$ & Mean of $P$--$L$\ in anchor & 2.5\% & 1.5\% & 1.4\% & 0.7\%$^a$ \\
$\sigma_{{\rm host}-PL}/\sqrt{n}$ & Mean of $P$--$L$\ values in SN hosts & 1.5\% & 1.5\% & 0.6 \% & 0.6\% \\
$\sigma_{\rm SN}/\sqrt{n}$ & Mean of SN~Ia calibrators & 2.5\% & 2.5\% & 1.9\% & 1.9\% \\
$\sigma_{m-z}$ & SN~Ia $m$--$z$ relation & 1\% & 0.5\% & 0.5\% & 0.5\% \\
$R \sigma_{\lambda,1,2}$ & Cepheid reddening, zeropoints, anchor-to-hosts & 4.5\% & 0.3\% & 0.0\% & 1.4\% \\
$\sigma_{Z}$ & Cepheid metallicity, anchor-to-hosts & 3\% & 1.1\% & 0.6 \% & 1.0\% \\
$\sigma_{\rm PL}$ & $P$--$L$\ slope, $\Delta$ log $P$, anchor-to-hosts & 4\% & 0.5\% & 0.4\% & 0.6\% \\
$\sigma_{\rm WFPC2}$ & WFPC2 CTE, long-short & 3\% & 0\% & 0\% & 0\% \\
\hline
\multicolumn{2}{l}{subtotal, $\sigma_{H_0}$} & 10\% & 4.7 \% & 4.0\% & 2.9\% \\
\hline
\multicolumn{2}{l}{Analysis Systematics} & NA & 1.3\% & 1.0\% & 1.0\% \\
\hline
\multicolumn{2}{l}{Total, $\sigma_{H_0}$} & 10\% & 4.8 \% & 4.1\% & 3.1\% \\
\hline
\hline
\multicolumn{6}{l}{$^*$Derived from diagonal elements of the
covariance matrix propagated via the error matrices associated} \\
\multicolumn{6}{l}{\ \ with Equations 1, 3, 7, and 8.} \\
\multicolumn{6}{l}{$^a$For Milky Way parallax, this term is already
included with the term above.} \\
\end{tabular}
\end{center}
\end{small}
\end{table}
\section{Dark Energy and Neutrinos}
An independent and precise measurement of $H_0$ is an important complement to
the determination of cosmological model parameters. Alternatively, it serves as
a powerful test of model-constrained measurements at higher redshifts. It is
beyond the scope of this paper to provide a complete analysis of the impact of
the measurement of $H_0$ on the cosmological model from all extant data. We
encourage others to do so. However, one such example using the present
measurement of $H_0$ can be illustrative.
Making use of the simplest present hypothesis for the cosmological model
(namely $\Lambda$-cold-dark-matter without curvature, exotic neutrino physics,
or specific early-Universe physics), and using the single most powerful
cosmological data set (the 7-year WMAP results from \citealt{komatsu11}),
results in a predicted value of $H_0 = 71.0 \pm 2.5$ km s$^{-1}$ Mpc$^{-1}$.
This value agrees well with our determination of \hofin at better than the
combined 1$\sigma$ confidence level.
Alternatively, we can use the WMAP data together with the measured value of
$H_0$ to constrain added complexity to the model. In Figure 10 we show the use
of this data combination for constraining a redshift-independent dark energy
equation-of-state parameter ($w$), the number of relativistic species (e.g.,
neutrino number), and the sum of neutrino masses. The result for dark energy
is $w = -1.08 \pm 0.10$, about 20\% more precise than the same result derived
from the determination of $H_0$ in R09. If we had perfect knowledge of the
CMB, our overall 30\% increase in the precision of $H_0$ would yield the
same-sized improvement in the determination of $w$. However, the fractional
uncertainty in $\Omega_M h^2$ from the WMAP 7-year analysis is comparable to
our measurement of $H_0$; thus, greater precision in $w$ may still be wrung
from future higher-precision measurements of the CMB by WMAP or {\it Planck}.
The enhanced precision in measuring $H_0$ also provides a strong rebuff to
recent attempts to explain accelerated expansion without dark energy but
rather by our presence in the center of a massive void of gigaparsec scale.
Already such models are hard to fathom as they require an exotic location for
the observer, at the center of the void to within a part in a million
\citep{blomqvist10} to avoid an excess dipole in the CMB. It is also not yet
apparent if such a model is consistent with other observables of the CMB or the
late-time integrated Sachs-Wolfe effect. However, using measurements of
$H(z>1)$ to constrain void models of the Lemaitre, Tolmon, and Bondi variety
already predicts slower-than-observed local expansion with values of $H_0$=60
\citep{nadathur10} or 62 \citep{wiltshire07} km s$^{-1}$ Mpc$^{-1}$, more than
5 $\sigma$ below our measurement.
Comparable improvements to cosmological constraints on relativistic species are
also realized from R09, as shown in Figure 10. Most interesting may be the
effective number of relativistic species, $N_{\rm eff} = 4.2\pm0.75$, which is
nominally higher than the value of 3.046 expected from the three known neutrino
species plus tau-neutrino heating from $e^+ e^-$ collisions \citep{mangano05}.
While this nominal excess of relativistic species has been noted previously
\citep[e.g.,]{reid10,komatsu11,dunkley10}, and even interpreted as a possible
indication of the presence of a sterile neutrino \citep{hamann10}, we caution
that the cosmological model provides other avenues for reducing the
significance of this result including additional degrees of freedom for
curvature, dark energy, primordial helium abundance, and neutrino masses. The
30\% improvement in the present constraint on $H_0$ combined with improved high
resolution CMB data (e.g., \cite{dunkley10}) and ultimately with {\it Planck}
satellite CMB data should reduce the present uncertainty in $N_{\rm eff} $ by a
factor of $\sim$ 3 which may provide a more definitive conclusion on the
presence of excess radiation in the early Universe.
\section{Discussion}
Examination of the complete error budget for $H_0$ in the last two columns of
Table 5 indicates additional approaches for improved precision in future
measurements of $H_0$. Expanding the sample of well-measured parallaxes to
Milky Way Cepheids (especially those at log\,$P > 1$) with the GAIA satellite
could drive the precision of the first rung of the distance ladder well under
1\%. However, as we have found with the ``baker's dozen'' of present Milky Way
parallaxes, much of this precision would be lost without better
cross-calibration between the space and ground photometric systems used to
measure Cepheids, near and far.
The largest remaining term comes from the quite limited sample of ideal SN~Ia
calibrators, just 8 objects. The occurrence of an ideal SN~Ia in the small
volume within which {\it HST} can measure Cepheids ($R \approx 30$~Mpc) is
rare, on average only once every 2--3 yr. Given the recent proliferation of SN
surveys and instances of multiple, independent discoveries, we are confident
that all such SNe~Ia within this volume are being found. Collecting more will
require extending the range of Cepheid measurements --- without introducing new
systematics --- and patience. The forthcoming {\it James Webb Space Telescope
(JWST)} offers a promising route to extend Cepheid observations out to 50~Mpc
and to redder wavelengths, where uncertainties due to possible variations in
the extinction law and the dependence of Cepheid luminosities on metallicity
are further reduced. This extension would increase the SN sample suitable for
calibration by a factor of $\sim 5$, reaching $\sim 40$ ideal SNe~Ia observed
over the past 20 yr. Based on a 5\% distance precision per ideal SN, such a
sample would enable a determination of $H_0$ to better than 1\%. However,
discovering these Cepheids may require imaging at optical wavelengths where the
amplitude of the variations is significant, a requirement which will challenge
the short-wavelength capabilities of {\it JWST}.
\section{Summary and Conclusions}
We have improved upon the precision of the measurement of $H_0$ from
\citet{riess09a} by (1) more than doubling the sample of Cepheids observed in
the near-IR in SN~Ia host galaxies, (2) expanding the SN~Ia sample from 6 to 8
with the addition of SN 2007af and SN 2007sr, (3) increasing the sample of
Cepheids observed in NGC\,4258 by 20\%, (4) reducing the difference in
metallicity for the observed sample of Cepheids between the calibrator and the
SN hosts, and (5) calibrating all optical Cepheid colors with WFC3 to remove
cross-instrument zeropoint errors. Further improvements to the precision and
reliability of the measurement of $H_0$ come from the use of additional sources
of calibration for the first rung, foremost of these are the trigonometric
parallaxex of 13 Cepheids in the Milky Way.
Our primary analysis gives $H_0 =$ \hofin\ including systematic errors
determined from varying assumptions and priors used in the analysis. The
combination of this result alone with the WMAP 7-year constraints yields $w =
-1.08 \pm 0.10$ and improves constraints on a possible but still uncertain
excess in relativistic species above the number of known neutrino
flavors. The measured $H_0$ is also highly inconsistent with the simplest inhomogeneous matter models
invoked to explain the apparent acceleration of the Universe without dark energy.
Given that statistical errors still dominate over systematic errors,
future work is likely to further improve the precision of the determination of
$H_0$.
\bigskip
\medskip
We are grateful to William Januszewski for his help in executing this program
on {\it HST}. We are indebted to Mike Hudson for assisting with the
peculiar-velocity calculations from the PSCz survey, to David Larson for
contributions to the WMAP MCMC analysis, to Daniel Scolnic for donating some
useful routines, and to Mark Huber for an analysis of pre-discovery
observations of SN 2007sr. We thank Chris Kochanek and Kris Stanek
for their support of GO-11570. Financial support for this work was provided by
NASA through programs GO-11570 and GO-10802 from the Space Telescope Science
Institute, which is operated by AURA, Inc., under NASA contract NAS
5-26555. A.V.F.'s supernova group at U.~C. Berkeley is also supported by NSF
grant AST--0607485 and by the TABASGO Foundation. L.M.M. acknowledges support
from a Texas A\&M University faculty startup fund. The metallicity
measurements for NGC 5584 and NGC 4038/9 presented herein were obtained with
the W.~M. Keck Observatory, which is operated as a scientific partnership among
the California Institute of Technology, the University of California, and NASA;
the observatory was made possible by the generous financial support of the
W.~M. Keck Foundation.
\vfill
\eject
{\bf Figure Captions}
Figure 1: {\it HST} observations of the host galaxies used to measure
$H_0$. The data employed to observe Cepheids in 8 SN~Ia hosts and
NGC\,4258 have been collected over 15~yr with 4 cameras over $\sim
500$ orbits of {\it HST} time. Two-month long campaigns in $F555W$
and $F814W$ were initially used to discover Cepheids from their light
curves. Subsequent follow-up observations in $F555W$ enabled the
discovery of Cepheids with $P > 60$~d. Near-IR follow-up data have
been used to reduce the effects of host-galaxy extinction and
sensitivity to metallicity.
Figure 2: {\it HST} images of NGC\,5584. The positions of Cepheids
with periods in the range $P > 60$~d, $30 < P < 60$~d, and $10 < P <
30$~d are indicated by red, blue, and green circles, respectively. A
yellow circle indicates the position of the host galaxy's SN~Ia.
The orientation is indicated by the compass rose whose vectors have
lengths of 15$\arcsec$. The black and white regions of the images
show the WFC3 optical data and the color includes the WFC3-IR data.
Figure 3. As in Figure 2, for NGC\,4038/4039.
Figure 4: {\it HST} WFC3-$F160W$ image of NGC\,3370. Upper panel: The
positions of Cepheids with periods in the range $P > 60$~d, $30 < P < 60$~d,
and $10 < P < 30$~d are indicated by red, blue, and green circles,
respectively. A yellow circle indicates the position of the host galaxy's
SN~Ia. The orientation is indicated by the compass rose whose vectors have
lengths of 15$\arcsec$. The fields of view for the NIC2 follow-up fields from
\citet{riess09a} are indicated. Lower Panel: close-up showing the field of
NGC3370-blue as observed with WFC3-IR (left) and with 4.7 times more exposure
time with NIC2 (right).
Figure 5: Example of scene modeling for the $\sim 1''$ surrounding
typical short, medium, and long-period Cepheids in one WFC3 field,
NGC\,5584. For each Cepheid, the stamp on the left shows the region
around the Cepheid, the middle stamp shows the model of the stellar
sources, and the right stamp is the residual of the image minus the
model. The position of the Cepheid as determined from the optical
data is indicated by the circle.
Figure 6: WFC3-IR versus NIC2 $F160W$ Cepheid photometry. Some of the apparent
dispersion results from the random phases of the Cepheids observed with WFC3.
Figure 7: Near-IR Cepheid period-luminosity relations. For the 8
SN~Ia hosts and the distance-scale anchor, NGC\,4258, the Cepheid
magnitudes are from the same instrument and filter combination, WFC3
$F160W$. This uniformity allows for a significant reduction in
systematic error when utilizing the difference in these relations
along the distance ladder. The measured metallicity for all of the
Cepheids is comparable to solar (log\,[O/H] $\approx 8.9$). A single
slope has been fit to the relations and is shown as the solid line.
20\% of the objects were outliers from the relations (open diamonds)
and are flagged as such for the subsequent analysis. Filled points
with asterisks indicate Cepheids whose periods are shorter than the
incompleteness limit identified from their optical detection.
Figure 8: Relative distances from Cepheids and SNe~Ia. The bottom
abscissa shows the peak apparent visual magnitude of each SN~Ia (red
points) corrected for reddening and to the fiducial brightness (using
the luminosity vs. light-curve shape relations), $m_V^0$. The top
abscissa includes the intercept of the $m_V^0$--log\,$cz$ relation for
SNe~Ia, $a_v$ to provide SN~Ia distance measures, $m_V^0 + 5a_v$,
quantities which are independent of the choice of a fiducial SN Ia. The
right-hand ordinate shows the relative distances between the hosts
determined from the Cepheid $VIH$ Wesenheit relations. The left
ordinate shows the same thing, with the addition of the independent
geometric distance to NGC\,4258 (blue point) based on its
circumnuclear megamasers. The contribution of the nearby SN~Ia and
Cepheid data to $H_0$ can be expressed as a determination of
$m_{V,4258}^0$, the theoretical mean of 8 fiducial SNe~Ia in
NGC\,4258.
Figure 9: Uncertainties in the determination of the Hubble constant.
Uncertainties are squared to show their contribution to the quadrature sum.
These terms are given in Table 5.
Figure 10: Confidence regions in the plane of $H_0$ and the equation-of-state
parameter of dark energy, $w$ and neutrino properties. The localization of the
third acoustic peak in the WMAP 7-year data \citep{komatsu11} produces a
confidence region which is narrow but highly degenerate with the dark energy
equation of state (upper panel). The improved measurement of $H_0$, \hofin,
from the SH0ES program is complementary to the WMAP constraint, resulting in a
determination of $w = -1.08 \pm 0.10$ assuming a constant $w$. This result is
comparable in precision to determinations of $w$ from baryon acoustic
oscillations and high-redshift SNe~Ia, but is independent of both. The inner
regions are 68\% confidence and the outer regions are 95\% confidence. The
modest tilt of the SH0ES measurement of 0.2\% in $H_0$ for a change in $w$ of
0.1 shown as the dotted lines in the upper panel, results from the mild
dependence of $a_v$ on $w$, corresponding to the change in $H$ for changes from
$w=-1$ at the mean SN rdshift of $z = 0.04$. The measurement of $H_0$ is made
at $j_0 = 1$ (i.e., $w = -1$). Constraints on the mass and number of
relativisitic species (e.g., neutrinos) are shown in the middle and lower
panels, respectively.
\bibliographystyle{apj}
|
\section{Algorithm}
\subsection{Computing $q$-gram Frequencies on Uncompressed Strings}
\label{subsection:uncompressed}
We describe two algorithms (Algorithm~\ref{algo:naive}
and Algorithm~\ref{algo:naive_sa}) for computing
the $q$-gram frequencies of a given uncompressed string $T$.
A na\"ive algorithm for computing the
$q$-gram frequencies is given in Algorithm~\ref{algo:naive}.
The algorithm constructs an associative array, where keys consist
of $q$-grams, and the values correspond to the occurrence frequencies
of the $q$-grams.
The time complexity depends on the implementation of the associative
array,
but requires at least $O(q|T|)$ time
since each $q$-gram is considered explicitly, and
the associative array is accessed $O(|T|)$ times:
e.g. $O(q|T|\log|\Sigma|)$ time and $O(q|T|)$ space using a
simple trie.
\begin{algorithm2e}[t]
\SetKwInput{KwOut}{Report}
\SetKw{KwReport}{Report}
\KwIn{string $T$, integer $q \geq 1$}
\KwOut{$(P, |\mathit{Occ}(T,P)|)$ for all $P\in\Sigma^q$ where
$\mathit{Occ}(T,P) \neq \emptyset$.}
$\mathbf{S} \leftarrow \emptyset$\tcp*[l]{empty associative array}
\For{$i\leftarrow 1$ \KwTo $|T|-q+1$}{
$\mathit{qgram} \leftarrow T[i:i+q-1]$\;
\lIf{$\mathit{qgram} \in\mathrm{keys}(\mathbf{S})$}{\label{algo:naive:increment}
$\mathbf{S}[\mathit{qgram}] \leftarrow
\mathbf{S}[\mathit{qgram}] + 1$\;
}
\lElse{
$\mathbf{S}[\mathit{qgram}] \leftarrow 1$\tcp*[l]{new $q$-gram}\label{algo:naive:addweight}
}
}
\lFor{$\mathit{qgram}\in\mathrm{keys}(\mathbf{S})$}{\KwReport{$(\mathit{qgram},\mathbf{S}[\mathit{qgram}])$}}
\caption{A na\"ive algorithm for computing $q$-gram frequencies.}
\label{algo:naive}
\end{algorithm2e}
\begin{algorithm2e}[t]
\SetKwInput{KwOut}{Report}
\SetKw{KwReport}{Report}
\SetKw{KwOr}{or}
\SetKw{KwAnd}{and}
\KwIn{string $T$, integer $q \geq 1$}
\KwOut{
$(i, |\mathit{Occ}(T,P)|)$
for all $P\in\Sigma^q$ and some position $i\in \mathit{Occ}(T,P)$.
}
$\mathit{SA} \leftarrow \mathit{SUFFIXARRAY}(T)$;
$\mathit{LCP} \leftarrow \mathit{LCPARRAY}(T,SA)$;
$\mathit{count} \leftarrow 1$\;
\For{$i\leftarrow 2$ \KwTo $|T|+1$}{
\If{$i = |T|+1$ \KwOr $\mathit{LCP}[i] < q$}{
\lIf{$\mathit{count} > 0$}{
\KwReport{$(\mathit{SA}[i-1], \mathit{count})$}; $\mathit{count} \leftarrow 0$\;
}
}
\lIf{$i \leq |T|$ \KwAnd $\mathit{SA}[i] \leq |T| - q + 1$}{
$\mathit{count} \leftarrow \mathit{count}+1$\;\label{algo:naive_sa:increment}
}
}
\caption{A linear time algorithm for computing $q$-gram frequencies.}
\label{algo:naive_sa}
\end{algorithm2e}
The $q$-gram frequencies of string $T$ can be calculated in
$O(|T|)$ time using suffix array $\mathit{SA}$ and lcp array $\mathit{LCP}$,
as shown in Algorithm~\ref{algo:naive_sa}.
For each $1\leq i\leq |T|$, the suffix $\mathit{SA}[i]$ represents
an occurrence of $q$-gram $T[\mathit{SA}[i]:\mathit{SA}[i]+q-1]$,
if the suffix is long enough, i.e. $\mathit{SA}[i] \leq |T| - q +1$.
The key is that since the suffixes are lexicographically
sorted, intervals on the suffix array where the values in the lcp array
are at least $q$ represent occurrences of the same $q$-gram.
The algorithm runs in $O(|T|)$ time, since $\mathit{SA}$ and $\mathit{LCP}$ can be
constructed in $O(|T|)$. The rest is a simple $O(|T|)$ loop.
A technicality is that we encode the output for a $q$-gram as
one of the positions in the text where the $q$-gram occurs,
rather than the $q$-gram itself.
This is because there can be a total of $O(|T|)$ different $q$-grams,
and if we output them as length-$q$ strings,
it would require at least $O(q|T|)$ time.
\subsection{Computing $q$-gram Frequencies on SLP}
We now describe the core idea of our algorithms,
and explain two variations which utilize variants of the two algorithms for uncompressed strings
presented in Section~\ref{subsection:uncompressed}.
For $q = 1$, the $1$-gram frequencies are simply the frequencies of
the alphabet and the output is
$(a,\sum\{\mathit{vOcc}(X_i)\mid X_i = a\})$ for each $a\in\Sigma$,
which takes only $O(n)$ time.
For $q \geq 2$, we make use of Lemma~\ref{lemma:mkSLP} below.
The idea is similar to the {\em $mk$ Lemma}~\cite{charikar05:_small_gramm_probl},
but the statement is more specific.
\begin{lemma}
\label{lemma:mkSLP}
Let $\mathcal{T} = \{X_i\}_{i=1}^n$ be an SLP that represents string $T$.
For an interval $[u:v]$ $(1\leq u < v \leq |T|)$,
there exists exactly one variable $X_i=X_\ell X_r$
such that for some $[u':v'] \in \mathit{itv}(X_i)$,
the following holds:
$[u:v] \subseteq [u':v']$,
$u \in [u':u'+|X_\ell|-1] \in \mathit{itv}(X_\ell)$
and
$v \in [u'+|X_\ell|:v'] \in \mathit{itv}(X_r)$.
\end{lemma}
\begin{proof}
Consider length $1$ intervals $[u:u]$ and $[v:v]$
corresponding to leaves in the derivation tree.
$X_i$ corresponds to the lowest common ancestor
of these intervals in the derivation tree.
\hfill $\Box$
\end{proof}
\begin{wrapfigure}[9]{r}{0.36\textwidth}
\vspace{-1.43cm}
\centerline{\includegraphics[width=0.35\textwidth]{SLP_ngram.eps}}
\caption{
Length-$q$ intervals corresponding to $X_i = X_\ell X_r$.
}
\label{fig:SLP-kgram}
\end{wrapfigure}
From Lemma~\ref{lemma:mkSLP}, each occurrence of
a $q$-gram ($q \geq 2$) represented by some length-$q$ interval of $T$,
corresponds to a single variable $X_i = X_{\ell}X_r$, and is split in
two by intervals corresponding to $X_{\ell}$ and $X_r$.
On the other hand, consider all length-$q$ intervals that
correspond to a given variable. Counting the frequencies of the $q$-grams
they represent, and summing them up for all variables give the
frequencies of all $q$-grams of $T$.
For variable $X_i = X_\ell X_r$, let
$t_i = \mathit{suf}(X_\ell, q-1) \mathit{pre}(X_r,q-1)$.
Then, all $q$-grams represented by length $q$ intervals that
correspond to $X_i$ are those in $t_i$. (Fig.~\ref{fig:SLP-kgram}).
If we obtain the frequencies of all $q$-grams in $t_i$, and then
multiply each frequency by $\mathit{vOcc}(X_i)$,
we obtain frequencies for the $q$-grams occurring in all intervals
derived by $X_i$.
It remains to sum up the $q$-gram frequencies of $t_i$
for all $1\leq i\leq n$.
We can regard it as obtaining the {\em weighted} $q$-gram
frequencies in the set of strings $\{t_1,\ldots,t_n\}$,
where each $q$-gram in $t_i$ is weighted by $\mathit{vOcc}(X_i)$.
We further reduce this problem to a weighted $q$-gram frequency
problem for a single string $z$ as in Algorithm~\ref{algo:slpmain}.
String $z$ is constructed by concatenating $t_i$
such that $q \leq |t_i| \leq 2(q-1)$,
and the weights of $q$-grams starting at each position in $z$
is held in array $w$.
On line~\ref{algo:slpmain:append0}, $0$'s instead of $\mathit{vOcc}(X_i)$ are appended to
$w$ for the last $q-1$ values corresponding to $t_i$.
This is to avoid counting unwanted $q$-grams that are generated by the
concatenation of $t_i$ to $z$ on line~\ref{algo:slpmain:zappend},
which are not substrings of each $t_i$.
The weighted $q$-gram frequency problem for a single string
(Line~\ref{algo:slpmainweightedfreqs}) can be solved
with a slight modification of Algorithm~\ref{algo:naive} or \ref{algo:naive_sa}.
The modified algorithms are shown respectively in
Algorithms~\ref{algo:weighted_naive} and \ref{algo:weighted_naive_sa}.
\begin{theorem}
Given an SLP ${\mathcal T} = \{X_i\}_{i=1}^n$ of size $n$
representing a string $T$,
the $q$-gram frequencies of $T$ can be computed in $O(qn)$ time
for any $q > 0$.
\end{theorem}
\begin{proof}
Consider Algorithm~\ref{algo:slpmain}.
The correctness is straightforward from the above arguments, so we
consider the time complexity.
Line~\ref{algo:slpmain:varocc} can be computed in $O(n)$ time.
Line~\ref{algo:slpmain:prefsuf} can be computed in $O(qn)$ time by a simple dynamic
programming.
For $\mathit{pre}()$:
If $X_i=a$ for some $a\in\Sigma$, then $\mathit{pre}(X_i,q-1) = a$.
If $X_i=X_{\ell}X_r$ and $|X_{\ell}| \geq q-1$, then $\mathit{pre}(X_i, q-1) = \mathit{pre}(X_{\ell}, q-1)$.
If $X_i=X_{\ell}X_r$ and $|X_{\ell}| < q-1$, then $\mathit{pre}(X_i, q-1)
= \mathit{pre}(X_{\ell}, q-1)\mathit{pre}(X_r, q - 1 - |X_{\ell}|)$.
The strings $\mathit{suf}()$ can be computed similarly.
The computation amounts to copying $O(q)$ characters for each
variable, and thus can be done in $O(qn)$ time.
For the loop at line~\ref{algo:slpmain:mainloop}, since the length of
string $t_i$ appended to $z$, as well as the number of elements
appended to $w$ is at most $2(q-1)$ in each loop,
the total time complexity is $O(qn)$.
Finally, since the length of $z$ and $w$ is $O(qn)$,
line~\ref{algo:slpmainweightedfreqs} can be calculated in $O(qn)$
time
using the weighted version of Algorithm~\ref{algo:naive_sa} (Algorithm~\ref{algo:weighted_naive_sa}).
\hfill $\Box$
\end{proof}
Note that the time complexity for using the weighted version of
Algorithm~\ref{algo:naive} for line~\ref{algo:slpmainweightedfreqs}
of Algorithm~\ref{algo:slpmain} would be at least $O(q^2n)$:
e.g. $O(q^2n\log|\Sigma|)$ time and $O(q^2n)$ space using a trie.
\begin{algorithm2e}[t]
\caption{Calculating $q$-gram frequencies of an SLP for $q\geq 2$}
\label{algo:slpmain}
\SetKw{KwAnd}{and}
\SetKwInput{KwOut}{Report}
\KwIn{SLP ${\mathcal T} = \{X_i\}_{i=1}^n$ representing string $T$, integer $q\geq 2$.}
\KwOut{all $q$-grams and their frequencies which occur in $T$.}
\SetKw{KwReport}{Report}
Calculate $\mathit{vOcc}(X_i)$ for all $1\leq i\leq n$\; \label{algo:slpmain:varocc}
Calculate $\mathit{pre}(X_i,q-1)$ and $\mathit{suf}(X_i,q-1)$ for all $1\leq
i\leq n-1$ \; \label{algo:slpmain:prefsuf}
$z \leftarrow \varepsilon$; $w \leftarrow []$\;
\For{$i \leftarrow 1$ \KwTo $n$}{\label{algo:slpmain:mainloop}
\If{$X_i = X_{\ell}X_r$ \KwAnd $|X_i| \geq q$}{
$t_i = \mathit{suf}(X_\ell,q-1)\mathit{pre}(X_r,q-1)$;
$z$.append($t_i$)\; \label{algo:slpmain:zappend}
\lFor{$j \leftarrow 1$ \KwTo $|t_i|-q+1$}{
$w$.append($\mathit{vOcc}(X_i)$)\;
}
\lFor{$j \leftarrow 1$ \KwTo $q-1$}{
$w$.append(0)\; \label{algo:slpmain:append0}
}
}
}
\KwReport
$q$-gram frequencies in $z$, where each $q$-gram
$z[i:i+q-1]$
is {\em weighted} by $w[i]$.\label{algo:slpmainweightedfreqs}
\end{algorithm2e}
\begin{algorithm2e}[t]
\caption{A variant of Algorithm~\ref{algo:naive} for weighted $q$-gram frequencies.}
\label{algo:weighted_naive}
\SetKwInput{KwOut}{Report}
\SetKw{KwReport}{Report}
\KwIn{string $T$, array of integers $w$ of length $|T|$, integer $q \geq 1$}
\KwOut{$(P, \sum_{i\in\mathit{Occ}(T,P)}w[i])$ for all $P\in\Sigma^q$ where
$\sum_{i\in\mathit{Occ}(T,P)}w[i] > 0$.}
$\mathbf{S} \leftarrow \emptyset$\tcp*[l]{empty associative array}
\For{$i\leftarrow 1$ \KwTo $|T|-q+1$}{
$\mathit{qgram} \leftarrow T[i:i+q-1]$\;
\lIf{$\mathit{qgram} \in\mathrm{keys}(\mathbf{S})$}{
$\mathbf{S}[\mathit{qgram}] \leftarrow
\mathbf{S}[\mathit{qgram}] + w[i]$\;
}
\lElse{\lIf{$w[i]>0$}{$\mathbf{S}[\mathit{qgram}] \leftarrow w[i]$\tcp*[l]{new $q$-gram}}}
}
\lFor{$\mathit{qgram}\in\mathrm{keys}(\mathbf{S})$}{\KwReport{$(\mathit{qgram},\mathbf{S}[\mathit{qgram}])$}}
\end{algorithm2e}
\begin{algorithm2e}[t]
\caption{A variant of Algorithm~\ref{algo:naive_sa} for weighted
$q$-gram frequencies.}
\label{algo:weighted_naive_sa}
\SetKwInput{KwOut}{Output}
\SetKw{KwReport}{Report}
\SetKw{KwOr}{or}
\SetKw{KwAnd}{and}
\KwIn{string $T$, array of integers $w$ of length $|T|$, integer $q \geq 1$}
\KwOut{$(i, \sum_{i\in\mathit{Occ}(T,P)}w[i])$ for all $P\in\Sigma^q$ where
$\sum_{i\in\mathit{Occ}(T,P)}w[i] > 0$
and some position $i\in\mathit{Occ}(T,P)$.}
$\mathit{SA} \leftarrow \mathit{SUFFIXARRAY}(T)$;
$\mathit{LCP} \leftarrow \mathit{LCPARRAY}(T,SA)$;
$\mathit{count} \leftarrow 1$\;
\For{$i\leftarrow 2$ \KwTo $|T|+1$}{
\If{$i = |T|+1$ \KwOr $\mathit{LCP}[i] < q$}{
\lIf{$\mathit{count} > 0$}{
\KwReport{$(\mathit{SA}[i-1], \mathit{count})$}; $\mathit{count} \leftarrow 0$\;
}
}
\lIf{$i \leq |T|$ \KwAnd $\mathit{SA}[i] \leq |T| - q + 1$}{
$\mathit{count} \leftarrow \mathit{count}+w[\mathit{SA}[i]]$\;
}
}
\end{algorithm2e}
\section{Applications and Extensions}
We showed that for an SLP ${\mathcal{T}}$ of size $n$ representing string $T$,
$q$-gram frequency problems on $T$ can be reduced to
{\em weighted} $q$-gram frequency problems on a string $z$ of length $O(qn)$,
which can be much shorter than $T$.
This idea can further be applied to obtain efficient compressed
string processing algorithms for
interesting problems which we briefly introduce below.
\subsection{$q$-gram Spectrum Kernel}
A string kernel is a function that computes the inner product between two strings
which are mapped to some feature space.
It is used when classifying string or text
data using methods such as Support Vector Machines (SVMs), and is usually the
dominating factor in the time complexity of SVM learning and classification.
A $q$-gram spectrum kernel~\cite{leslie02:_spect_kernel} considers the
feature space of $q$-grams.
For string $T$, let $\phi_q(T) = (|\mathit{Occ}(T,p)|)_{p\in\Sigma^q}$.
The kernel function is defined as
$K_q(T_1,T_2) =
\langle\phi_q(T_1),\phi_q(T_2)\rangle =
\sum_{p\in\Sigma^q} |\mathit{Occ}(T_1,p)| |\mathit{Occ}(T_2,p)|$.
The calculation of the kernel function amounts to
summing up the product of occurrence frequencies
in strings $T_1$ and $T_2$ for all
$q$-grams which occur in both $T_1$ and $T_2$.
This can be done in $O(|T_1|+|T_2|)$ time using suffix arrays.
For two SLPs ${\mathcal T_1}$ and ${\mathcal T_2}$ of
size $n_1$ and $n_2$ representing strings $T_1$ and $T_2$,
respectively, the $q$-gram spectrum kernel $K_q(T_1,T_2)$ can be computed
in $O(q(n_1+n_2))$ time by a slight modification of our algorithm.
\subsection{Optimal Substring Patterns of Length $q$}
Given two sets of strings, finding string patterns that are
frequent in one set and not in the other, is an important
problem in string data mining, with many problem formulations and the
types of patterns to be considered,
e.g.: in Bioinformatics~\cite{brazma98:_approac},
Machine Learning (optimal patterns~\cite{arimura98:_fast_algor_discov_optim_strin}),
and more recently KDD
(emerging patterns~\cite{chan03:_minin_emerg_subst}).
A simple optimal $q$-gram pattern discovery problem can be defined as follows:
Let $\mathbf{T_1}$
and $\mathbf{T_2}$
be two multisets of strings.
The problem is to find the $q$-gram $p$ which gives
the highest (or lowest) score according to some scoring function
that depends only on $|\mathbf{T_1}|$, $|\mathbf{T_2}|$,
and the number of strings respectively in $\mathbf{T_1}$ and $\mathbf{T_2}$ for
which $p$ is a substring.
For uncompressed strings, the problem can be solved in $O(N)$ time,
where $N$ is the total length of the strings in both $\mathbf{T_1}$ and
$\mathbf{T_2}$, by applying the algorithm of~\cite{HuiCPM92} to two sets of strings.
For the SLP compressed version of this problem,
the input is two multisets of SLPs,
each representing strings in $\mathbf{T_1}$ and $\mathbf{T_2}$.
If $n$ is the total number of variables used in all
of the SLPs, the problem can be solved in $O(qn)$ time.
\subsection{Different Lengths}
The ideas in this paper can be used to consider all substrings of length {\em not only} $q$,
but {\em all lengths up-to} $q$, with some modifications.
For the applications discussed above,
although the number of such substrings increases to $O(q^2n)$,
the $O(qn)$ time complexity can be maintained
by using standard techniques of suffix
arrays~\cite{gusfield97:_algor_strin_trees_sequen,Kasai01}.
This is because there exist only $O(qn)$ substring with distinct frequencies
(corresponding to nodes of the suffix tree),
and the computations of the extra substrings can be summarized with
respect to them.
\ignore{
\subsection{Collage System}
Collage system~\cite{KidaCollageTCS} is a more general framework for modeling
various compression methods.
In addition to the simple concatenation operation used in SLPs,
it includes operations for repetition and prefix/suffix truncation
of variables.
For example, while a LZ77 encoded representation of size $m$ may require
$O(m^2\log m)$ size when represented as an SLP,
it can be represented as a collage system of size $O(m\log m)$~\cite{GasieniecSWAT96}.
Our algorithm can be extended to run in $O((q+h)n)$ time on
collage system of size $n$, where $h$ is the height of the derivation tree.
The increase in complexity comes from the handling of truncated
variables. Details will be presented in a forthcoming paper.
}
\section{Conclusion}
We presented an $O(qn)$ time and space algorithm for calculating all
$q$-gram frequencies in a string, given an SLP of size $n$
representing the string.
This solves, much more efficiently, a more general problem than
considered in previous work.
Computational experiments on various real texts showed that the algorithms
run faster than algorithms that work on the
uncompressed string, when $q$ is small.
Although larger values of $q$ allow us to capture longer character dependencies,
the dimensionality of the features increases, making the space of
occurring $q$-grams sparse.
Therefore, meaningful values of $q$ for
typical applications can be fairly small in practice (e.g. $3\sim 6$),
so our algorithms have practical value.
A future work is extending our algorithms that work on SLPs,
to algorithms that work on collage systems~\cite{KidaCollageTCS}.
A Collage System is a more general framework for modeling various
compression methods.
In addition to the simple concatenation operation used in SLPs,
it includes operations for repetition and prefix/suffix truncation
of variables.
\ignore{
For example, while a LZ77 encoded representation of size $m$ may require
$O(m^2\log m)$ size when represented as an SLP,
it can be represented as a collage system of size $O(m\log m)$~\cite{GasieniecSWAT96}.
}
This is the first paper to show the potential of the compressed
string processing approach in developing efficient and {\em practical} algorithms
for problems in the field of string mining and classification.
More and more efficient algorithms for various processing of
text in compressed representations are becoming available.
We believe texts will eventually be stored in compressed form by default,
since not only will it save space, but it will also have the added benefit of being able to conduct
various computations on it more efficiently later on, when needed.
\section{Computational Experiments}
We implemented 4 algorithms (NMP, NSA, SMP, SSA) that
count the frequencies of all $q$-grams in a given text.
NMP (Algorithm~\ref{algo:naive}) and NSA (Algorithm~\ref{algo:naive_sa})
work on the uncompressed text.
SMP (Algorithm~\ref{algo:slpmain} +
Algorithm~\ref{algo:weighted_naive})
and
SSA (Algorithm~\ref{algo:slpmain} +
Algorithm~\ref{algo:weighted_naive_sa})
work on SLPs.
The algorithms were implemented using the C++ language.
We used {\tt std::map} from the Standard Template Library (STL)
for the associative array implementation.
\footnote{We also used {\tt std::hash\_map} but omit the results due
to lack of space. Choosing the hashing function to use is difficult,
and we note that its performance was unstable and sometimes very bad
when varying $q$.}
For constructing suffix arrays, we used the divsufsort
library\footnote{\url{http://code.google.com/p/libdivsufsort/}}
developed by Yuta Mori. This implementation is not linear time
in the worst case, but has been empirically shown to be one of
the fastest implementations on various data.
All computations were conducted on a Mac Xserve (Early 2009)
with 2 x 2.93GHz Quad Core Xeon processors and 24GB Memory,
only utilizing a single process/thread at once.
The program was compiled using the GNU C++ compiler ({\tt g++}) 4.2.1
with the {\tt -fast} option for optimization.
The running times are measured in seconds, starting from after reading
the uncompressed text into memory for NMP and NSA, and
after reading the SLP that represents the text into memory for SMP and
SSA.
Each computation is repeated at least 3 times, and the average is taken.
\subsection{Fibonacci Strings}
\begin{wrapfigure}[12]{r}{0.48\textwidth}
\vspace{-1.2cm}
\begin{center}
\includegraphics[width=0.47\textwidth]{graph_data_exp_fib.32.0115.csvk=50.eps}
\caption{Running times of NMP, NSA, SMP, SSA on Fibonacci strings
for $q=50$.}
\label{fig:graph_fib}
\end{center}
\end{wrapfigure}
The $i$ th Fibonacci string $F_i$ can be represented by the following
SLP:
$X_1 = \mathtt{b}$, $X_2 = \mathtt{a}$,
$X_i = X_{i-1} X_{i-2}$ for $i > 2$,
and $F_i = \mathit{val}(X_i)$.
Fig.~\ref{fig:graph_fib} shows the running times on
Fibonacci strings $F_{20}, F_{25}, \ldots, F_{95}$, for $q=50$.
Although this is an extreme case since Fibonacci strings can
be exponentially compressed, we can see that SMP and
SSA that work on the SLP are clearly faster than NMP and NSA which
work on the uncompressed string.
\subsection{Pizza \& Chili Corpus}
\begin{wrapfigure}[14]{r}{0.5\textwidth}
\vspace{-1.0cm}
\begin{center}
\includegraphics[width=0.47\textwidth]{speedUp_pizza.eps}
\caption{Time ratios NMP/SMP and NSA/SSA
plotted against ratio $|z|/|T|$.}
\label{fig:speed_up}
\end{center}
\end{wrapfigure}
We also applied the algorithms on texts
XML, DNA, ENGLISH, and PROTEINS, with sizes 50MB, 100MB, and 200MB,
obtained from the Pizza \& Chili
Corpus\footnote{\url{http://pizzachili.dcc.uchile.cl/texts.html}}.
We used RE-PAIR~\cite{LarssonDCC99} to obtain SLPs for this data.
Table~\ref{table:pizza} shows the running times for all algorithms and
data, where $q$ is varied from $2$ to $10$.
We see that for all corpora,
SMP and SSA running on SLPs are actually faster than
NMP and NSA running on uncompressed text, when $q$ is small.
Furthermore, SMP is faster than SSA when $q$ is smaller.
Interestingly for XML, the SLP versions are faster even for $q$
up to $9$.
Fig.~\ref{fig:speed_up} shows the same results
as time ratio: NMP/SMP and NSA/ SSA, plotted against ratio:
(length of $z$ in Algorithm~\ref{algo:slpmain})/(length of uncompressed text).
As expected, the SLP versions are basically faster than their uncompressed
counterparts, when $|z|/\mbox{(text length)}$ is less than $1$,
since the SLP versions run the weighted versions of the uncompressed algorithms
on a text of length $|z|$.
SLPs generated by other grammar based compression algorithms showed
similar tendencies (data not shown).
\begin{table}[t]
\caption{Running times in seconds for data from the Pizza \& Chili Corpus.
Bold numbers represent the fastest time for each data and $q$.
Times for SMP and SSA are prefixed with $\triangleright$,
if they become fastest when all algorithms start from the SLP representation,
i.e., NMP and NSA require time for decompressing the SLP (denoted
by decompression time).
The bold horizontal lines show the boundary where $|z|$ in
Algorithm~\ref{algo:slpmain}
exceeds the uncompressed text length.
}
\label{table:pizza}
\begin{center}
\scriptsize
\setlength{\tabcolsep}{1pt}
\renewcommand{\rmdefault}{ptm}
\renewcommand{\sfdefault}{phv}
\renewcommand{\ttdefault}{pcr}
\normalfont
\input{table1.tex}
\end{center}
\end{table}
\section{Introduction}
A major problem in managing large scale string data is its sheer size.
Therefore, such data is normally stored in compressed form.
In order to utilize or analyze the data afterwards,
the string is usually decompressed, where we must again confront the
size of the data.
To cope with this problem, algorithms that work directly on compressed
representations of strings without explicit decompression
have gained attention, especially for the string pattern matching problem~\cite{amir92:_effic_two_dimen_compr_match}
where
algorithms on compressed text can actually run faster than
algorithms on the uncompressed text~\cite{ShibataCIAC2000}.
There has been growing interest in what problems can be
efficiently solved in this kind of
setting~\cite{lifshits07:_proces_compr_texts,hermelin09:_unified_algor_accel_edit_distan}.
Since there exist many different text compression schemes,
it is not realistic to develop different algorithms for each scheme.
Thus, it is common to consider algorithms on texts represented as
{\em straight line programs} (SLPs)~\cite{NJC97,lifshits07:_proces_compr_texts,hermelin09:_unified_algor_accel_edit_distan}.
An SLP is a context free grammar in the Chomsky normal form that
derives a single string.
Texts compressed by any grammar-based compression algorithms
(e.g.~\cite{SEQUITUR,LarssonDCC99}) can be represented as SLPs,
and those compressed by the LZ-family (e.g.~\cite{LZ77,LZ78}) can be quickly
transformed to SLPs~\cite{rytter03:_applic_lempel_ziv}.
Recently, even {\em compressed self-indices} based on SLPs have
appeared~\cite{claudear:_self_index_gramm_based_compr},
and SLPs are a promising representation of compressed strings
for conducting various operations.
In this paper, we explore a more advanced field of
application for compressed string processing:
mining and classification on
string data given in compressed form.
Discovering useful patterns hidden in strings as well as
automatic and accurate classification of
strings into various groups, are important problems in
the field of data mining and machine learning with many
applications.
As a first step toward {\em compressed} string mining and
classification, we consider the problem of
finding the occurrence frequencies for all
$q$-grams contained in a given string.
$q$-grams are important features of string data,
widely used for this purpose in many fields such as text and natural
language processing, and bioinformatics.
In~\cite{inenaga09:_findin_charac_subst_compr_texts},
an $O(|\Sigma|^2n^2)$-time $O(n^2)$-space algorithm
for finding the {\em most frequent} $2$-gram from an
SLP of size $n$ representing text $T$ over alphabet $\Sigma$ was presented.
In~\cite{claudear:_self_index_gramm_based_compr},
it is mentioned
that the most frequent $2$-gram can be found in $O(|\Sigma|^2n\log n)$-time
and $O(n\log|T|)$-space, if the SLP is pre-processed and a self-index is built.
It is possible to extend these two algorithms
to handle $q$-grams for $q > 2$, but would respectively require
$O(|\Sigma|^qqn^2)$ and $O(|\Sigma|^qqn\log n)$ time,
since they must essentially enumerate and count the occurrences of all
substrings of length $q$,
regardless of whether the $q$-gram occurs in the string.
Note also that any algorithm that works on the uncompressed text
$T$ requires exponential time in the worst case, since $|T|$ can be as
large as $O(2^n)$.
The main contribution of this paper is an $O(qn)$ time and space
algorithm that computes the occurrence frequencies for {\em all}
$q$-grams in the text, given
an SLP of size $n$ representing the text.
Our new algorithm solves the more general problem
and greatly improves the computational complexity compared to previous work.
We also conduct computational experiments on various real texts,
showing that when $q$ is small, our algorithm and its variation actually run faster
than algorithms that work on the uncompressed text.
Our algorithms have profound applications in the field of string mining
and classification, and several applications and extensions are discussed.
For example, our algorithm leads to an $O(q(n_1+n_2))$ time algorithm for
computing the $q$-gram spectrum kernel~\cite{leslie02:_spect_kernel}
between SLP compressed texts of size $n_1$ and $n_2$.
It also leads to an $O(qn)$ time algorithm for finding
the optimal $q$-gram (or emerging $q$-gram) that discriminates between
two sets of SLP compressed strings, when $n$ is the total size of the
SLPs.
\subsubsection{Related Work}
There exist many works on {\em compressed text
indices}~\cite{navarro07:_compr},
but the main focus there is on fast search for a {\em given} pattern.
The compressed indices basically replace or simulate operations on
uncompressed indices using a smaller data structure.
Indices are important for efficient string processing,
but note that simply replacing the underlying index used in a mining
algorithm will generally increase time complexities of the algorithm
due to the extra overhead required to access the compressed index.
On the other hand, our approach is a new mining algorithm which exploits
characteristics of the compressed representation to achieve faster
running times.
Several algorithms for finding characteristic sequences from compressed texts have been proposed, e.g.,
finding the longest common substring of two strings~\cite{matsubara_tcs2009},
finding all palindromes~\cite{matsubara_tcs2009},
finding most frequent substrings~\cite{inenaga09:_findin_charac_subst_compr_texts}, and
finding the longest repeating substring~\cite{inenaga09:_findin_charac_subst_compr_texts}.
However, none of them have reported results of computational
experiments, implying that this paper is the first to show the
practical usefulness of a compressed text mining algorithm.
\input{preliminaries.tex}
\input{algorithm.tex}
\input{applications.tex}
\input{experiments.tex}
\input{conclusion.tex}
\bibliographystyle{splncs03}
\section{Preliminaries}
Let $\Sigma$ be a finite {\em alphabet}.
An element of $\Sigma^*$ is called a {\em string}.
For any integer $q > 0$, an element of $\Sigma^q$ is called an \emph{$q$-gram}.
The length of a string $T$ is denoted by $|T|$.
The empty string $\varepsilon$ is a string of length 0,
namely, $|\varepsilon| = 0$.
For a string $T = XYZ$, $X$, $Y$ and $Z$ are called
a \emph{prefix}, \emph{substring}, and \emph{suffix} of $T$, respectively.
The $i$-th character of a string $T$ is denoted by $T[i]$ for $1 \leq i \leq |T|$,
and the substring of a string $T$ that begins at position $i$ and
ends at position $j$ is denoted by $T[i:j]$ for $1 \leq i \leq j \leq |T|$.
For convenience, let $T[i:j] = \varepsilon$ if $j < i$.
For a string $T$ and integer $q \geq 0$, let $\mathit{pre}(T,q)$ and $\mathit{suf}(T,q)$
represent respectively, the length-$q$ prefix and suffix of $T$.
That is, $\mathit{pre}(T,q) = T[1:\min(q,|T|)]$ and
$\mathit{suf}(T,q) = T[\max(1,|T|-q+1):|T|]$.
For any strings $T$ and $P$,
let $\mathit{Occ}(T,P)$ be the set of occurrences of $P$ in $T$, i.e.,
$\mathit{Occ}(T,P) = \{k > 0 \mid T[k:k+|P|-1] = P\}$.
The number of elements $|\mathit{Occ}(T,P)|$ is called
the \emph{occurrence frequency} of $P$ in $T$.
\subsection{Straight Line Programs}
\begin{algorithm2e}[t]
\caption{Calculating $\mathit{vOcc}(X_i)$ for all $1\leq i \leq n$.}
\label{algo:varocc}
\KwIn{SLP ${\mathcal T} = \{X_i\}_{i=1}^n$ representing string $T$.}
\KwOut{$\mathit{vOcc}(X_i)$ for all $1\leq i\leq n$}
$\mathit{vOcc}[X_n] \leftarrow 1$\;
\lFor{$i\leftarrow 1$ \KwTo $n-1$}{
$\mathit{vOcc}[X_i] \leftarrow 0$\;
}
\For{$i\leftarrow n$ \KwTo $2$}{
\If{ $X_i = X_{\ell}X_r$ }{
$\mathit{vOcc}[X_{\ell}] \leftarrow \mathit{vOcc}[X_{\ell}] + \mathit{vOcc}[X_i]$;
$\mathit{vOcc}[X_r] \leftarrow \mathit{vOcc}[X_r] + \mathit{vOcc}[X_i]$\;
}
}
\end{algorithm2e}
\begin{wrapfigure}[13]{r}{0.5\textwidth}
\vspace{-1.56cm}
\centerline{\includegraphics[width=0.48\textwidth]{slp.eps}}
\caption{
The derivation tree of
SLP $\mathcal T = \{X_i\}_{i=1}^{7}$ with $X_1 = \mathtt{a}$, $X_2 = \mathtt{b}$, $X_3 = X_1X_2$,
$X_4 = X_1X_3$, $X_5 = X_3X_4$, $X_6 = X_4X_5$, and $X_7 = X_6X_5$,
representing string $T = \mathit{val}(X_7) = \mathtt{aababaababaab}$.
}
\label{fig:SLP}
\end{wrapfigure}
A {\em straight line program} ({\em SLP}) $\mathcal T$ is a sequence of assignments
$X_1 = expr_1, X_2 = expr_2, \ldots, X_n = expr_n$,
where each $X_i$ is a variable and each $expr_i$ is an expression, where
$expr_i = a$ ($a\in\Sigma$), or $expr_i = X_{\ell} X_r$~($\ell,r < i $).
Let $\mathit{val}(X_i)$ represent the string derived from $X_i$.
When it is not confusing, we identify a variable $X_i$
with $\mathit{val}(X_i)$.
Then, $|X_i|$ denotes the length of the string $X_i$ derives.
An SLP $\mathcal{T}$ {\em represents}
the string $T = \mathit{val}(X_n)$.
The \emph{size} of the program $\mathcal T$ is the number $n$ of
assignments in $\mathcal T$. (See Fig.~\ref{fig:SLP})
The substring intervals of $T$ that each variable derives can
be defined recursively as follows: $\mathit{itv}(X_n) = \{ [1:|T|]\}$, and
$\mathit{itv}(X_i) =
\{ [u + |X_\ell|:v] \mid X_k = X_\ell X_i, [u:v] \in \mathit{itv}(X_k) \}
\cup
\{ [u:u+|X_i|-1] \mid X_k = X_i X_r, [u:v] \in \mathit{itv}(X_k) \}
$ for $i < n$.
For example, $\mathit{itv}(X_5) = \{ [4:8], [9:13]\}$ in Fig.~\ref{fig:SLP}.
Considering the transitive reduction of set inclusion,
the intervals $\cup_{i=1}^n \mathit{itv}(X_i)$ naturally form a
binary tree (the derivation tree).
Let $\mathit{vOcc}(X_i) = |\mathit{itv}(X_i)|$ denote
the number of times a variable $X_i$ occurs in the derivation of $T$.
$\mathit{vOcc}(X_i)$ for all $1\leq i\leq n$ can be computed in $O(n)$
time by a simple iteration on the variables,
since $\mathit{vOcc}(X_n) = 1$ and for $i < n$,
$\mathit{vOcc}(X_i) =
\sum \{ \mathit{vOcc}(X_k) \mid X_k = X_\ell X_i \}
+
\sum \{\mathit{vOcc}(X_k) \mid X_k = X_iX_r \}$.
(See Algorithm~\ref{algo:varocc})
\subsection{Suffix Arrays and LCP Arrays}
The suffix array $\mathit{SA}$~\cite{manber93:_suffix} of any string $T$
is an array of length $|T|$ such that
$\mathit{SA}[i] = j$, where $T[j:|T|]$ is the $i$-th lexicographically smallest suffix of $T$.
The \emph{lcp} array of any string $T$ is an array of length $|T|$ such that
$\mathit{LCP}[i]$ is the length of the longest common prefix of
$T[\mathit{SA}[i-1]:|T|]$ and $T[\mathit{SA}[i]:|T|]$ for $2 \leq i \leq |T|$,
and $\mathit{LCP}[1] = 0$.
The suffix array for any string of length $|T|$
can be constructed in $O(|T|)$
time~(e.g.~\cite{Karkkainen_Sanders_icalp03})
assuming an integer alphabet.
Given the text and suffix array, the lcp array can also be calculated
in $O(|T|)$ time~\cite{Kasai01}.
|
\section{Introduction}
Measuring cosmological parameters and the equation of state of dark
energy to high accuracy is the goal of many current and upcoming
experiments (e.g. PS1\footnote{PS1 http://ps1sc.org},
DES\footnote{DES https://www.darkenergysurvey.org/},
Euclid\footnote{Euclid http://sci.esa.int/euclid}, HALO\footnote{HALO
Rhodes et al., in preparation}, LSST\footnote{LSST
http://www.lsst.org/} and WFIRST\footnote{WFIRST
http://wfirst.gsfc.nasa.gov}). Variables such as the size and depth
of the survey (amongst other things) have a significant effect on the
ability of a survey to constrain cosmological parameters.
Consequently, significant effort must be spent in accurately
predicting what these telescopes will see, well before construction
begins. This will allow us to both to influence the design phase and
to understand the capabilities of the instrument once the design has
been set. In order to make predictions for these upcoming missions,
statistical tools must be used to estimate the accuracy they will be
able to achieve. The current standard for prediction uses the Fisher
matrix methodology \citep{tth97,f35}. Traditionally, Fisher matrices
have been generated with data covariance matrices that assume an
underlying Gaussian matter and radiation distribution
\citep{tth97,k97}, which is accurate for CMB estimates when the
Universe was still linear. However, this is not an accurate
representation of the low-redshift Universe at smaller scales. Error
bars generated using these Gaussian assumptions may be biased when
compared with those generated using methods that account for the
non-Gaussian, non-linear nature of the Universe.
In this paper we compare the $1\sigma$, two-parameter $\Omega_m -
\sigma_8$ error estimates from the full maximum likelihood analyses
with a Gaussian Fisher analysis. Using weak gravitational lensing
power spectrum analysis of mock galaxy shear catalogues generated with
the {\small SUNGLASS} pipeline \citep{kht+11}, we show the importance
of using accurate non-linear covariance matrices when estimating
errors for future experiments. Analytic approximations of the
correlation function covariance matrix under these assumptions have
been shown to underestimate the errors on cosmic variance by a factor
of up to $\sim30$, which makes breaking the $\Omega_m - \sigma_8$
degeneracy more difficult \citep{svh+07}. Simulations are able to
provide accurate covariance matrices because they do not make
assumptions about the underlying Gaussianity of the Universe and
consequently, the resulting covariance matrices include the
off-diagonal components.
Fisher matrix analyses are attractive because of their relative speed
and minimal computational requirements when compared with maximum
likelihood estimates. However, this is offset by the loss of
information due to Gaussian assumptions in the generation of the
Fisher matrix. To compensate for some of these Gaussian assumptions,
we propose using a non-Gaussian weak lensing shear covariance matrix
generated from simulations to generate the Fisher matrix. The
resulting Fisher matrix still assumes a multivariate Gaussian
parameter estimate distribution and power spectrum distribution.
However, it now contains the non-linear information found in the
off-diagonal components of the covariance matrix from the simulations,
giving us a `non-linear' Fisher matrix. In this paper, we will show
the effect of this simple modification to the calculation of the
Fisher matrix on the error estimates and compare this with the maximum
likelihood and Gaussian Fisher error contours in the
$\Omega_m-\sigma_8$ plane. We also perform a multi-parameter analysis
of $\Omega_m,~ \sigma_8, ~ h, ~n_s, ~w_0$ and $w_a$ and compare the
Gaussian and non-linear Fisher matrix error estimates.
It is possible to produce data vectors for covariance matrices by
performing a simple 2-D binning of the galaxies in the
survey. However, further information may be gained by splitting the
distribution up in to redshift bins and performing a tomographic
analysis \citep[e.g.][]{h99,h02b,jt03}. In this work, we perform both
2-D and 3-bin tomographic analyses in order to generate covariance
matrices that are used to calculate the maximum likelihood estimates
and the Gaussian and non-linear Fisher matrices.
The outline of this paper is as follows. Section \ref{sec:sunglass}
will detail how the simulations and mock galaxy catalogues were
generated. Section \ref{sec:fisher} will introduce the Fisher matrix
formalism and the maximum likelihood formalism. Section
\ref{sec:results} shows the results of the analyses on the maximum
likelihood estimates and the Gaussian and non-linear Fisher matrices
from the 2-D analyses in Section \ref{sec:unbinned} and the tomographic
analyses in Section \ref{sec:tomo}. Section \ref{sec:multi} compares
the multi-parameter Gaussian and non-linear Fisher estimates.
Finally, a summary of the findings will be presented in Section
\ref{sec:disc}.
\section{Details of the simulations}
\label{sec:sunglass}
\begin{table}
\begin{center}
\begin{tabular}{cccccc}
\hline\\
$N$ & Area (sq deg) & $z_{\rm max}$ & $n_g$/sq arcmin
& $z_{\rm med}$ \\
\hline\\
100 & 100 & 1.5 & 15 & 0.82 \\
\hline
\end{tabular}
\end{center}
\caption{Table of parameters for the mock galaxy shear catalogues
used in this paper. $N$ is the number of independent lightcones,
$z_{\rm max}$ is the maximum redshift in the lightcone and
$n_{g}$/sq arcmin is the number of `galaxies' per square arcmin in
the catalogue and $z_{\rm med}$ is the median redshift of the
catalogue. The suite of lightcones is used together to form a
survey with an effective area of 10,000 sq. deg.}
\label{tab:lightconetable}
\end{table}
The suite of weak lensing simulations used in this work was generated
using the {\small SUNGLASS} pipeline \citep[for a detailed
introduction, see][]{kht+11}. We have 100 independent simulations
generated with the cosmological N-body simulations code {\small
GADGET2} \citep{sp05}. The simulations were made with a flat
concordance $\Lambda$CDM cosmology, consistent with the WMAP 7-year
results \citep{jbd+11}: $\Omega_m = 0.272$, $\Omega_{\Lambda} =
0.728$, $\Omega_b = 0.045$, $\sigma_8 = 0.809$, $n_s = 0.963$ and
$h=0.71$ in units of 100 km s$^{-1}$ Mpc$^{-1}$. There are $512^3$
particles in a box of $512h^{-1}$~Mpc which leads to a particle mass
of $7.5 \times 10^{10}\ensuremath{ {\rm M}_{\odot} }$. The simulations were all started from a
redshift of $z=60$ and allowed to evolve to the present with 26
snapshots being stored in redshifts $0.0 \le z \le 1.5$.
Lightcones were generated through the simulation snapshots to
determine the average convergence in an angular pixel using the `no
radial binning' method introduced in \cite{kht+11}:
\begin{equation}
\kappa_p(r_s) = \sum_k \frac{K(r_k,r_s)}{\Delta\Omega_p \bar{n}
(r_k)r_k^2} - \int_0^{r_s}\!dr~ K(r,r_s),
\end{equation}
where $\bar{n}$ is the number density of particles in the simulation,
$\Delta\Omega_p = \Delta\theta_x\Delta\theta_y$ is the pixel area,
$r_s$ is the comoving radial distance of the lensing source plane and
$r_k$ is the comoving radial distance of each particle, $k$, in the
lightcone. $K(r,r_s)$ is the scaled lensing kernel:
\begin{equation}
K(r,r_s) = \frac{3 H_0^2 \Omega_m}{2c^2} \frac{(r_s - r)r}{r_sa(r)}.
\end{equation}
The convergences were calculated using $2048^2$ azimuthal bins on
source redshift planes that were separated by $z = 0.1$ to create 15
planes from $0.0 < z < 1.5$. Once the convergences were calculated,
shear values were determined in Fourier space, where the shear is
given by $\gamma = \gamma_1 + i\gamma_2$.
The shear and convergence values in these source redshift planes were
interpolated back onto the individual particles in the lightcones to
generate mock galaxy shear catalogues. The B-modes (the unphysical,
imaginary component of the convergence) in the mock catalogues were
calculated directly from the shear. These catalogues were then sampled
to reproduce a standard galaxy redshift distribution
\citep[e.g.][]{ebt+91},
\begin{equation}
n(z) \propto z^\alpha \exp \left[ - \left(\frac{z}{z_0}\right)^\eta
\right],
\end{equation}
where $z$ is the redshift, $z_0$, $\alpha$ and $\eta$ set the depth,
low-redshift slope and high-redshift cut-off for a given galaxy
survey. We take $\alpha=\eta=2$ and $z_0=0.78$ in this work, which
gives a median redshift $z_{med} = 0.82$. It is assumed that galaxies
trace the dark matter distribution perfectly and the final mock galaxy
shear catalogues contain 15 galaxies per square arcminute. There is
no ellipticity noise in the catalogues, however there is a shot-noise
contribution related to the discrete sampling of the particles in the
mock catalogues. Table \ref{tab:lightconetable} summarises the mock
galaxy catalogues used in this work.
\section{Methodology}
\label{sec:fisher}
The shear, $\gamma$, convergence, $\kappa$, and B-mode, $\beta$,
fields are related to each other in Fourier-space on a flat-sky by
\begin{equation}
\kappa(\mbox{\boldmath $\ell$}) + i\beta(\mbox{\boldmath $\ell$}) = e^{2 i \varphi_\ell} [\gamma_1(\mbox{\boldmath $\ell$}) +
i\gamma_2(\mbox{\boldmath $\ell$}) ],
\end{equation}
where $\varphi_\ell$ is the angle between the angular wave-vector and
an axis on the sky. For each of the mock galaxy shear catalogues the
shear, convergence, and B-mode auto- and cross-power spectra have been
estimated. The tomographic weak lensing shear, convergence and B-mode
power cross-spectra, for two different source redshifts $z$ and $z'$,
are given by
\begin{eqnarray}
C_\ell^{\gamma\gamma} (z,z') &=& \langle \gamma_1(\boldsymbol{\ell},z)
\gamma_1(\mbox{\boldmath $\ell$},z') \rangle + \langle \gamma_2(\boldsymbol{\ell},z)
\gamma_2(\mbox{\boldmath $\ell$},z') \rangle,\\
C_\ell^{\kappa\kappa}(z,z') &=& \langle
\kappa(\boldsymbol{\ell},z)\kappa(\mbox{\boldmath $\ell$},z') \rangle,\\
C_\ell^{\beta\beta}(z,z') &=& \langle \beta(\boldsymbol{\ell},z)
\beta(\mbox{\boldmath $\ell$},z') \rangle.
\end{eqnarray}
The power spectra are related to each other by $C_\ell^{\gamma\gamma}
(z,z')= C_\ell^{\kappa\kappa} (z,z')+ C_\ell^{\beta\beta}(z,z')$. The
auto-spectra are calculated when $z'=z$. In practice we bin the
sources into redshift slices. For a survey that has $N_z$ slices the
expectation-value of the tomographic shear power spectrum in redshift
bins labeled $i$ and $j$, is given by
\begin{equation}
C^{\gamma\gamma}_{ij}(\ell) = \frac{9H_0^4\Omega_m^2}{4c^4}
\int^{r_{\rm max}}_0 \! \frac{dr}{a^2(r) }~ P\left(\frac{\ell}{r} ,
r\right) g_{i}(r) g_{j}(r),
\end{equation}
\citep{k92,js09}, where $r$ is the comoving distance, $r_{\rm max}$ is
the maximum comoving distance and $P(\ell/r,r)$ is the 3D matter
density power spectrum. The lensing efficiency function is
\begin{equation}
g_{i}(r) = \int_r^{r_{\rm max}} \!dr~ p_{i} (r') \left( 1 -
\frac{r}{r'} \right),
\end{equation}
where $p_{i}$ is the normalised probability distribution of `galaxies'
in the bin.
When determining the angle averaged shear power spectrum in the
simulations, we must take into consideration the conventions used in
the Fourier transform software FFTW\footnote{The Fastest Fourier
Transform in the West http://www.fftw.org}. Thus, the discretised
tomographic shear power spectrum becomes
\begin{equation}
\frac{\ell (\ell + 1) \widehat{C}_{ij}^{\gamma\gamma}(\ell)} {2\pi} =
\!\!\! \sum_{\ell \, in \, shell} \!\!\!\! \frac{\gamma_1(\mbox{\boldmath $\ell$},z_i)
\gamma_1(\mbox{\boldmath $\ell$},z_j) + \gamma_2(\mbox{\boldmath $\ell$},z_i)\gamma_2(\mbox{\boldmath $\ell$},z_j)}{n_b^2
\Delta \ln \ell},
\end{equation}
where $\widehat{C}_\ell^{\gamma\gamma}$ is the estimated power, $n_b$
is the total number of bins in the Fourier transform, $z_i$ is the
$i^{th}$ redshift slice, and $\Delta \ln \ell$ is the thickness of a
shell in log $\ell$-space. The modes in this power spectrum are
corrected for mode discreteness errors by scaling by the expected
number of modes. To compactify the notation we shall denote the
tomographic shear power by $C^{\gamma\gamma}_X(\ell)$, where $X=(i,j)$
is a pair of redshift slices. For our 2-D analysis there is only one
bin so $i=j$. Using the information from these power spectra, we have
the the information required to make cosmological parameter estimates.
\subsection{Shear power covariance matrix generation}
The shear cross-spectra covariance matrix, $M^{XX'}_{\ell \ell'}$ is
defined by
\begin{equation}
M^{XX'}_{\ell\ell'} = \langle \Delta C^{\gamma\gamma}_X(\ell) \Delta
C^{\gamma\gamma}_{X'} (\ell') \rangle,
\end{equation}
where $\Delta C^{\gamma\gamma}_X(\ell)=C^{\gamma\gamma}_X(\ell)-\langle
C^{\gamma\gamma}_X(\ell)\rangle$ and angled-brackets denotes ensemble
averaging. The covariance matrix is an important element of parameter
estimation, containing information on the strength of the correlations
between variates, in this case the shear power spectrum modes. The
accuracy of this matrix improves by increasing the numbers of
realisations included in the calculation of the matrix. For this
work, 100 independent realisations of 2-D power spectra provides an
accuracy of $\Delta C_\ell \sim 10\%$ \citep[as shown in Figure 7
of][]{kht+11}.
\subsubsection{Gaussian covariance matrix}
If the shear field is isotropic and we assume the shear field is
Gaussian we can write down an equation for the tomographic shear power
covariance matrix and make a correction for the fraction of the sky
covered by the survey:
\begin{eqnarray}
M^{ij,kl}_{\ell\ell'} \!\!\! \!\! & =& \!\!\!\!\!\!
\frac{\delta^K_{\ell\ell'}} {(2\ell+1) f_{\mathrm{sky}}} \Big(
[C_{ik}^{\gamma\gamma} (\ell)+N_i(\ell) \delta^K_{ik}]
[C_{jl}^{\gamma\gamma}(\ell)+N_j(\ell) \delta^K_{jl}] \nonumber
\\
& & \,\,\, + \,\,\, [C_{il}^{\gamma\gamma}(\ell)+N_i(\ell)
\delta^K_{il}][C_{jk}^{\gamma\gamma}(\ell)+N_j(\ell) \delta^K_{jk}]
\Big) ,
\label{eq:theoryCoV}
\end{eqnarray}
where $f_{\mathrm{sky}}$ is the fraction of the sky covered by the
survey and $N_\ell$ is a shot-noise term due to intrinsic ellipticity
in the shear field. For the purposes of this work although we have
discrete galaxy density, we set the intrinsic shear variance to zero
so that $N_\ell=0$. In principle our results can be combined with a
Gaussian noise covariance to model different weak lensing surveys.
\subsubsection{Simulation covariance matrix estimation}
The covariance matrix of the tomographic shear power spectrum for the
suite of mock galaxy shear catalogues is estimated by
\begin{equation}
M^{XX'}_{\ell\ell'} = \frac{1} {N-1} \sum_N \Delta
C_{X}^{\gamma\gamma}(\ell) \Delta C_{X'}^{\gamma\gamma} (\ell'),
\label{eq:CoV}
\end{equation}
where $N$ is the number of mock catalogue realisations and here
$\Delta C_{X}^{\gamma\gamma} (\ell) = \widehat{C}_{X}^{\gamma\gamma}
(\ell) - \langle \widehat{C}_{X}^{\gamma\gamma}(\ell) \rangle$ where $\langle
\widehat{C}_{X}^{\gamma\gamma}(\ell) \rangle$ is the ensemble average of
the tomographic shear power spectrum across all realisations.
We calculate inverse covariance matrix using a singular value
decomposition (SVD) on the covariance matrix \citep{ptv+92}. However,
the resulting inverse is biased due to noise. To correct for this, we
multiply the inverse by a factor \citep{hss07}:
\begin{equation}
[\widehat{M}^{XX'}_{\ell\ell'}]^{-1} = \frac{N_S - N_p - 2} {N_S - 1}
[M^{XX'}_{\ell\ell'}]^{-1},
\end{equation}
where $N_S$ is the number of realisations, $N_p$ is the total number
of bins in all of the power spectra and
$[\widehat{M}^{XX'}_{\ell\ell'}]^{-1}$ is an unbiased estimation of
the inverse covariance matrix.
\subsection{Fisher matrix generation}
The Fisher matrix is given by
\begin{equation}
F_{ij} = \frac{1}{2} \sum_{\ell \ell'} \sum_{XX'} \frac {\partial
C^{\gamma\gamma}_X(\ell)} {\partial \theta_i} \left[
\widehat{M}^{XX'}_{\ell\ell'}\right]^{-1} ~ \frac {\partial
C^{\gamma\gamma}_{X'}(\ell')} {\partial \theta_j} ,
\label{eq:Fisher}
\end{equation}
where $i$ and $j$ label cosmological parameters (e.g. $\Omega_m$ and
$\sigma_8$) and the partial derivatives are the gradient of the
expectation value of the shear power spectrum in parameter space. In
order to determine that gradients of the power spectra, we use a
five-point function \citep[][equation 25.3.6]{as68} and the ensemble
average power spectra. The theoretical prediction for the power
spectrum was generated using a code provided by Benjamin Joachimi
\citep[as tested in][]{kht+11,js09}. This code uses the \cite{spj+03}
non-linear power spectrum estimate, the \cite{eh98} matter transfer
function and a numerically calculated linear growth factor. All
further theoretical power spectrum predictions in this paper were
performed using this same code.
The Fisher matrix provides parameter estimates through two methods.
The first by calculating the area of the Fisher matrix error ellipse
that encloses a two-parameter 68\% confidence limit in the two
parameter plane. The inverse of this area is proportional to the
Figure-of-Merit that is often quoted in studies
\citep[e.g.][]{abc+06,w08,aab+09,sl10}. The second is the
single-parameter marginal errors, which are given by
\begin{equation}
\Delta \theta_i = \left[F^{-1}\right]_{ii}^{1/2}.
\end{equation}
\subsection{Maximum likelihood parameter estimation}
To perform a maximum likelihood analysis on the suite of mock shear
catalogues, we use a Gaussian likelihood estimator. Despite the fact
that the simulations are non-Gaussian, using a Gaussian likelihood has
been shown to produce accurate results \citep{kht+11}. The likelihood
is given by
\begin{equation}
L(\hat{C}_\ell^{\gamma\gamma}|\sigma_8,\Omega_m) = \frac{1}
{(2\pi)^{N/2} ({\rm det} \widehat{M}^{XX'}_{\ell \ell'})^{1/2}}
\exp\left[- \frac{\chi^2} {2} \right],
\label{eq:Like}
\end{equation}
where $N$ is the number of independent mock catalogue realisations,
and
\begin{equation}
\chi^2 = \sum_{\ell \ell'} \sum_{XX'} \Delta C_{X}^{\gamma\gamma}
(\ell) \left[\widehat{M}^{XX'}_{\ell\ell'}\right]^{-1} \Delta
C_{X'}^{\gamma\gamma}(\ell'),
\end{equation}
where $\Delta C_{X}^{\gamma\gamma} (\ell) =
\widehat{C}_X^{\gamma\gamma}(\ell) - \langle C_{X}^{\gamma\gamma}
(\ell) \rangle$, and $\langle C^{\gamma\gamma}_\ell \rangle$ is the
expected angular power spectrum given by equation (8). While this
likelihood analysis is computationally expensive, it takes into
account the full non-Gaussian and non-linear nature of the simulations
and should provide the most accurate error estimates. The maximum
likelihood is related to the Fisher matrix by
\begin{equation}
F_{ij} = \left \langle - \frac{\partial^2\ln L}{\partial \theta_i
\partial \theta_j}\right\rangle.
\end{equation}
\section{Results}
\label{sec:results}
\subsection{2-D two-parameter analysis}
\label{sec:unbinned}
\begin{figure}
\psfig{file=nz_PSwErrors_gamma.ps,angle=-90,width=\linewidth,clip=}
\caption{Mean 2-D shear power spectrum for the suite of unbinned mock
galaxy shear catalogues.The smooth (red) line is the theoretical
prediction for the power spectrum, the (black) line shows the mean
power spectrum for the suite of mock catalogues with errors on the
mean shown. The diagonal (dark blue) dashed line shows the shot
noise prediction and the (magenta) triangles show the measured
B-modes. The bottom panel shows the the percentage difference of
the data from the expected power spectrum with errors.}
\label{fig:nzPS}
\end{figure}
The first step to making parameter estimates from the mock galaxy
shear catalogues is to determine the 2-D shear angular power spectra
for each realisation. Figure \ref{fig:nzPS} shows the mean power
spectra from the suite with the (red) long line showing the
theoretical prediction, the (black) line showing the mean measured
power spectrum from the suite of catalogues with errors on the mean.
The (magenta) triangles show the measured B-modes and the dashed (dark
blue) line shows the shot-noise arising from discrete particle
sampling in the mock catalogues. The bottom panel shows the
percentage difference of the data from the expected power spectrum.
The simulations recover the expected power spectrum within 5\% between
$150 < \ell < 2000$.
The shot noise prediction for these catalogues was determined by
filling a suite of simulation volumes with randomly placed particles
to mimic Poisson noise. The {\small SUNGLASS} pipeline was run on
these noise simulations and the power spectrum of the mock catalogues
is the dashed line in the figure. We know that the shot-noise in these
catalogues is not purely Poissonian because the simulations start from
a glass pre-initial distribution \citep{w94}. However, the noise tends
toward Poissonian as the simulations evolve. We are not able to model
this noise accurately because it evolves with structures in the
simulation \citep{bge95}. Thus, this Poisson estimate should be a
reasonable approximation for the noise in these catalogues.
A data-vector was constructed from the shear band-power spectra with
wavenumbers from $150 < \ell < 1500$. From this, we estimated the
covariance matrix in equation (\ref{eq:CoV}). Similarly, a Gaussian
covariance matrix was generated using equation (\ref{eq:theoryCoV}).
To test the effect of the non-Gaussian nature of the simulations
covariance matrix, we looked at the diagonal components of both the
simulation and Gaussian field covariance matrices and determined their
$\Delta C_{\ell}$ values. For the case of the simulations this is
simply the square root of the diagonal components. For the Gaussian
field this is
\begin{equation}
\Delta C_\ell = \frac{\sqrt{2}C_\ell}{\sqrt{\ell(2\ell+1) f_{\rm sky} \Delta
\ln \ell}}.
\end{equation}
\begin{figure}
\centerline{\psfig{file=CoV_Error_Check.ps,angle=-90,width=\linewidth,clip=}}
\caption{Diagonal components of the covariance matrix for the
simulations (black line) and a Gaussian field with the same power
spectrum (red dashed line).}
\label{fig:CoVError}
\end{figure}
Figure \ref{fig:CoVError} shows the diagonals of the covariance
matrices as a function of wavenumber $\ell$. The (red) dashed line is
from the Gaussian error and the (black) line is from the
simulations. At low wavenumbers, which is still in the reasonably
linear regime, the two errors agree reasonably well. However, in the
higher wavenumber, non-linear regime, there is a factor of three
difference between the errors, suggesting that the non-Gaussian
contribution to the covariance is significant.
Fisher matrices in the $\Omega_m - \sigma_8$ plane were calculated
using equation (\ref{eq:Fisher}) for both the Gaussian and simulation
covariance matrices. The Fisher matrices were multiplied by 100, the
total number of independent mock catalogues, to provide an error
estimate for a survey of 10,000 square degrees. Additionally, the
maximum likelihood was calculated using the simulation covariance
matrix and the likelihoods from the 100 mock catalogues were combined
\citep[as shown in ][]{kht+11}. For all of these calculations it is
assumed that all other cosmological parameters are known.
In Figure \ref{fig:LikeFish} the two-parameter $1\sigma$ contours for
the $\Omega_m - \sigma_8$ plane are shown. The inner (orange shaded)
ellipse shows the Gaussian Fisher estimate, the outer (grey shaded)
ellipse shows the non-linear Fisher estimate and the thick (black)
line contour is the combined maximum likelihood estimate. The red
point in the middle represents the fiducial $\Omega_m$ and $\sigma_8$
parameters. The size and shape of the simulation Fisher ellipse and
the maximum likelihood contour are very similar, showing that the
non-linear Fisher calculation is a good method for making future
cosmological parameter estimates, provided the off-diagonal terms are
included. The area of the non-linear Fisher estimate is 5.1 times
larger than the area of the Gaussian Fisher estimate which is a clear
indication that the off-diagonal terms in the covariance matrix are
essential for accurate estimation of the parameter errors.
\begin{figure}
\centerline{\psfig{file=Likelihood_Fisher.ps,angle=-90,width=\linewidth, clip=}}
\caption{Comparison of the $1\sigma$ two-parameter contours for the
$\Omega_m-\sigma_8$ parameters for Gaussian Fisher contour (inner
orange shaded ellipse) with the non-linear Fisher contour (middle
grey shaded ellipse) and the full maximum likelihood analysis
(thick black line contour). The (red) point in the middle of the
ellipses represents the fiducial $\Omega_m-\sigma_8$ parameters in
this calculation.}
\label{fig:LikeFish}
\end{figure}
In addition to the area of the contours we also determined the
marginal errors on the parameters from the Fisher matrices. The
non-linear $\Omega_m$ marginal error is 1.23 times larger than the
Gaussian marginal errors and $\sigma_8$ is 1.17 times larger. From
Figure \ref{fig:CoVError}, we would expect the marginal errors to be
much larger. However, the off-diagonal terms in the simulation
covariance matrix act to reduce the marginal errors but increase the
total area. This is demonstrated by setting the off-diagonal terms in
the simulation covariance matrix are to zero. In this case the
non-linear Fisher ellipse becomes far narrower but the marginal errors
increase significantly.
Figure \ref{fig:Marginal} shows how the marginal errors change in the
non-linear Fisher matrix as a function of maximum wavenumber
$\ell_{\rm max}$. The marginal errors for both the non-linear and
Gaussian Fisher matrices are remarkably similar across all values of
$\ell_{\rm max}$ with the largest gain of information occurring for
both $\Omega_m$ and $\sigma_8$ between $250 < \ell_{\rm max} < 500$
and again at $\ell_{\rm max} > 1000$. This similarity is due to an
effect that the off-diagonal components of the simulation covariance
matrix are having on the error ellipse, as discussed earlier. This
figure shows that if the marginal error is the value of interest, the
Gaussian Fisher matrix appears to provide a result that is comparable
with the error obtained with the non-linear Fisher error estimate at
wavenumbers between $150 < \ell_{\rm max} < 1500$. However, when the
errors are marginalised over two parameters, the area of the contours
shows that the errors are being underestimated by a factor of five.
\begin{figure}
\centerline{\psfig{file=Fisher_MarginalError.ps,angle=-90,width=\linewidth,clip=}}
\caption{Marginal errors of $\Omega_m$ and $\sigma_8$ as a function
of maximum wavenumber in the Fisher calculation. The dashed lines
are the marginal errors on $\Omega_m$ as a function of maximum
wavenumber $\ell_{\rm max}$ and the continuous lines are for
$\sigma_8$.}
\label{fig:Marginal}
\end{figure}
\subsection{Tomographic shear power}
\label{sec:tomo}
It is possible to perform the same analysis that was performed in
Section \ref{sec:unbinned} on tomographically-binned mock catalogues.
Tomography introduces extra information into the analysis. In this
analysis the suite of mock galaxy shear catalogues was split into
three redshift bins; bin 1: $0.0 \le z \le 0.5$, bin 2: $0.5 < z \le
1.0$ and bin 3: $1.0 < z \le 1.5$.
Figure \ref{fig:PSComp} shows the auto- and cross-power spectra for
the 3-bin tomographic analysis of the mock galaxy shear catalogues.
If we first focus on the auto-power spectra ($C_{11},~ C_{22}$ and
$C_{33}$), we can see that the wavenumbers that we model accurately
increase with redshift. For $C_{11}$, we find agreement for
wavenumbers from $150 < \ell < 700$. The power spectra from $C_{22}$
and $C_{33}$ are slightly higher than the expected power spectrum but
still within 3\% and are reliable up to $\ell = 1000$ before
shot-noise becomes dominant.
The cross-bins $C_{12}, ~C_{13}$ and $C_{23}$ are damped which is due
to the shot noise dominance in the lower redshift bin. Consequently,
we only use the results of these power spectra up to the $\ell$ range
accurately recovered by the auto-power spectra of the lower redshift
bin.
The power spectra in the auto- and cross-bins were turned into a
data-vector with wavenumbers from $150 < \ell < 1800$ included. A
data covariance matrix was generated and from this, the correlation
coefficient matrix where the correlation coefficients are given by
\begin{equation}
r^{XX'}_{\ell\ell'} = \frac{\widehat{M}^{XX'}_{\ell\ell'}}
{\sqrt{\widehat{M}^{XX}_{\ell\ell} \widehat{M}^{X'X'}_{\ell'\ell'}}}.
\end{equation}
This matrix shows how (anti-)correlated each of the $\ell$ modes
are.
Figure \ref{fig:CCM_Tomo} shows the correlation coefficient matrix for
the tomographic data vector. The `block' nature to the matrix
indicates each tomographic bin pair. Each of the blocks along the
diagonal represents the auto-correlation between each of the
tomographic power spectrum analyses. The off-diagonal blocks show the
cross-correlations between the tomographic pair power spectra. As
expected, the higher wavenumbers in each block are highly correlated.
Additionally, the $C_{11}$ auto-power spectrum is highly correlated.
The $C_{22}$ auto-power spectrum is significantly less correlated and
the $C_{33}$ auto-power spectrum bin has very low correlations between
the $\ell$ modes. The indication of this is that the lower redshift
tomographic bins should contribute fewer power spectrum bins to a
covariance matrix being used for any kind of analysis.
\begin{figure*}
\begin{minipage}{148mm}
\begin{tabular}{cc}
\psfig{file=Tomo_PS_1.ps,angle=-90,width=0.475\linewidth,clip=} &
\psfig{file=Tomo_PS_2.ps,angle=-90,width=0.475\linewidth,clip=}\\
\psfig{file=Tomo_PS_3.ps,angle=-90,width=0.475\linewidth,clip=} &
\psfig{file=Tomo_PS_4.ps,angle=-90,width=0.475\linewidth,clip=}\\
\psfig{file=Tomo_PS_5.ps,angle=-90,width=0.475\linewidth,clip=} &
\psfig{file=Tomo_PS_6.ps,angle=-90,width=0.475\linewidth,clip=}
\end{tabular}
\end{minipage}
\caption{2-D shear power spectra for a 3-bin tomography analysis of the
mock galaxy shear catalogues. The long (red) line is the theoretical
prediction for the shear power spectrum, the (black) line is the
mean power spectrum for the 100 mock catalogues with errors on the
mean, the (dark blue) dashed line is the shot noise estimate and the
(magenta) triangles are the measured B-modes. }
\label{fig:PSComp}
\end{figure*}
\begin{figure*}
\centering
\psfig{file=CCM_Tomo_test.ps,angle=-0,width=\linewidth,clip=}
\caption{Correlation coefficient matrix for 3 tomographic bins. The
matrix is arranged so that each sub-square is a tomographic bin
pair.}
\label{fig:CCM_Tomo}
\end{figure*}
With the information gained from these power spectra and the
correlation coefficient matrix, we are able to assemble a data-vector
that contains the $C_\ell$ from wavenumbers that are accurately
reproducing the expected power spectrum and have a reasonably low
correlation. Thus, we select 30 bins in total from the three
auto-power spectra, with the highest wavenumber being around $\ell =
1000$ and just 10 bins in total from the cross-power spectra, with the
highest wavenumber being around $\ell = 500$. A covariance matrix was
generated using this data-vector. A full maximum likelihood analysis
was performed with this covariance matrix and a non-linear Fisher
matrix was also calculated. An equivalent Gaussian covariance matrix
was also generated and the Gaussian Fisher matrix determined.
\subsection{Two-parameter tomographic analysis}
Figure \ref{fig:LikeFish_Tomo} shows the $1\sigma$, two-parameter
contours for the Gaussian Fisher (orange shaded ellipse) and the
non-linear Fisher (grey shaded ellipse). The thick black line shows
the simulation maximum likelihood analysis. As in the case for the
2-D analysis, the non-linear Fisher contour is very close to the
maximum likelihood contour. The area of the non-linear Fisher contour
is 5.1 times larger than the area of the Gaussian Fisher contour. The
marginal error of the non-linear Fisher estimate is 1.39 times larger
in $\Omega_m$ and 1.24 times larger in $\sigma_8$. These numbers are
very similar to those calculated in the 2-D analysis.
The area of the contours and the size of the marginal errors in the
tomographic Fisher and likelihood estimates is smaller than those in
the unbinned analyses, showing that the tomographic analysis does
indeed contain more information.
\begin{figure}
\centerline{\psfig{file=Likelihood_Fisher_Tomo.ps,angle=-90,width=\linewidth, clip=}}
\caption{Comparison of the $1\sigma$, two-parameter contours for the
$\Omega_m-\sigma_8$ parameters for the Gaussian Fisher contour
(inner orange shaded ellipse), the simulation Fisher contour
(middle grey shaded ellipse) and the full simulation maximum
likelihood analysis (thick black line contour). The (red) point in
the middle of the contours represents the fiducial
$\Omega_m-\sigma_8$ parameters in this calculation.}
\label{fig:LikeFish_Tomo}
\end{figure}
\subsection{Multi-parameter Fisher analysis}
\label{sec:multi}
\begin{table}
\begin{center}
\begin{tabular}{l|ccc}
\multicolumn{4}{c}{\textbf{Tomographic Multi-Parameter Analysis}}\\
\hline
& NL & Gauss & Gauss / NL \\
\hline\\
$\Delta \Omega_m$ & 0.009 & 0.010 & 1.11 \\
$\Delta \sigma_8$ & 0.019 & 0.021 & 1.11 \\
$\Delta h$ & 0.198 & 0.352 & 1.78 \\
$\Delta n_s$ & 0.143 & 0.201 & 1.41 \\
$\Delta w_0$ & 0.107 & 0.129 & 1.21 \\
$\Delta w_a$ & 0.343 & 0.584 & 1.70 \\
\hline
\end{tabular}
\end{center}
\caption{Marginal errors of a 3-bin tomographic, six-parameter
Fisher matrix analysis using both Gaussian and non-linear Fisher
matrices. For this configuration, the Gaussian marginal errors are
always larger than the non-linear Fisher errors. However, the
non-linear Fisher volume is 3.7 times larger than the Gaussian
volume.}
\label{tab:marginal}
\end{table}
In the previous section, we showed that a Fisher matrix analysis in
the $\Omega_m - \sigma_8$ plane accurately determines the errors on
the parameters when a non-Gaussian data covariance matrix is used in
the calculation. This was shown by comparing the 68\% error contour
with a maximum likelihood analysis of the simulation suite. In this
Section we perform a Fisher matrix analysis over the cosmological
parameters: $\Omega_m, ~\sigma_8, ~h, ~n_s, ~w_0$ and $w_a$. Note that
$\Omega_\Lambda = 1 - \Omega_m$. We compare the Gaussian Fisher
analysis with the non-Gaussian (simulation) Fisher analysis. Based on
the findings in the two-parameter analysis, we assume that the Fisher
errors calculated with the simulation data covariance matrix are
consistent with a maximum likelihood analysis of the simulations over
the same multi-parameter space.
For a Euclid-like survey (Table \ref{tab:lightconetable}), it makes
little sense to perform a 2-D analysis over multiple parameters. We
found that the size of the marginal errors was so large as to provide
no constraining information (e.g. $\Delta w_0 = 11.3$ and $\Delta w_a
= 49.8$). However, the additional information provided by performing
a tomographic analysis of the power spectra yielded far better
constraints and illuminated some features of the Fisher matrix
analysis that were not immediately obvious. Figure
\ref{fig:Tomo_Multi} shows the projected two-parameter, $1\sigma$
contours, marginalised over the the multiple parameters from the
Fisher matrix analysis. The (red) point at the center shows the
fiducial parameters used in the analysis, the (grey) shaded ellipses
show the non-linear Fisher analysis and the dashed (orange) ellipses
show the Gaussian Fisher analysis. In all cases, except the
$\Omega_m-\sigma_8$ plane, the projected area of the Gaussian contours
is larger than the area of the non-linear Fisher contours which gives
the impression that the error estimates from the Gaussian Fisher
analysis are more conservative the non-linear estimates. The marginal
errors of the Gaussian contours are larger in all cases (Table
\ref{tab:marginal}). However, volume of the six-parameter space
generated,
\begin{equation}
V \propto \sqrt{\det(F^{-1})},
\end{equation}
shows that the non-linear Fisher volume is 3.7 times larger than the
Gaussian volume, which is what we would expect given the off-diagonal
terms included in the simulation data covariance matrix. The
implication of this is that although the projected areas of the
Gaussian Fisher appear larger, the overall volume is smaller. This can
be explained with an example (in three dimensions for simplicity of
explanation): Take a spherical ball and a thin plate with a slightly
larger radius than the ball. If we look at the ball from any of the
three axes, the area will appear to be the same circle in each
projection. If we place the plate at a $45^\circ$ angle to each of
the three axes and then look at it in projection, it will appear to
have a larger area than the ball in all projections, however we know
that the volume of the plate is far smaller than the ball. With this
knowledge in hand, it is easy to see how misleading the projections of
the Fisher matrices can be in these complex multiple-parameter spaces.
\begin{figure*}
\centerline{\psfig{file=Fisher_Tomo_Multi.ps,angle=-90,width=\linewidth,clip=}}
\caption{Multiple parameter Fisher analysis. The (grey) shaded
ellipse is the non-linear Fisher contour and the (orange) dashed
ellipse is the Fisher contour for the Gaussian Fisher matrix. The
volume of the 6-dimensional space is larger in the simulation
Fisher matrix, however some of the projected contours appear
larger for the Gaussian Fisher matrix. This could be very
misleading.}
\label{fig:Tomo_Multi}
\end{figure*}
The eigenvalues and eigenvectors of the Fisher matrices were
determined to find the equation for the plane with the most
information in the full six-parameter space. For the Gaussian, this
is:
\begin{eqnarray}
X &=& 0.865~\Omega_m + 0.498~\sigma_8 + 0.049~n_s + 0.030~h -
\nonumber\\ && 0.035~w_0 - 0.006~w_a,
\end{eqnarray}
and for the simulations:
\begin{eqnarray}
Y &=& 0.841~\Omega_m + 0.536~\sigma_8 + 0.034~n_s + 0.028~h -
\nonumber \\ & & 0.056~w_0 - 0.015~w_a.
\end{eqnarray}
When using the fiducial parameters,
\begin{eqnarray}
X & = & 0.741, \\
Y & = & 0.771,
\end{eqnarray}
and the errors on the planes are $\Delta X = 8.6 \times 10^{-8}$ and
$\Delta Y = 1.4 \times 10^{-6}$. This shows an order-of-magnitude
difference in the thickness of the 6-D error ellipses.
\section{Discussion and Conclusions}
\label{sec:disc}
This paper makes a comparison between Fisher matrices in the $\Omega_m
- \sigma_8$ plane and six-parameter Fisher matrices ($\Omega_m,~
\sigma_8, ~h, ~n_s, ~w_0$ and $w_a$), generated using covariance
matrices from a Gaussian random field and from full N-body
simulations. The $1\sigma$ two-parameter contours from the
two-parameter $\Omega_m-\sigma_8$ Fisher matrices are also compared
with the contour from a maximum likelihood analysis of the full suite
of simulations.
This work uses the {\small SUNGLASS} pipeline to generate 100
independent simulations of 512$h^{-1}$~Mpc with $512^3$ particles and
a standard $\Lambda$CDM cosmology. The pipeline turns these
simulations into 100 independent mock galaxy shear catalogues of 100
square degrees and a galaxy redshift distribution with a median of
$z_m = 0.82$ and 15 `galaxies' per square arcminute. When these
catalogues are combined, they provide an effective survey area of
10,000 sq. deg.
We perform both 2-D and 3-bin tomographic angular shear power spectrum
analyses on each of the mock catalogues and generate covariance
matrices from the resulting data-vectors. We also generate a Gaussian
field covariance matrix to compare with the more realistic
non-Gaussian analyses.
Using these covariance matrices, we generate Fisher matrices for the
$\Omega_m - \sigma_8$ plane and show that the $1\sigma$, two-parameter
contour for the non-linear Fisher matrix has an area that is 5.1 times
larger than the theoretical prediction in both the 2-D and tomographic
analyses. This indicates that the theoretical prediction is
significantly under-predicting the errors on these parameters (the
Figure-of-Merit is over-optimistic), even though their marginal errors
are similar.
To quantify if the contours generated using the Fisher matrix are a
reasonable estimate of the true errors, we compared these outputs with
a combined maximum likelihood analysis of the full simulation suite.
The resulting contours in both the 2-D and tomographic analyses
closely matched the contours generated with the simulation Fisher
matrix. From this, we can conclude that it is sensible to use a
Fisher matrix analysis for parameter estimates, using a covariance
matrix with all off-diagonal terms included.
Based on the success of the non-linear Fisher errors in matching the
maximum likelihood estimates in the $\Omega_m-\sigma_8$ plane, we also
perform both 2-D and 3-bin tomographic analyses to generate
multi-parameter Fisher matrices of $\Omega_m, ~\sigma_8, ~h, ~n_s,
~w_0$ and $w_a$. We compare the non-linear Fisher matrix with the
Gaussian Fisher matrix under the assumption that the simulation Fisher
matrices are providing accurate error estimates. With the survey
parameters used in this paper, the 2-D analysis finds marginal errors
on the parameters that are so large that they provide no constraining
power.
The tomographic analyses find reasonable marginal errors and
demonstrate the value of the additional information obtained from the
tomography. However, the marginal errors of the Gaussian Fisher
matrix are larger than the non-linear Fisher marginal errors for every
parameter which is a counterintuitive result. We show that the
projected contours from the Gaussian Fisher matrix are also larger
than the non-linear Fisher matrix, even though the volume of the
non-linear Fisher is 3.7 times larger than the Gaussian Fisher. This
warns us that the projected Fisher contours can be misleading over
complex multi-variate spaces and that larger 2-D contours do not
necessarily indicate a larger error volume.
\section*{Acknowledgments}
AK would like to thank Benjamin Joachimi and Tom Kitching for their
very useful discussions on this work, the European DUEL RTN project
MRTN-CT-2006-036133 and the University of Edinburgh for studentship
support.
\bibliographystyle{mn2e}
|
\section{#1}}
\newcommand{\sect}[1]{\subsection{#1}}
\newcommand{}{}
\newcommand{section}{section}
\newcommand{Section}{Section}
\newtheorem{thm}{Theorem}[section]
\newtheorem*{nonum}{Theorem}
\newtheorem{algorithm}[thm]{Algorithm}
\newtheorem{axiom}[thm]{Axiom}
\newtheorem{case}[thm]{Case}
\newtheorem{claim}[thm]{Claim}
\newtheorem{conclusion}[thm]{Conclusion}
\newtheorem{condition}[thm]{Condition}
\newtheorem{conjecture}[thm]{Conjecture}
\newtheorem{corollary}[thm]{Corollary}
\newtheorem{criterion}[thm]{Criterion}
\newtheorem*{thmMainTrue}{Theorem \ref{thm.main}}
\newtheorem*{thmMainGen}{Theorem \ref{thm.q.main}}
\newtheorem*{thmMainLag}{Theorem \ref{t.Lag.Perm}}
\newtheorem*{corClassify}{Corollary \ref{cor.classify}}
\theoremstyle{definition}
\newtheorem{exam}[thm]{Example}
\newtheorem{defn}[thm]{Definition}
\theoremstyle{plain}
\newtheorem{exercise}[thm]{Exercise}
\newtheorem{fact}[thm]{Fact}
\newtheorem{lem}[thm]{Lemma}
\newtheorem{notation}[thm]{Notation}
\newtheorem{problem}[thm]{Problem}
\newtheorem{prop}[thm]{Proposition}
\newtheorem{rem}[thm]{Remark}
\newtheorem{solution}[thm]{Solution}
\newtheorem{summary}[thm]{Summary}
\newcommand{\term}[1]{\textbf{#1}}
\newcommand{\mtrx}[1]{\begin{matrix} #1 \end{matrix}}
\newcommand{\pmtrx}[1]{\begin{pmatrix} #1 \end{pmatrix}}
\newcommand{\cmtrx}[1]{\left\{\begin{matrix}#1\end{matrix}\right\}}
\newcommand{\arry}[2]{\begin{array}{#1} #2 \end{array}}
\newcommand{\tbl}[2]{\begin{tabular}{#1} #2 \end{tabular}}
\newcommand{\RHScase}[1]{\left\{\begin{array}{ll} #1 \end{array}\right.}
\newcommand{\mathbb{C}}{\mathbb{C}}
\newcommand{\mathbb{N}}{\mathbb{N}}
\newcommand{\mathbb{R}}{\mathbb{R}}
\newcommand{\mathbb{Q}}{\mathbb{Q}}
\newcommand{\mathbb{Z}}{\mathbb{Z}}
\newcommand{\mathcal{A}}{\mathcal{A}}
\newcommand{\mathcal{B}}{\mathcal{B}}
\newcommand{\mathcal{C}}{\mathcal{C}}
\newcommand{\mathcal{H}}{\mathcal{H}}
\newcommand{\mathcal{L}}{\mathcal{L}}
\newcommand{\mathcal{O}}{\mathcal{O}}
\newcommand{\mathcal{Q}}{\mathcal{Q}}
\newcommand{\mathcal{R}}{\mathcal{R}}
\newcommand{\mathcal{T}}{\mathcal{T}}
\newcommand{\varepsilon}{\varepsilon}
\newcommand{\dset}[1]{\{1,\dots, #1\}}
\newcommand{\ddset}[2]{\{#1,\dots, #2\}}
\newcommand{\inv}[1]{#1^{-1}}
\input{thesis-commands}
\begin{document}
\title[Self-Inverses in Rauzy Classes]{Self-inverses in Rauzy Classes}
\author[J. Fickenscher]{Jonathan Fickenscher}
\address{Department of Mathematics, Rice University, Houston, TX~77005, USA}
\email{<EMAIL>}
\date{\today}
\maketitle
\pagestyle{headings}
\begin{abstract}
\noindent Thanks to works by M. Kontsevich and A. Zorich followed by C. Boissy, we have a classification of all Rauzy Classes of any given genus. It follows from these works that Rauzy Classes are closed under the operation of inverting the permutation. In this paper, we shall prove the existence of self-inverse permutations in every Rauzy Class by giving an explicit construction of such an element satisfying the sufficient conditions. As a corollary, we will give another proof that every Rauzy Class is closed under taking inverses. In the case of generalized permutations, generalized Rauzy Classes have been classified by works of M. Kontsevich, H. Masur and J. Smillie, E. Lanneau, and again C. Boissy. We state the definition of self-inverse for generalized permutations and prove a necessary and sufficient condition for a generalized Rauzy Class to contain self-inverse elements.
\end{abstract}
\setcounter{tocdepth}{2}
\tableofcontents
\input{chapter1}
\input{chapter2}
\input{chapter3}
\input{appendix-Lagrangian}
\bibliographystyle{abbrv}
|
\section{Introduction}\label{intro}
This paper lies in the lineage of recent works studying the
influence of some perturbations on the asymptotic spectrum of
classical random matrix models.
Such questions come from Statistics
(cf. \cite{John}) and appeared in the framework of empirical
covariance matrices.
In the pioneering work \cite{BBP05}, J. Baik, G. Ben Arous and S.
P\'ech\'e dealt with random sample covariance matrices
$(S_N)_N$ defined by
\begin{eqnarray}{\label{spike}}
S_N=\frac{1}{p} Y_N Y_N^* \mbox{~~with~~} Y= \Sigma_N^{\frac{1}{2}} B_N
\end{eqnarray}
where $B_N$ is a $N \times p$ complex matrix such that the entries $(B_N)_{ij}$ are i.i.d centered standard Gaussian and ${\Sigma}_N$
is a deterministic positive $N\times N$ matrix having all but finitely
many eigenvalues equal to one. This model can be seen as a multiplicative perturbation of the so-called white Wishart matrix for which ${\Sigma}_N=I_N$. Besides, the size of the samples $N$
and the size of the population $p=p_N$ are assumed of the same order
(as $N \to \infty$, $N/p \rightarrow c>0$). The global limiting behavior
of the spectrum of $S_N$ is not affected by such a matrix $\Sigma_N$.
Thus, the limiting spectral measure is the well-known
Marchenko-Pastur law (\cite{MP67})
defined by
\begin{equation} \label{MP} \mu_{\mbox{\tiny{MP}},{c}}(dx)=\max\{1-\frac{1}{c},0\}\delta_0
+f(x) 1\!\!{\sf I}_{[(1-\sqrt{c})^2;
(1+\sqrt{c})^2]}(x)dx\end{equation}
with $$ f(x)=\frac{\sqrt{\left(x-(1-\sqrt{c})^2\right)
\left((1+\sqrt{c})^2-x\right)}}{2\pi c x}.$$
When ${\Sigma}_N=I_N$, the largest eigenvalue of $S_N$ converges towards the right hand point of the support of the Marchenko-Pastur law (see \cite{Ge,BaiSilYi,YBK}). When ${\Sigma}_N\neq I_N$,
in \cite{BBP05} the authors pointed out a striking phase
transition phenomenon for the asymptotic behaviour of the largest eigenvalue of $S_N$ (at the convergence and fluctuations levels)
according to the value of the largest eigenvalue(s) of $\Sigma_N$.
They showed in particular that when
the largest eigenvalue of ${\Sigma}_N$ is far from one, the largest eigenvalue of $S_N$ converges outside the limiting
Marchenko-Pastur support. In \cite{BaikSil06},
J. Baik and
J. Silverstein extended the result of \cite{BBP05} on the convergence of the extremal eigenvalues of complex or real non
necessarily Gaussian matrices $S_N$ under finite four
moments assumptions on the distribution of the entries of $B_N$. \\
When $S_N$ is still defined by
(\ref{spike}),
but now the limiting spectral distribution of $\Sigma_N$ is some compactly supported measure $\nu$ on $[0;+ \infty[$, under finite second
moments assumptions on the distribution of the entries of $B_N$, the spectral distribution of $S_N$ converges almost surely towards
a probability measure $\mu_{\tiny{LSD}}$ which only depends on $c$ and $\nu$; denoting by
$g_{\mu_{\tiny{LSD}}}(z)= \int \frac{1}{z-x} d\mu_{\tiny{LSD}}(x)$ the Stieltjes transform of $\mu_{\tiny{LSD}}$,
for $z \in \mathbb{C}^+$, $g_{\mu_{\tiny{LSD}}}(z)$ is the unique solution $Z$ in $ \{ Z \in \mathbb{C}, -\frac{(1-c)}{z} -cZ \in \mathbb{C}^+ \} $ of the equation
\begin{equation}\label{MPeq} Z=\int \frac{1}{z- t(1-c +czZ)}d\nu(t) \end{equation}
(see \cite{MP67,BaiSil95,GS,KY,SSt,Wa,Yin86}).
Very recently R. Rao and J. Silverstein \cite{RaoSil09} and
Z.~D. Bai and J. Yao \cite{BaiYao08b} dealt with such a model assuming moreover that $\Sigma_N$
has a finite number of eigenvalues fixed outside the support of $\nu$ called spikes (or converging outside the support of $\nu$ in \cite{RaoSil09}), whereas the distance between the other eigenvalues of $\Sigma_N$ and the support of $\nu$ uniformly goes to zero. (Note that the assumptions in \cite{RaoSil09} are a bit more general).
Under finite four
moments assumptions on the distribution of the entries of $B_N$, the authors characterized the spikes of $\Sigma_N$ that generate jumps of eigenvalues of $S_N$
and described the corresponding limiting points
outside the support of the limiting spectral distribution $\mu_{\tiny{LSD}}$ of $S_N$.\\
Several authors considered an additive analogue of the above setting that is, the
influence on the asymptotic spectrum of
the addition of some Hermitian deterministic perturbation $A_N$ to the rescaled so-called Hermitian Wigner $N\times N$ matrix $W_N$.
\noindent Recall that, according to Wigner's work \cite{Wigner55,Wigner58}
and further results of different authors (see \cite{Bai99} for a review),
provided the common distribution $\mu $ of the entries is centered
with variance $\sigma ^2$, the large $N$-limiting spectral distribution
of the rescaled complex Wigner matrix $X_N=\frac{1}{\sqrt{N}} W_N$ is
the semicircle distribution $\mu _{\sigma }$ whose density is given by
\begin{equation}\label{scl}
\frac{d\mu _{\sigma }}{dx}(x)= \frac{1}{2 \pi \sigma ^2} \sqrt{4\sigma ^2- x^2}
\, 1 \hspace{-.20cm}1_{[-2\sigma , 2\sigma ]}(x).
\end{equation}
Moreover, if the fourth moment of the measure $\mu $ is finite,
the largest (resp. smallest) eigenvalue of $X_N$
converges almost surely towards the right (resp. left) endpoint $2\sigma $
(resp. $-2\sigma $) of the semicircular support
(cf. \cite{BaiYin88} or Theorem 2.12 in \cite{Bai99}).
\noindent Let $A_N$ be a deterministic Hermitian matrix such that the spectral measure of $A_N$ weakly converges to some probability measure $\nu$
and $\left\| A_N\right\| $ is uniformly bounded in $N$.
When $N$ becomes large, free probability provides us a good understanding
of the global behaviour of the spectrum of $M_N=X_N+A_N$ where $X_N$ is a rescaled complex Wigner matrix.
Indeed,
the spectral distribution of $M_N$ weakly converges to the free convolution $\mu _\sigma \boxplus \nu $
almost surely and in expectation
(cf \cite{AGZ09,MinSpe10} and \cite{Voiculescu91,Dykema93} for pioneering works).
We refer the reader to \cite{VDN92} for an introduction to free probability theory.
\noindent Dealing with small rank perturbation of a G.U.E matrix $W_N^G$,
S. P\'ech\'e pointed out an analogue phase transition phenomenon as in the sample covariance setting
for the convergence and the fluctuations of the largest eigenvalue of $M_N^G=W_N^G/{\sqrt{N}}+A_N$
with respect to the largest eigenvalue $\theta $ (independent of $N$) of $A_N$ \cite{Peche06}.
These investigations imply that, if $\theta $ is far enough from zero ($\theta > \sigma $),
then the largest eigenvalue of $M_N^G$ jumps above the support $[-2\sigma , 2\sigma ]$
of the limiting spectral measure and converges (in probability)
towards $\rho _\theta =\theta + \frac{\sigma ^2}{\theta }$.
Note that Z. F\"uredi and J. Koml$\acute{\text{o}}$s already exhibited such a phenomenon
in \cite{FurKom81} dealing with non-centered symmetric matrices.\\
In \cite{F�P�07}, D. F\'eral and S. P\'ech\'e proved that the results of \cite{Peche06}
still hold for a non-necessarily Gaussian Wigner Hermitian matrix $W_N$
with sub-Gaussian moments and in the particular case of a rank one perturbation matrix $A_N$
whose entries are all $\frac{\theta }{N}$ for some real number $\theta $.
In \cite{CDF09}, the authors considered a deterministic Hermitian matrix $A_N$
of arbitrary fixed finite rank $r$
and built from a family of $J$ fixed non-null real numbers
$\theta _1 > \cdots > \theta _J$ independent of $N$
and such that each $\theta _j$ is an eigenvalue of $A_N$ of fixed multiplicity $k_j$
(with $\sum_{j=1}^J k_j=r$).
They dealt with general Wigner matrices associated to some symmetric measure satisfying a Poincar\'e inequality.
They proved that eigenvalues of $A_N$ with absolute value strictly greater than $\sigma $
generate some eigenvalues of $M_N$ which converge
to some limiting points outside the support of $\mu _\sigma $.
In \cite{CDFF10}, the authors investigated the asymptotic behavior of the eigenvalues of generalized spiked perturbations of Wigner matrices associated to some symmetric measure satisfying a Poincar\'e inequality.
In this paper, the perturbation matrix $A_N$ is a deterministic Hermitian matrix whose spectral measure
converges to some probability measure $\nu $ with compact support and such that $A_N$ has a fixed number of fixed eigenvalues (spikes)
outside the support of $\nu $, whereas the distance between the other eigenvalues
and the support of $\nu $ uniformly goes to zero as $N$ goes to infinity.
It is established that only a particular subset of the spikes will generate
some eigenvalues of $M_N$ which will converge to some limiting points
outside the support of the limiting spectral measure. The phenomenon is completely analogous to the one described in \cite{RaoSil09} and
\cite{BaiYao08b} in the
sample covariance setting.\\
Now, one can wonder in the spiked deformed Wigner matrix setting as well as in the spiked sample covariance matrix setting, when some eigenvalues separate from the bulk, how the corresponding eigenvectors of the deformed model project onto those of the perturbation.
There are some results concerning finite rank perturbations: \cite{Paul} in the real Gaussian sample covariance matrix setting, and \cite{BGRao09} dealing with finite rank additive or multiplicative perturbations of unitarily invariant matrices.
For a general perturbation, up to our knowledge nothing has been done concerning eigenvectors of deformed Wigner matrices.
Dealing with sample covariance matrices, S. P\'ech\'e and O. Ledoit \cite{PL} introduced a tool to study the average behaviour of the eigenvectors but it seems that this did not allow
them to focus on the eigenvectors associated with the eigenvalues that separate from the bulk.
As already said, the limiting spectral distribution of the deformed Wigner model is described by the free convolution
of the respective limiting spectral distributions. Moreover,
the authors explained in \cite{CDFF10} that the phenomenon of the eigenvalues separating from the bulk can be fully described in terms of free probability
involving the subordination function related
to the free additive convolution of a semicircular distribution with the limiting spectral distribution of the perturbation.
Actually, as we will show below, the analogue results in the sample covariance matrix setting can also be
described in terms of free probability
involving the subordination function related
to the free multiplicative convolution of a Marchenko-Pastur distribution with the limiting spectral distribution of the perturbation.
Moreover, as already noticed by P. Biane in \cite{PB}, free probability again has something to tell us about eigenvectors of deformed matricial models. Indeed, in this paper, we are going to describe in the deformed Wigner matrix setting as well as in
the sample covariance matrix one, how the eigenvectors of the deformed model associated to the eigenvalues that separate from the bulk project onto those associated to the spikes of the perturbation, pointing out that the subordination functions relative to the free additive or multiplicative convolution play an important part in this asymptotic behavior. Note that the proof is exactly the same in the additive and the multiplicative cases.\\
In the sample covariance matrix model as well as in the deformed Wigner model, the convergence of the eigenvalues that separate from the bulk is deduced from a striking exact separation phenomenon, roughly stating that to each gap in the spectrum of the deformed model there corresponds a gap in the spectrum of the perturbation, these gaps splitting in exactly the same way the corresponding spectrum. For general deformed models, that is, dealing with other matrices than Wigner matrices in the additive case or other matrices than white Wishart matrices in the multiplicative case, such an exact separation phenomenon is not expected in full generality. Nevertheless,
we express all the results in this paper in terms of the free additive respectively multiplicative subordination functions since we conjecture that,
for other deformed models than deformed Wigner matrices and sample covariance matrices,
the limiting values of the eigenvalues that separate from the bulk as well as the limiting values of the orthogonal projection of the corresponding eigenvectors onto those associated to the spikes of the perturbation will be given by the same quantities
provided one deals with the corresponding subordination functions relative to the limiting spectral distribution of the non-deformed model. By the way, note that one can check that the results of F. Benaych-Georges and R. N. Rao in \cite{BGRao09}, concerning finite rank multiplicative or additive perturbation of a unitarily invariant matrix,
about the convergence of the extremal eigenvalues and of the projection of the corresponding eigenvectors onto those of the perturbation can be rewritten in terms of subordination functions as conjectured.
\\
The paper is organized as follows. In Section 2, we introduce the additive and multiplicative deformed models we consider in this paper; we also introduce some basic notations that will be used throughout the paper.
Section 3 is devoted to definitions and results concerning free convolutions and subordination functions, some of them being necessary to state our main result Theorem \ref{cvev1p1} in Section 4. Note that we present a common formulation for the additive and multiplicative deformed models
and a common proof in Section 5 and Section 6, postponing in Section 7 the technical results that need a specific study for each model. Finally, an Appendix gathers several tools that will be used in the paper.
\section{Models and Notations}
Let $\mu$ be a probability measure with variance $\sigma ^2$ which satisfies a Poincar\'e inequality with constant $C_{PI}$ (the definition of
such an inequality is recalled in the Appendix). Note that this condition implies that $\mu$ has moments of any order (see Corollary 3.2 and Proposition 1.10 in \cite{Ledoux01}).
In this paper, we will deform the
following classical matricial models.
\begin{itemize}
\item Normalized Wigner matrices $X^W_N=\frac{1}{\sqrt{N}} W_N $ \\
such that
$W_N$ is a $N \times N$ Wigner Hermitian matrix associated to
the distribution $\mu $: \\
$(W_N)_{ii}$, $\sqrt{2} \Re ((W_N)_{ij})_{i < j}$, $\sqrt{2}
\Im ((W_N)_{ij})_{i<j}$
are i.i.d., with distribution $\mu$.
\item Sample covariance matrices $X^S_N=\frac{1}{p}B_N B_N^*$\\such that
$B_N$ is a $N \times p$ matrix such that
$\sqrt{2} \Re ((B_N)_{ij})_{1\leq i\leq N, 1\leq j\leq p}$, $\sqrt{2}
\Im ((B_N)_{ij})_{{1\leq i\leq N, 1\leq j\leq p}}$
are i.i.d., with distribution $\mu $.
We assume that $\frac{N}{p} \rightarrow c>0$ when $N$ goes to infinity.
\end{itemize}
{\bf For Wigner matrices, we will assume moreover that $\mu$ is symmetric} since we will use results of \cite{CDFF10}
where this assumption is needed.\\
\noindent We will deform these models by respectively addition and multiplication by a deterministic Hermitian perturbation matrix $A_N$; {\bf in the multiplicative perturbation case, $A_N$ will be assumed to be nonnegative definite}.
In both cases, we assume that:\\
{\bf Assumption A}:\\
The eigenvalues $\gamma_i=\gamma_i(N)$ of $A_N$
are such that the spectral measure $\mu _{A_N} := \frac{1}{N} \sum_{i=1}^N \delta _{\gamma _i}$ weakly
converges to some probability measure $\nu $ with compact support.
We assume that there exists a fixed integer $r\geq 0$ (independent of $N$)
such that $A_N$ has $N-r$ eigenvalues $\beta _j(N)$ satisfying
\begin{equation}\label{univconv}
\max _{1\leq j\leq N-r} {\rm dist}(\beta _j(N),{\rm supp}(\nu ))\mathop{\longrightarrow } _{N \rightarrow \infty } 0,\end{equation}
where ${\rm supp}(\nu )$ denotes the support of $\nu $. \\
We also assume that there are $J$ fixed real numbers $\theta _1 > \ldots > \theta _J$
independent of $N$ which are outside the support of $\nu $ and such that each $\theta _j$
is an eigenvalue of $A_N$ with a fixed multiplicity $k_j$ (with $\sum_{j=1}^J k_j=r$).
The $\theta_j$'s will be called {\bf the spikes or the spiked eigenvalues} of $A_N$. The set of the spikes of $A_N$ will be denoted by
$\Theta$:
$$\Theta:=\{\theta_1; \ldots, \theta_J\}.$$ \\
{\bf In the sample covariance matrix setting we assume $\theta _J>0$.}\\
\noindent We will consider simultaneously the two deformed models:
$$M_N^W= X_N^W+A_N=\frac{1}{\sqrt{N}} W_N +A_N,$$
$$M_N^S=A_N^{\frac{1}{2}}X_N^S A_N^{\frac{1}{2}}=\frac{1}{p}A_N^{\frac{1}{2}}B_N B_N^*A_N^{\frac{1}{2}}.$$
{\bf When the approaches are the same for the two models we adopt the notation $M_N$ standing for both $M_N^W$
and $M_N^S$. When the studies are specific to one of the two models, we will use the superscripts.}\\
\noindent Actually, we assume without loss of generality in the sample covariance setting that the variance of $\mu$ is 1
since it corresponds to considering the rescaled matrix $\frac{M^S_N}{\sigma^2}$.\\
\noindent Throughout this paper, we will use the following notations.
\begin{itemize}
\item[-] $\mathbb{C}^+$ will denote the complex upper half plane $\{z \in \mathbb{C}, \, \Im z >0\}$. Similarly,
$\mathbb{C}^-$ will stand for
$\{z \in \mathbb{C}, \, \Im z <0\}$.
\item[-] For a function $f$ differentiable in some neighborhood of a point $x$ in $\mathbb{R}$, we will denote by $f^{'}(x)$
the derivative of $f$.
\item[-] For a vector subspace ${\cal V}$ of $\mathbb{C}^N$, we will denote by ${\cal V}^{\bot}$ its orthogonal supplementary
subspace
and by
$P_{\cal V}$ the orthogonal projection onto
${\cal V}$.
\item[-] $\langle ~, ~\rangle $ will denote the Hermitian inner product on $\mathbb{C}^N$ defined by $\langle a,b\rangle =b^*a$ for any $a,b$ in $\mathbb{C}^N$.
\item[-] $\Vert ~ ~\Vert_2 $ will denote the Euclidean norm on $\mathbb{C}^N$.
\item[-] We will denote by $M_{m\times q}(\mathbb{C})$ the set of $ m\times q$ matrices with complex entries.
$\Vert ~ \Vert $ will denote the operator norm
and for any matrix $M$, $\Vert M \Vert_2= \{Tr(MM^*)\}^{\frac{1}{2}}$.
\item[-] For any matrix $M$ in $M_{N\times N}(\mathbb{C})$, we will denote its kernel by Ker($M$).
\item[-] $E_{ij}$ in $M_{m\times q}(\mathbb{C})$ stands for the matrix such that $(E_{ij})_{kl}=\delta_{ik}\delta_{jl}$.
\item[-] For any $N\times N$ Hermitian matrix $M$, we will denote by
$$\lambda_1(M) \geq \ldots \geq \lambda_N(M)$$
\noindent its ordered eigenvalues.
\item[-] For a probability measure $\tau $ on $\mathbb {R}$,
we denote by $\mbox{supp}(\tau)$ its support and by $^c \mbox{supp}(\tau) $ its complement in $\mathbb {R}$.
\item[-] For a probability measure $\tau $ on $\mathbb {R}$,
we denote by $g_\tau $ its Stieltjes transform defined for $z \in \mathbb {C}\setminus \mathbb {R}$ by
$$g_\tau (z) = \int_\mathbb {R} \frac{d\tau (x)}{z-x}.$$
\item[-] $G_N$ denotes the resolvent of $M_N$ and
$g_N$ the mean of the Stieltjes transform of the spectral measure of $M_N$, that is,
$$g_N(z) = \mathbb {E}({\rm tr}_N G_N(z)), \, z \in \mathbb {C}\setminus \mathbb {R},$$
where ${\rm tr}_N$ is the normalized trace: ${\rm tr}_N =\frac{1}{N} {\rm Tr}$.\\
When it is necessary to distinguish the deformed Wigner matrix setting and the sample covariance matrix one, we will specify
the resolvent or the Stieltjes transform by using the corresponding superscript as follows:
$G_N^W, g_N^W$ and $G_N^S, g_N^S$.
\item[-] $C, K $ denote nonnegative constants which may vary from line to line.
\end{itemize}
As already mentioned in the introduction, the assumptions on $W_N$ and $A_N$ ensure that they are asymptotically free,
and then
the spectral distribution of $M_N^W$ weakly converges to the free convolution $\mu _\sigma \boxplus \nu $
almost surely and in expectation
(cf \cite{AGZ09,MinSpe10} and \cite{Voiculescu91,Dykema93} for pioneering works).
\noindent Concerning the sample covariance matrix model $M_N^S$, as already noticed in the introduction, its limiting spectral measure only depends on $c$ and $\nu$.
Note that when the entries of $B_N$ are Gaussian (that is, if $\mu$ is Gaussian) we can assume that $A_N$ is diagonal by the invariance under unitary conjugation of the distribution of $X^S_N$. Then, since according to Corollary 4.3.8
in \cite{HP},
$X^S_N$ and $A_N$ are asymptotically free, we can conclude that {\bf the limiting spectral distribution of $M_N^S$ is actually the free multiplicative convolution of the limiting spectral measure of $X_N^S$, that is, $ \mu_{\mbox{\tiny{MP}},{c}}$, with $\nu$, denoted by $ \mu_{\mbox{\tiny{MP}},{c}}\boxtimes \nu$.}
Thus, free additive and multiplicative convolutions provide a good understanding of the limiting global behaviour of the spectrum of the above deformed models. Moreover, \cite{BGRao09,BGRao10,CDFF10} show us that free probability can also allow to locate
isolated eigenvalues of deformed matricial models.
In particular, in \cite{CDFF10}, the authors point out that the subordination function relative to the free additive convolution provides a good understanding of the outliers of deformed Wigner matrices.
We will see in this paper that the subordination function relative to the free (additive or multiplicative) convolution
plays again an important part in the asymptotic behaviour of the eigenvectors relative to the outliers. We introduce in the following section some results concerning free convolution that will be fundamental later on.
\section{ Free convolution}\label{freeconv}
Free convolutions appear as natural analogues of the classical convolutions in the context of free probability theory.
Denote by ${\cal M}$ the set of probability measures supported on the real line and by ${\cal M}^+$ the ones supported on $[0; + \infty[$. For $\mu$ and $\nu$ in ${\cal M}$ one defines the free additive convolution $\mu \boxplus \nu$ of $\mu$ and $\nu$ as the distribution of $X+Y$ where $X$ and $Y$ are free self adjoint random variables with distribution $\mu$ and $\nu$. For $\mu$ and $\nu$ in ${\cal M}^+$, the free multiplicative convolution $\mu \boxtimes \nu$ of $\mu$ and $\nu$ is the distribution of $X^{\frac{1}{2}}YX^{\frac{1}{2}}$ where $X$ and $Y$ are free positive random variables with distribution $\mu$ and $\nu$. We refer the reader to \cite{VDN92} for an introduction to free probability theory
and to \cite{Voiculescu86,Voiculescu87} and \cite{BercoviciVoiculescu} for free convolutions.
In this section, we recall the analytic approach developped in \cite{Voiculescu86,Voiculescu87} to calculate the free convolutions of measures, we present
the important subordination property, and describe more deeply subordination functions relative to free additive convolution by a semi-circular distribution and free multiplicative convolution by a Marchenko-Pastur distribution.
We also recall characterizations of the complement of the support of these convolutions.
\subsection{Additive Free convolution}
Let $\tau $ be a probability measure on $\mathbb {R}$.
Its Stieltjes transform $g_\tau $ is analytic on the complex upper half-plane $\mathbb {C}^+$.
There exists a domain $$D_{\alpha , \beta } = \{ u+iv \in \mathbb {C}, |u| < \alpha v, v > \beta \}$$
on which $g_\tau $ is univalent.
Let $K_\tau $ be its inverse function, defined on $g_\tau (D_{\alpha , \beta })$, and
$$R_\tau (z) = K_\tau (z) - \frac{1}{z}.$$
Given two probability measures $\tau $ and $\nu $,
there exists a unique probability measure $\lambda $ such that
$$R_\lambda = R_\tau + R_\nu $$
on a domain where these functions are defined.
The probability measure $\lambda $ is called
the additive free convolution of $\tau $ and $\nu $ and denoted by $\tau \boxplus \nu $.
\subsubsection{Subordination property}
The free additive convolution of probability measures has an important property,
called subordination, which can be stated as follows.
\begin{proposition}
let $\tau $ and $\nu $ be two probability measures on $\mathbb {R}$;
there exists a unique analytic map $F^{(a)}: \mathbb {C}^+ \rightarrow \mathbb {C}^+$
such that \begin{equation} \label{subeq}\forall z \in \mathbb {C}^+ , ~~~~g_{\tau \boxplus \nu}(z)= g_\nu (F^{(a)}(z)),\end{equation}
$$\, F^{(a)}(\overline{z})=\overline{F^{(a)}(z)}, \, \Im F^{(a)}(z)\geq \Im z, \, \lim_{y \rightarrow + \infty} \frac{F^{(a)}(iy)}{iy}=1.$$
\end{proposition}
This phenomenon was first observed by D. Voiculescu under a genericity assumption in \cite{Voiculescu93},
and then proved in generality in \cite{Biane98} Theorem 3.1.
Later, a new proof of this result was given in \cite{BelBer07},
using a fixed point theorem for analytic self-maps of the upper half-plane.
\subsubsection{Free convolution by a semicircular distribution}
In \cite{Biane97b}, P. Biane provides a deep study of the free additive convolution by a semicircular distribution.
We first recall here some of his results that will be useful in our approach. Let $\nu $ be a probability measure on $\mathbb {R}$.
When $\tau$ in (\ref{subeq}) is the semi-circular distribution $\mu_\sigma$, let us denote by $F^{(a)}_{\sigma,\nu}$ the subordination function. In \cite{Biane97b},
P. Biane introduces the set
\begin{equation}\label{Ome}\Omega _{\sigma , \nu }:=\{ u+iv \in \mathbb {C}^+, v > v_{\sigma , \nu }(u)\},\end{equation}
where the function $v_{\sigma , \nu }: \mathbb {R} \rightarrow \mathbb {R}^+$ is defined by
$$v_{\sigma , \nu }(u) = \inf \left\{v \geq 0, \int_{\mathbb {R}} \frac{d\nu (x)}{(u-x)^2+v^2} \leq \frac{1}{\sigma ^2}\right\}.$$
The boundary of $\Omega _{\sigma , \nu }$ is the graph of the continuous function $v_{\sigma , \nu }$.
P. Biane proves the following
\begin{proposition}\cite{Biane97b}\label{homeo}
The map
\begin{equation}\label{defdeH} H_{\sigma , \nu }: z \longmapsto z+\sigma ^2 g_\nu (z)\end{equation}
is a homeomorphism from $\overline{\Omega _{\sigma , \nu }}$ to $\mathbb {C}^+ \cup \mathbb {R}$
which is conformal from $\Omega _{\sigma , \nu }$ onto $\mathbb {C}^+$.
$F^{(a)}_{\sigma , \nu }: \left\{\begin{array}{ll} \mathbb {C}^+ \cup \mathbb {R} \rightarrow \overline{\Omega _{\sigma ,\nu }}\\
z \rightarrow z-\sigma ^2g_{\mu _\sigma \boxplus \nu }(z) \end{array}\right.$
is the inverse function of $H_{\sigma , \nu }$.
\end{proposition}
\noindent Considering $ H_{\sigma , \nu }$ as an analytic map defined in the whole upper half-plane $\mathbb {C}^+$,
it can be easily seen that \begin{equation}\label{imageinverse}\Omega _{\sigma , \nu }=( H_{\sigma , \nu })^{-1}(\mathbb {C}^+). \end{equation}
\noindent {\bf In the following, we will denote $H_{\sigma , \nu }$ by $H$ to simplify the writing.}
\begin{remarque}\label{imagedeF}
Note that according to Proposition \ref{homeo},
$$F^{(a)}_{\sigma , \nu }(\mathbb{R}) =\partial \Omega_{\sigma , \nu } =\{u+i v_{\sigma , \nu }(u), u\in \mathbb{R}\}$$
\noindent so that we have the following equivalence
$$u\in F^{(a)}_{\sigma , \nu }(\mathbb{R}) \cap \mathbb{R} \Longleftrightarrow v_{\sigma , \nu }(u)=0.$$
\noindent The following characterization of the elements of the complement of the support of $\nu$ which are in the image of $\mathbb{R}$ by $F^{(a)}_{\sigma , \nu }$ readily follows:
\begin{equation}\label{imagereelle}u \in F^{(a)}_{\sigma , \nu }(\mathbb{R}) \cap \mathbb{R}\setminus {\rm supp~} (\nu) \Longleftrightarrow
u \in ^c{\rm supp~} (\nu) , H'(u) \geq 0.\end{equation}
\end{remarque}
\noindent
In \cite{Biane97b}, P. Biane obtains a description of the support of $\mu _\sigma \boxplus \nu $ from which,
when $\nu $ is a compactly supported probability measure,
the authors deduce in \cite{CDFF10} a characterization of the complement of the support
of $\mu _\sigma \boxplus \nu $ involving the support of $\nu $ and $H$.
\begin{proposition}\label{Caract}
$$x \in ^c {\rm supp}(\mu _\sigma \boxplus \nu ) \Leftrightarrow
\exists u \in {\cal O}^{(a)} {\rm ~such~that~} x=H(u)$$
where ${\cal O}^{(a)}$ is the open set
\begin{eqnarray} {\cal O}^{(a)}&:=&\left\{ u \in ^c {\rm supp}(\nu ), \, \, H'(u) > 0\right\} \nonumber \\& =&\left\{ u \in ^c {\rm supp}( \nu),\, \sigma^2 \int \frac{1}{(u-t)^2}d \nu(t) <1\right\} \label{defThetaW}.\end{eqnarray}
\end{proposition}
\begin{remarque}\label{oainclu} ${\cal O}^{(a)} \subset \partial \Omega_{\sigma , \nu }$
\end{remarque}
\noindent This readily follows from Remark \ref{imagedeF}.
\begin{remarque}\label{Hcroit}
Note that
if $u_1 < u_2$ are in $\left\{ u \in ^c {\rm supp~} (\nu), H'(u)\geq 0 \right\}$, one has
$H(u_1)\leq H(u_2).$
Indeed, by Cauchy-Schwarz inequality, we have
\begin{eqnarray*}
H(u_2)-H(u_1) & = & (u_2-u_1) \bigg [ 1 - \sigma ^2 \int_{\mathbb {R}}\frac{d\nu (x)}{(u_1-x)(u_2-x)} \bigg ] \\
& \geq & (u_2-u_1) \bigg [ 1 - \sigma ^2 \sqrt{ (-g'_{\nu }(u_1))(-g'_{\nu }(u_2))} \bigg ] \geq 0.
\end{eqnarray*}
\end{remarque}
\subsection{Multiplicative free convolution}
Let $\tau \neq \delta_0$ be a probability measure on $[0;+ \infty[$. Define the analytic function $$\Psi_\tau (z)=\int \frac{tz}{1-tz}d\tau(t)=\frac{1}{z}g_\tau(\frac{1}{z}) -1,$$
\noindent for complex values of $z$ such that $\frac{1}{z}$ is not in the support of $\tau$.
$\Psi_\tau$ determines uniquely the measure $\tau$ and it is univalent in the left half-plane $\{z \in \mathbb{C},\, \Re z <0\}$.
\noindent Then one may determine an analytic function $S_\tau$ such that
$$\Psi_\tau \left[ \frac{z}{z+1} S_\tau(z) \right]=z$$
\noindent in some domain (which will contain at least some interval to the left of zero) and then
$S_{\mu \boxtimes \nu}=S_\mu S_\nu.$ (see \cite{Voiculescu87}).
\subsubsection{Subordination property}
Free multiplicative convolution also presents a subordination phenomenon first proved in \cite{Biane98} (see also \cite{BelBer07}).
\begin{proposition}\label{submult}
Let $\tau \neq \delta_0$ and $\nu \neq \delta_0$ be two probability measures on $[0;+\infty[$. There exists a unique analytic
map $F^{(m)}_{\tau,\nu}$ defined on $\mathbb{C}\setminus [0;+\infty[$ such that
\begin{equation}\label{eqsubmult}\forall \, z \in \mathbb{C}\setminus [0;+\infty[, \, \Psi_{\nu\boxtimes \tau}(z) =\Psi_\nu(F^{(m)}_{\tau,\nu}(z))\end{equation}
and $$\forall \, z \in \mathbb{C}^+, \; F^{(m)}_{\tau,\nu}(z) \in \mathbb{C}^+,\, F^{(m)}_{\tau,\nu}(\overline{z})
=\overline{F^{(m)}_{\tau,\nu}(z)},\,
\, \arg (F^{(m)}_{\tau,\nu}(z))\geq \arg(z).$$
\end{proposition}
\subsubsection{Multiplicative free convolution with a Marchenko-Pastur distribution}
Let us determinate the subordination function relative to the free multiplication by a Marchenko-Pastur distribution.
We can deduce from (\ref{MPeq}) that for any $z \in \mathbb{C}^+$,
\begin{equation}\label{MPeq2} g_{ \mu_{\mbox{\tiny{MP}},c}\boxtimes \nu}(z)=\int \frac{1}{z- t(1-c +cz g_{ \mu_{\mbox{\tiny{MP}},{c}}\boxtimes \nu}(z))}d\nu(t)\end{equation}
and then that $\forall z \in \mathbb{C}\setminus [0;+\infty[$,
\begin{equation}\label{psi}\Psi_{ \mu_{\mbox{\tiny{MP}},c}\boxtimes \nu} (z)=\Psi_\nu \left(z-cz +cg_{ \mu_{\mbox{\tiny{MP}},c}\boxtimes \nu}(\frac{1}{z})\right).\end{equation}
\noindent Note that $$z-cz +cg_{ \mu_{\mbox{\tiny{MP}},c}\boxtimes \nu}(\frac{1}{z})=g_{\tau_{c,\nu}}(\frac{1}{z})$$
where {\bf $\tau_{c,\nu}$ is the limiting spectral distribution of $\frac{1}{p} B_N^* A_N B_N$}.\\
\noindent It is clear that $\forall z \in \mathbb{C}^+, \; g_{\tau_{c,\nu}}(\frac{1}{z}) \in \mathbb{C}^+,\, g_{\tau_{c,\nu}}(\frac{1}{\overline{z}})
=\overline{ g_{\tau_{c,\nu}}(\frac{1}{z})}.$ Moreover, since, using (\ref{MPeq2}), we have
$$g_{\tau_{c,\nu}}(\frac{1}{z})=z \left[(1-c) +c
\int \frac{1}{1- tg_{\tau_{c,\nu}}(\frac{1}{z})}d\nu(t)\right],$$
\noindent it is easy to see that
$\arg (g_{\tau_{c,\nu}}(\frac{1}{z}))\geq \arg(z).$\\
Therefore,
denoting by $F^{(m)}_{c,\nu}$ the subordination function in (\ref{eqsubmult}) when $\tau= \mu_{\mbox{\tiny{MP}},c}$, we have that
$$ F^{(m)}_{c,\nu}(z)= z-cz +cg_{ \mu_{\mbox{\tiny{MP}},c}\boxtimes \nu}(\frac{1}{z})=g_{\tau_{c,\nu}}(\frac{1}{z}).$$
Now, we are going to present the characterization of the complement of the support of
$\tau_{c,\nu}$ provided by Choi and Silverstein in \cite{CHOISILVER} . Note that the supports of $ \mu_{\mbox{\tiny{MP}},c}\boxtimes \nu$ and $\tau_{c,\nu}$ obviously coincide on $]0;+\infty[$.\\
~~
\noindent According to \cite{BaiSil06} p 113, for each $z \in \mathbb{C^+}$, $g_{\tau_{c,\nu}}(z)$ is the unique solution $Z$ in $\mathbb{C}^{-}$
of the equation \begin{equation} Z =\frac{1}{z-c\int \frac{t}{1-tZ}d\nu(t)}\end{equation}
so that \begin{equation}\label{composition} z= {\cal Z}_{c,\nu}(g_{\tau_{c,\nu}}(z))\end{equation}
where
\begin{equation}\label{defzcnu}{\cal Z}_{c,\nu}(x)=\frac{1}{x}+c \int \frac{t}{1-tx}d\nu(t).\end{equation}
\noindent {\bf In the following, we will denote ${\cal Z}_{c,\nu}$ by ${\cal Z}$ to simplify the writing.}
\begin{proposition}\label{ChoiSilver}\cite{CHOISILVER}
If $u \in ^c {\rm supp}( \tau_{c,\nu})$, then $s=g_{\tau_{c,\nu}}(u)$ satisfies
\begin{enumerate}
\item $s \in \mathbb{R}\backslash\{0\},$
\item $\frac{1}{s} \in ^c {\rm supp}( \nu)$,
\item ${\cal Z}{'}(s)<0$.
\end{enumerate}
Conversely if $s$ satisfies (1), (2) and (3), then $u = {\cal Z}(s)\in ^c {\rm supp}( \tau_{c,\nu})$.
\end{proposition}
\noindent In particular, letting $z$ converge towards any element $u$ of $^c \mbox{supp}( \tau_{c,\nu})$ in (\ref{composition})
leads to \begin{equation}\label{compodansunsens}u= {\cal Z}(g_{\tau_{c,\nu}}(u)).\end{equation}
\noindent In \cite{CHOISILVER}, the authors proved also that $\lim_{\begin{array}{ll}z\rightarrow u\\z \in \mathbb{C}^+ \end{array}} g_{\tau_{c,\nu}}(z)=g_{\tau_{c,\nu}}(u)$ exists for any $u$ in $\mathbb{R}\setminus\{0\}$.
\noindent We include here for the convenience of the reader some basic facts that will be used later on.
\begin{remarque}\label{compositionlimite}
For any $u \in \mathbb {R}\setminus \{0\}$ such that $g_{\tau_{c,\nu}}(u)\in \mathbb{R}\setminus\{0\}$ and $\frac{1}{g_{\tau_{c,\nu}}(u)} \in ^c {\rm supp~}(\nu)$, we have ${\cal Z}[g_{\tau_{c,\nu}}(u)]=u$.
\end{remarque}
\noindent This readily follows by letting $z$ goes to $u$ in (\ref{composition}).\\
\noindent Let us introduce the open set
\begin{eqnarray} \label{defThetaS} {\cal O}^{(m)}&:=&\left\{ u \in ^c \mbox{supp}( \nu)\setminus \{0\},\, {\cal Z}^{'}(\frac{1}{u})<0\right\} \nonumber \\& =&\left\{ u \in ^c \mbox{supp}( \nu)\setminus \{0\},\, c \int \frac{t^2}{(u-t)^2}d \nu(t) <1\right\} \label{defThetaS}.\end{eqnarray}
\begin{remarque}\label{CScomp}
For any $u$ in ${\cal O}^{(m)}$, $g_{\tau_{c,\nu}}[{\cal Z}(\frac{1}{u})]=\frac{1}{u}$.
\end{remarque}
\noindent Let us prove Remark \ref{CScomp}. According to Proposition \ref{ChoiSilver}, for any $u \in {\cal O}^{(m)}$,
${\cal Z}(\frac{1}{u}) \in ^c \mbox{supp}( \tau_{c,\nu})$ and then according to the same Proposition \ref{ChoiSilver},
$\frac{1}{g_{\tau_{c,\nu}}[{\cal Z}(\frac{1}{u})]}$ also belongs to ${\cal O}^{(m)}$. Now, for any $a\neq b$ in ${\cal O}^{(m)}$,
\begin{equation}\label{croi} {\cal Z}(\frac{1}{b})-{\cal Z}(\frac{1}{a})={(b-a)} \left[ 1-c \int \frac{t^2}{({a}-t)({b}-t)} d \nu(t) \right];\end{equation}
\noindent by Cauchy-Schwartz inequality,
\begin{eqnarray}\hspace*{-5mm}\left| c \int \frac{t^2}{({a}-t)({b}-t)} d \nu(t) \right| &\leq & \left\{c\int \frac{t^2}{({a}-t)^2} d \nu(t)\right\}^{\frac{1}{2}} \left\{c\int \frac{t^2}{(b-t)^2} d \nu(t)\right\}^{\frac{1}{2}} \label{borne}\\\hspace*{-5mm}& < &1.
\nonumber
\end{eqnarray}
Hence we can conclude that
\begin{equation} \label{difference} \mbox{~ for any~} a\neq b\; \mbox{~in~} {\cal O}^{(m)}, \, {\cal Z}(\frac{1}{b}) \neq {\cal Z}(\frac{1}{a}).\end{equation}
Since using (\ref{compodansunsens}) we have ${\cal Z}[g_{\tau_{c,\nu}}[{\cal Z}(\frac{1}{u})]]={\cal Z}(\frac{1}{u})$, we can then deduce that $g_{\tau_{c,\nu}}[{\cal Z}(\frac{1}{u})]=\frac{1}{u}.$ $\Box$\\
~~
\begin{remarque}\label{croissance globaleS} Using (\ref{croi}) and (\ref{borne}), we have
for any $a \leq b$ in the set $\left\{ u \neq 0,\, u \in ^c {\rm supp}( \nu),\, {\cal Z}^{'}(\frac{1}{u})\leq 0\right\}$ that ${\cal Z}(\frac{1}{a}) \leq {\cal Z}(\frac{1}{b})$.
\end{remarque}
\begin{remarque}\label{nonnul}
For any $u \in {\cal O}^{(m)}\cap ]0;+\infty[$, we have $ {\cal Z}(\frac{1}{u})>0$.
\end{remarque}
\noindent Indeed, assume that $ {\cal Z}(\frac{1}{u}) \leq 0$.
According to Proposition \ref{ChoiSilver}, ${\cal Z}(\frac{1}{u})\in ^c \mbox{supp}( \tau_{c,\nu})$; $ {\cal Z}(\frac{1}{u}) \leq~0$ implies that ${\cal Z}(\frac{1}{u})$ is on the left hand side of $ \mbox{supp}( \tau_{c,\nu})$ and therefore that
$g_{\tau_{c,\nu}}({\cal Z}(\frac{1}{u}))\leq 0$. This leads to a contradiction with Remark \ref{CScomp} saying that $g_{\tau_{c,\nu}}({\cal Z}(\frac{1}{u}))=\frac{1}{u}>0.$
\begin{remarque}\label{pasdansimagedeg}
For any $u\neq 0$ in $^c {\rm supp~}( \nu)$ such that ${\cal Z}^{'}(\frac{1}{u})> 0$, we have $\frac{1}{u} \notin
g_{\tau_{c,\nu}} (\mathbb{R}\setminus\{0\})$.
\end{remarque}
\noindent Indeed, let us assume that there exists $v \in \mathbb{R}\setminus\{0\}$ such that $\frac{1}{u} =
g_{\tau_{c,\nu}}(v)$. According to Proposition \ref{ChoiSilver}, $v$ belongs to $\mbox{supp}( \tau_{c,\nu})$.
Using Remark \ref{compositionlimite}, since $g_{\tau_{c,\nu}}(v)\in \mathbb{R}\setminus\{0\}$ and $\frac{1}{g_{\tau_{c,\nu}}(v)} \in ^c \mbox{supp~}(\nu)$, we have ${\cal Z}[g_{\tau_{c,\nu}}(v)]=v$. It follows that for any $y>0$,
\begin{eqnarray}
1&=& \frac{{\cal Z}[g_{\tau_{c,\nu}}(v+iy)]-{\cal Z}[g_{\tau_{c,\nu}}(v)]}{iy} \nonumber\\
&=&
\frac{{\cal Z}[g_{\tau_{c,\nu}}(v+iy)]-{\cal Z}[g_{\tau_{c,\nu}}(v)]}{g_{\tau_{c,\nu}}(v+iy)-g_{\tau_{c,\nu}}(v)} \times \frac{g_{\tau_{c,\nu}}(v+iy)-g_{\tau_{c,\nu}}(v)} {iy} .\label{fractions}
\end{eqnarray}
Since $g_{\tau_{c,\nu}}(v+iy)$ converges towards $g_{\tau_{c,\nu}}(v) =\frac{1}{u}$ and $ {\cal Z} $ is holomorphic in a neighborhood of $g_{\tau_{c,\nu}}(v)$, letting $y $ tends to zero the first factor on the right hand side of (\ref{fractions}) converges towards ${\cal Z}^{'} (g_{\tau_{c,\nu}}(v))= {\cal Z}^{'}(\frac{1}{u})>0.$ This implies that $\Re\left[ \frac{g_{\tau_{c,\nu}}(v+iy)-g_{\tau_{c,\nu}}(v)} {iy}\right]$ converges towards $\frac{1}{{\cal Z}^{'}(\frac{1}{u})}>0$ when $y$ tends to zero. Now, for any $y>0$, we have
\begin{eqnarray*}\Re\left[ \frac{g_{\tau_{c,\nu}}(v+iy)-g_{\tau_{c,\nu}}(v)} {iy}\right]&=&
\frac{\Im \left[ g_{\tau_{c,\nu}}(v+iy)-g_{\tau_{c,\nu}}(v)\right]} {y}\\
&=&\frac{\Im g_{\tau_{c,\nu}}(v+iy)}{y} <0
\end{eqnarray*}
\noindent which leads to a contradiction. $\Box$
\section{Main results}
As noticed in the previous section, we have the following characterization of the complement of the support of the limiting spectral distribution of $M_N^W$.
\begin{equation}\label{car1}x \in ^c{\rm supp}(\mu _\sigma \boxplus \nu ) \Leftrightarrow
\exists u \in {\cal O}^{(a)} {\rm ~such~that~} x=H(u) .\end{equation}
Moreover, we can deduce, using Proposition \ref{homeo} and Remark \ref{oainclu}, that :\\
$x \mapsto F^{(a)}_{\sigma,\nu}\left(x\right)$ is a bijection from $ ^c{\rm supp}(\mu _\sigma \boxplus \nu )$ onto ${\cal O}^{(a)}$ with inverse $H$.\\
\noindent Since the supports of $ \mu_{\mbox{\tiny{MP}},c}\boxtimes \nu$ and $\tau_{c,\nu}$ coincide on $]0;+ \infty[$, we can also deduce from the previous section the following characterization of the restriction to $\mathbb{R}\setminus\{0\}$ of the complement of the support of the limiting spectral distribution of $M_N^S$:
\begin{equation}\label{car2}x \neq 0, x \in ^c \mbox{supp}( \mu_{\mbox{\tiny{MP}},c}\boxtimes \nu) \Leftrightarrow \exists u \in {\cal O}^{(m)}, {\cal Z} (\frac{1}{u}) \neq 0, {\rm ~such~that~}
x={\cal Z} (\frac{1}{u}).\end{equation}
\noindent Moreover, we can deduce from Remark \ref{compositionlimite} and Remark \ref{CScomp} that:\\
$x \mapsto F^{(m)}_{c,\nu}\left(\frac{1}{x}\right) (=g_{\tau_{c,\nu}}(x))$ is a bijection from $ ^c \mbox{supp}( \mu_{\mbox{\tiny{MP}},c}\boxtimes \nu)\setminus \{0\}$ onto the set $\left\{\frac{1}{u}, u \in {\cal O}^{(m)}, {\cal Z} (\frac{1}{u})\neq 0\right\}$ with inverse ${\cal Z}$.\\
Note that the limiting mass at zero was studied in \cite{CHOISILVER}:
$$ \mu_{\mbox{\tiny{MP}},c}\boxtimes \nu (0) =\left\{\begin{array}{ll}\nu(0) \mbox{~if~} c(1-\nu(0))\leq 1,\\
1- \frac{1}{c} \mbox{~if~} c(1-\nu(0))> 1.\end{array} \right.$$
Actually, according to \cite{CDFF10} (resp. \cite{RaoSil09,BaiYao08b}),
the spikes $\theta_j$ in $\Theta=\{\theta_1;\ldots;\theta_J\}$ of the perturbation matrix $A_N$ that will generate eigenvalues of $M_N^W$ (resp. $M_N^S$) which deviate from the bulk are exactly those belonging to ${\cal O}^{(a)}$ (resp. ${\cal O}^{(m)}$)
and the corresponding limiting points
outside the support of $\mu _\sigma \boxplus \nu $ (resp. $\tau_{c,\nu}$) will be given by
$H(\theta_j)$ (resp. ${\cal Z} (\frac{1}{\theta_j})$).
Note that the results in \cite{RaoSil09,BaiYao08b} are not formulated in that way since the authors do not
deal with subordination function but as already mentioned we choose to express all the results
using these functions $H$ and ${\cal Z}$ related to the subordination functions for further generalizations.
Hence adopting the notations of the first column of the following array standing for both the corresponding elements of the second column (deformed Wigner matrix case) and the third column (sample covariance matrix case),
we can present
a common formulation of these results.
\begin{equation}\label{correspondance}\mbox{
\begin{tabular}{|c|c|c|} \hline
$M_N$& $M_N^W$ & $ M_N^S$\\ \hline
$\mu_{\mbox{\tiny{LSD}}}$& $ \mu _\sigma \boxplus \nu$ & $ \mu_{\mbox{\tiny{MP}},c}\boxtimes \nu$\\ \hline
${\cal O} $&${\cal O}^{(a)}$ &${\cal O}^{(m)}$\\ \hline
$ \Theta_o$ & $\Theta \cap {\cal O}^{(a)}=\{\theta_i \in \Theta, H' (\theta_j)>0\}$ &$ \Theta \cap {\cal O}^{(m)}=\{\theta_i \in \Theta, {\cal Z}^{'}(\frac{1}{\theta_j})<0\} $\\ \hline
$ \rho _{\theta _j}$& $H(\theta_j)$& ${\cal Z}(\frac{1}{\theta_j}) $\\ \hline
\end{tabular}}
\end{equation}
\begin{theoreme}{\label{ThmASCV}}\cite{RaoSil09,BaiYao08b,CDFF10}
Let $\theta_j$ be in $ \Theta_o$
and denote by $n_{j-1}+1, \ldots , n_{j-1}+k_j$ the descending ranks of $\theta _j$ among the eigenvalues of $A_N$.
Then the $k_j$ eigenvalues $(\lambda_{n_{j-1}+i}(M_N), \, 1 \leq i \leq k_j)$
converge almost surely outside the support of $\mu_{\mbox{\tiny{LSD}}}$
towards $\rho _{\theta _j}$.
Moreover, these eigenvalues asymptotically separate from the rest of the spectrum since
(with the conventions that $\lambda_0(M_N)=+\infty$ and $\lambda_{N+1}(M_N)=-\infty$)
there exists $0< \delta_0 $ such that
\noindent almost surely for all large N, \begin{equation}\label{sepraj}\lambda_{n_{j-1}}(M_N) > \rho _{\theta _j} + \delta_0 \, \mbox{~and~} \, \lambda_{n_{j-1}+k_j +1}(M_N) < \rho _{\theta _j} - \delta_0
.\end{equation}
\end{theoreme}
The aim of this paper is to study how the corresponding eigenvectors of the deformed model project onto those of the perturbation.
Here is the main result of the paper still adopting the notations of the first column of the array (\ref{correspondance}) in order to present
a unified approach.
\begin{theoreme}\label{cvev1p1}
Let $\theta_j$ be in $\Theta_o$
and denote by $n_{j-1}+1, \ldots , n_{j-1}+k_j$ the descending ranks of $\theta _j$ among the eigenvalues of $A_N$.
Let
$\xi(j)$ be a normalized eigenvector of $M_N$ relative to one of the eigenvalues $(\lambda_{n_{j-1}+q}(M_N)$, $1\leq q\leq k_j)$.
Then,
when $N$ goes to infinity,
\begin{itemize}
\item[(i)] $\left\| P_{\mbox{Ker~}(\theta_j I_N-A_N)}\xi(j)\right\|^2_2 \stackrel{a.s}{\rightarrow} \tau(\theta_j)$\\
~~
\noindent where \begin{equation}\label{tau}\tau(\theta_j)=\left\{\begin{array}{ll} H'(\theta_j) \mbox{~if~} M_N=M_N^W,\\
- \frac{{\cal Z}^{'}(\frac{1}{\theta_j})}{\theta_j {\cal Z}(\frac{1}{\theta_j}) } \mbox{~if~} M_N=M_N^S.
\end{array}\right.\end{equation}
\noindent Note that we have explicitly $$H'(\theta_j) = 1 -\sigma^2 \int \frac{1}{(\theta_j-x)^2}d\nu(x),$$
$$- \frac{{\cal Z}'(\frac{1}{\theta_j})}{\theta_j {\cal Z}(\frac{1}{\theta_j}) }=
\frac{1- c\int \frac{x^2}{(\theta_j-x)^2}d\nu(x)}{1+c \int \frac{x}{(\theta_j-x)}d\nu(x)}.$$
\item[(ii)] for any $\theta_l$ in $\Theta\setminus\{\theta_j\}$,
$$\left\| P_{\mbox{Ker~}(\theta_l I_N-A_N)}\xi(j)\right\|_2 \stackrel{a.s}{\rightarrow} 0.$$
\end{itemize}
\end{theoreme}
\noindent {\bf Example:} Let us consider the perturbation matrix
$$A_N = \displaystyle{{\rm{ diag
}}(2, \frac{3}{2},0,\underbrace{-1,\ldots
,-1}_{\frac{N}{2}}, \underbrace{1,\ldots,
1 }_{\frac{N}{2}-3})},$$ whose limiting spectral distribution is
$\nu =\frac{1}{2}\delta_1 + \frac{1}{2}\delta_{-1}$. Thus, the set of the spikes of $A_N$ is $\Theta=\{2; \frac{3}{2};0\}$.
Let us consider the corresponding deformed Wigner model assuming moreover that $\sigma^2=\frac{1}{2}$. Then,
$H(u)=u+\frac{1}{4}\frac{1}{(u-1)}+\frac{1}{4}\frac{1}{(u+1)}$. One can check that the support of $\mu_{\frac{1}{\sqrt{2}}}\boxplus \nu$ has two connected components which are symmetric with respect to zero.
Since $H'(2)=\frac{13}{18}>0$ and $2$ is the largest eigenvalue of $A_N$, according to Theorem \ref{ThmASCV}, when $N$ goes to infinity,
the largest eigenvalue of the deformed Wigner model $M_N^W$ converges almost surely towards $H(2)=\frac{7}{3}$ (on the right hand side of the support of
$\mu_{\frac{1}{\sqrt{2}}}\boxplus \nu$).
Note that, since $H'(\frac{3}{2})<0$, the second largest eigenvalue of $M_N^W$ sticks to the bulk.
Moreover, since $H'(0)=\frac{1}{2}>0$ and the descending rank of $0$ among the eigenvalues of $A_N$ is $\frac{N}{2}$,
according to Theorem \ref{ThmASCV}, when $N$ goes to infinity, $\lambda_{\frac{N}{2}}(M_N^W)$ converges almost surely towards $H(0)=0$ which is
between the two connected components of the support of $\mu_{\frac{1}{\sqrt{2}}}\boxplus \nu$.
Now, denote by $\{e_1,\ldots,e_N\}$ the canonical basis of $\mathbb{C}^N$.
Since $e_1$ is an eigenvector relative to $2$, $e_2$ is an eigenvector relative to $\frac{3}{2}$ and $e_{3}$ is an eigenvector relative to $0$, according to Theorem \ref{cvev1p1}, if $\xi$ denotes a normalized eigenvector associated to the largest eigenvalue of $M_N^W$, $^t \xi=\left( \xi^{(1)},\ldots,\xi^{(N)}\right)$, then
$\vert\xi^{(1)}\vert \stackrel{a.s}{\rightarrow} \sqrt{H'(2)}=\frac{\sqrt{13}}{3\sqrt{2}}$ and, for $i=2,3$, $\vert\xi^{(i)}\vert \stackrel{a.s}{\rightarrow} 0$
when $N$ goes to infinity. Similarly, if $\xi$ denotes a normalized eigenvector associated to $\lambda_{\frac{N}{2}}(M_N^W)$, $^t \xi=\left( \xi^{(1)},\ldots,\xi^{(N)}\right)$, then
$\vert\xi^{(3)}\vert \stackrel{a.s}{\rightarrow} \sqrt{H'(0)}=\frac{1}{\sqrt{2}}$ and, for $i=1,2$, $\vert\xi^{(i)}\vert \stackrel{a.s}{\rightarrow} 0$
when $N$ goes to infinity. \\
Actually, in order to establish Theorem \ref{cvev1p1}, we will first prove Proposition \ref{cvev} below since
when $k_j \neq 1$, the method used in this paper does not allow us to tackle directly the orthogonal projection of each eigenvector separately to prove (i).
\begin{proposition}\label{cvev} Let $\theta_j$ be in $\Theta_o$
and denote by $n_{j-1}+1, \ldots , n_{j-1}+k_j$ the descending ranks of $\theta _j$ among the eigenvalues of $A_N$.
Denote by
$\xi_1(j),\ldots, \xi_{k_j}(j)$ an orthonormal system of eigenvectors associated to $(\lambda_{n_{j-1}+n}(M_N)$, $1\leq n\leq k_j)$.
Then, for any $\theta_l $ in $\Theta$,
when $N$ goes to infinity, $$\sum_{n=1}^{k_j}\left\| P_{\mbox{Ker~}(\theta_l I_N-A_N)}\xi_n(j)\right\|^2_2 \stackrel{a.s}{\rightarrow} \delta_{l,j}k_j\tau(\theta_j)$$
where $\tau(\theta_j)$ is defined by (\ref{tau}).
\end{proposition}
In Section \ref{redu}, we will explain how we can
deduce (i) of Theorem \ref{cvev1p1} from Proposition \ref{cvev} using a perturbation trick and ideas of \cite{EI}.\\
~~
\section{Reduction of the proof of (i) of Theorem \ref{cvev1p1} to the case of a spike with multiplicity one}\label{redu}
Note first that, dealing with a spike $\theta_j$ in $\Theta_o$ with multiplicity one, the statements of Theorem \ref{cvev1p1} and Proposition \ref{cvev}
are equivalent. Thus, in this section, we show how to deduce (i) of Theorem \ref{cvev1p1} dealing with a spike $\theta_j$ with multiplicity $k_j \neq 1$
from the hypothesis that (i) is true dealing with a spike with multiplicity one.
We will need the following lemmas.
\begin{lemme}\label{Proj}
Let ${\cal V}$ be a vector subspace of $\mathbb{C}^N$ with an orthonormal basis $V_1,\ldots,V_k$.
Let $\alpha$ be in $[0;+\infty[$.
For any $m=1,\ldots,k$, let $\alpha_m$ be in $[0;+\infty[$.
Let $E$ be a vector subspace of $\mathbb{C}^N$ with an orthonormal basis $\xi_1,\ldots,\xi_k$.
Then, there exists a sequence $a_N \geq 0$, depending on the $ \vert \langle V_i,\xi_n\rangle\vert$, $ \alpha_m$, $i,n,m \in \{1,\ldots,k\}^3$,
such that, for any vector $u$ in the unit sphere of $ \mathbb{C}^N$,
$$\left| \left\|P_{\cal V} u\right\|_2^2-\alpha\right| \leq (2k+\alpha) \left\|P_{{E}^{\bot}} u \right\|_2 +\max_{m=1}^{k} \left| \alpha_m-\alpha \right| +a_N,$$
\noindent and, if, for any $ m,n$ in $\{1,\ldots,k\}^2$, \begin{equation}\label{hypconv} \vert \langle V_m,\xi_n\rangle\vert^2
\rightarrow \delta_{m,n} \alpha_m\mbox{~~~when $N$ goes to infinity},
\end{equation}
then $a_N $ converges to zero when $N$ goes to infinity.
\end{lemme}
\noindent {\bf Proof:} Throughout the proof, we will often use the following obvious inequalities without mentioning them:\\
for any vectors $u_1$ and $u_2$ in the unit sphere of $ \mathbb{C}^N$,
$\left| \langle u_1u_2\rangle \right| \leq \left\| u_1 \right\|_2 \left\| u_2 \right\|_2=1$,
$\left| \langle P_{E}u_1,u_2\rangle \right|\leq \left\| P_{E} u_1 \right\|_2 \left\|u_2\right\|_2 \leq \left\| u_1 \right\|_2 \left\| u_2 \right\|_2=1.$\\
~~
\noindent We have for each $m$ in $\{1,\ldots,k\}$, for any vector $u$ in the unit sphere of $ \mathbb{C}^N$,\\
~~~~$\left| ~\left| \langle u,V_m\rangle \right|^2 -\left| \langle P_{E} u,V_m\rangle \right|^2 ~ \right|$
\begin{eqnarray} &=&\left| \left( \left| \langle u,V_m\rangle \right| -\left| \langle P_{E}u,V_m\rangle \right|\right) \right| \nonumber \\
&&~~~\times \left( \left| \langle u,V_m\rangle \right| +\left| \langle P_{E}u,V_m\rangle \right|\right) \nonumber\\
&\leq& 2 \left| \left( ~\left| \langle u,V_m\rangle \right| -\left| \langle P_{E}u,V_m\rangle \right|\right)\right|.\label{inin}\end{eqnarray}
\noindent From $$ \langle u,V_m\rangle = \langle P_{E}u,V_m\rangle + \langle P_{ E^{\bot}} u,V_m\rangle,$$ \noindent
it readily follows that \begin{equation}\label{inin2} \left|~ \left| \langle u,V_m\rangle \right|-\left| \langle P_{E}u,V_m\rangle \right|~ \right| \leq \left| \langle P_{ E^{\bot}}u ,V_m\rangle \right|.\end{equation}
(\ref{inin}) and (\ref{inin2}) readily yield
\begin{eqnarray} \left| ~\left| \langle u,V_m\rangle \right|^2 -\left| \langle P_{E}u,V_m\rangle \right|^2 ~ \right|& \leq& 2 \left| \langle P_{ E^{\bot}}u ,V_m\rangle \right| \nonumber\\
& \leq & 2 \left\| P_{ E^{\bot}}u \right\|_2 \left\| V_m \right\|_2 \nonumber \\& \leq & 2 \left\| P_{ E^{\bot}}u \right\|_2 .
\label{proj1}\end{eqnarray}
\noindent Now, we have that
\begin{eqnarray*}\langle P_{E}u,V_m\rangle&=& \langle u,\xi_m\rangle \langle \xi_m,V_m\rangle\\
&+&\sum_{\begin{array}{ll}n=1\\n \neq m\end{array} }^{k}\langle u,\xi_n\rangle \langle \xi_n,V_m\rangle \\
&=&\langle u,\xi_m\rangle \langle \xi_m,V_m\rangle+\Delta_m(u) \end{eqnarray*}
where\begin{equation}\label{Delta} \left| \Delta_m(u)\right| \leq
\sum_{\begin{array}{ll}n=1\\n \neq m\end{array} }^{k} \left|\langle \xi_n,V_m\rangle \right|.\end{equation}
\noindent Then,
$$\left| \langle P_{E}u,V_m\rangle \right|^2=\left|\langle u,\xi_m\rangle \right|^2 \left|\langle \xi_m,V_m\rangle \right|^2+\nabla_m(u)$$
\noindent with \begin{equation}\label{nabla} \left| \nabla_m(u) \right| \leq 2 \left| \Delta_m(u)\right| +\left| \Delta_m(u)\right|^2. \end{equation}
Thus,
\begin{eqnarray}\left| \langle P_{E} u,V_m\rangle \right|^2&=&\alpha \left| \langle u, \xi_m\rangle \right|^2\nonumber \\ & & + \left(\alpha_m-\alpha\right)\left| \langle u,\xi_m\rangle \right|^2
\nonumber \\ & &+ \left(\left|\langle \xi_m,V_m\rangle \right|^2-\alpha_m\right) \left| \langle u,\xi_m\rangle \right|^2 \\&& + \nabla_m(u). \label{proj}
\end{eqnarray}
Now, using that $$1= \left\| u \right\|_2^2=\sum_{m=1}^{k} \left| \langle u, \xi_m \rangle \right|^2 + \left\| P_{ E^{\bot}}u \right\|_2^2,$$
we have
\begin{eqnarray*}\sum_{m=1}^{k}\left| \langle u,V_m\rangle \right|^2 -\alpha &=&\sum_{m=1}^{k} \left[ \left| \langle u,V_m\rangle \right|^2 -\alpha \left| \langle u, \xi_m \rangle \right|^2 \right]\\
&& -\alpha \left\| P_{ E^{\bot}}u \right\|_2^2\end{eqnarray*}
and then
\begin{eqnarray}\sum_{m=1}^{k}\left| \langle u,V_m\rangle \right|^2 -\alpha
&=&\sum_{m=1}^{k} \left[
\left| \langle u,V_m \rangle \right|^2 - \left| \langle P_{E} u,V_m\rangle \right|^2 \right] \nonumber\\
&&
+\sum_{m=1}^{k} \left[
\left| \langle P_{E} u,V_m\rangle \right|^2
-\alpha \left| \langle u, \xi_m\rangle \right|^2 \right] \nonumber\\
&& -\alpha \left\| P_{ E^{\bot}}u \right\|_2^2 \label{de}
.\end{eqnarray}
Using (\ref{de}),
(\ref{proj1}), (\ref{proj}), (\ref{Delta}) and (\ref{nabla}), we immediately get that
\begin{eqnarray}\left| \sum_{m=1}^{k}\left| \langle u,V_m\rangle \right|^2 -\alpha \right| &\leq &
2 k \left\| P_{ E^{\bot}} u \right\|_2 \nonumber\\
&&+\alpha\left\| P_{ E^{\bot}}u \right\|_2^2 \nonumber\\
&&+\max_{m=1}^{k} \left| \alpha_m-\alpha \right| \nonumber\\
&&+ a_{N},\label{estimgen}\end{eqnarray}
\noindent with\\
~~
$a_N =\sum_{m=1}^{k} \left| \left|\langle \xi_m,V_m\rangle \right|^2-\alpha_m
\right| $ $$ + 2 \sum_{\begin{array}{ll}m,n=1\\n \neq m\end{array} }^{k} \left|\langle \xi_n,V_m\rangle \right|+
\sum_{m=1}^{k}\left(\sum_{\begin{array}{ll}n=1\\n \neq m\end{array} }^{k} \left|\langle \xi_n,V_m\rangle \right|\right)^2.$$
\noindent Now, if (\ref{hypconv}) is satisfied, that is, for each $m,n$ in $\{1,\ldots,k\}^2$,
when $N$ goes to infinity,
$$\left| \langle V_m,\xi_n\rangle \right|^2 {\rightarrow} \delta_{m,n}\alpha_m,$$ it is clear that
$a_N$ converges towards zero when $N$ goes to infinity.
Lemma \ref{Proj} follows. $\Box$
\begin{lemme}\label {evProj}
Let $ M_N $ be an Hermitian $N \times N$ matrix. Assume that there is a sequence $(n(N))_{N\geq 0}$ in $\mathbb{N}$ and a fixed positive integer number $k$,
such that for any $l=1,\ldots,k$, $\lambda_{n(N)+l}(M_N)$ converges towards $\rho\in \mathbb{R}$ when $N$ goes to infinity and
there exists $ \delta_0>0$ such that,
for all large N, \begin{equation}\label{deviate}\lambda_{n(N)}(M_N) > \rho + \delta_0 \, \mbox{~and~} \, \lambda_{n(N)+k+1}(M_N) < \rho - \delta_0
\end{equation}
(with the conventions that $\lambda_0(M_N)=+\infty$ and $\lambda_{N+1}(M_N)=-\infty$).
For any $0<\epsilon<\epsilon_0$, let $M_N(\epsilon)$ be an Hermitian $N \times N$ matrix. Assume that
there exists $f_\epsilon \geq 0$, independent of $N$, decreasing to zero when $\epsilon$ decreases to zero,
such that for all large $N$, for any $0<\epsilon<\epsilon_0$,
\begin{equation}\label{nvelledif}\Vert M_N (\epsilon) - M\Vert \leq f_\epsilon.\end{equation}
Let $0<\tilde{\epsilon}_0<\epsilon_0$ be such that $f_{\tilde{\epsilon}_0}<\frac{\delta_0}{4}$.
Then for all large $N$ , for any $0<\epsilon< \tilde{\epsilon}_0$, for any $l=1,\ldots,k$,
\begin{equation}\label{convergence}\left| \lambda_{n(N)+l}(M_N)(\epsilon) -\rho\right| \leq \frac{\delta_0}{2},\end{equation}
\begin{equation}\label{sep}\lambda_{n(N)}(M_N(\epsilon)) > \rho+ \frac{\delta_0}{2}\mbox{~~and~~}
\lambda_{n(N)+k +1}(M_N(\epsilon)) < \rho - \frac{\delta_0}{2},
\end{equation}
\noindent and for any normalized eigenvector $\xi$ of $M_N$ relative to the eigenvalue $\lambda_{n(N)+l}(M_N)$ for some $l$ in $\{1,\ldots,k\}$,
$$\left\| P_{ E(\epsilon)^{\bot}}\xi\right\|_2 \leq \frac{2}{\delta_0} \left\{ f_\epsilon + \left| \lambda_{n(N)+l}(M_N)-\rho
\right|\right\},$$
\noindent where $E(\epsilon)$ denotes the vector subspace generated by the eigenvectors relative to the eigenvalues $\lambda_{n(N)+q}(M_N(\epsilon))$, $q=1,\ldots,k$.
\end{lemme}
\noindent {\bf Proof:}
\noindent According to Weyl inequalities (see Lemma \ref{Weyl} in the Appendix) and (\ref{nvelledif}),
for all large $N$, for all $0<\epsilon<\epsilon_0$,
$$\lambda_{n(N)+l}(M_N(\epsilon))\leq \lambda_{n(N)+l}(M_N) +f_\epsilon,\mbox{~~for~~} l=1, \ldots, k+1,$$
$$\lambda_{n(N)+l}(M_N(\epsilon))\geq \lambda_{n(N)+l}(M_N) -f_\epsilon, \mbox{~~for~~} l=0, \ldots, k.$$
\noindent
By assumptions of the lemma,
for all large $N$, for any $l=1,\ldots,k$,
$$\left| \lambda_{n(N)+l}(M_N) -\rho\right| \leq \frac{\delta_0}{4},$$
and
\noindent for all large N,
$$\lambda_{n(N)}(M_N) > \rho + \delta_0 \, \mbox{~and~} \, \lambda_{n(N)+k +1}(M_N) < \rho - \delta_0 .$$
Hence, choosing $0<\tilde{\epsilon}_0<\epsilon_0$ such that $f_{\tilde{\epsilon}_0}<\frac{\delta_0}{4}$, (\ref{sep}) and (\ref{convergence}) readily follow.
For all large $N$ and any $0<\epsilon<\tilde{\epsilon}_0$,
let $\xi_1(\epsilon),\ldots, \xi_{k}(\epsilon)$ be an orthonormal basis of $ E(\epsilon)$ such that there exists an $N \times N$ unitary matrix $V(\epsilon)$ whose $k$ first columns are $\xi_1(\epsilon),\ldots, \xi_{k}(\epsilon)$ and a $(N -k)\times (N-k)$ diagonal matrix $\Lambda_2(\epsilon)$ such that $$M_N(\epsilon)=V(\epsilon)\begin{pmatrix} \Lambda_1(\epsilon) & (0)\\(0)&\Lambda_2(\epsilon)\end{pmatrix}V(\epsilon)^*$$
where $$\Lambda_1(\epsilon)=\mbox{diag}\left(\lambda_{n(N)+1}(M_N(\epsilon)), \ldots, \lambda_{n(N)+k}(M_N(\epsilon))\right).$$
\noindent Let $\xi$ be a normalized eigenvector of $M_N$ relative to the eigenvalue $\lambda_{n(N)+l}(M_N)$ for some $l$ in $\{1,\ldots,k\}$.
Let us set $$R(\epsilon):=\left( M_N(\epsilon)- \rho I_N\right)\xi.$$
We have
$$R(\epsilon)=V(\epsilon) \begin{pmatrix}\Lambda_1(\epsilon)- \rho I_{k}& (0)\\(0)&\Lambda_2(\epsilon)- \rho I_{N-k}\end{pmatrix} V(\epsilon)^*\xi.$$
Define the vector $v_1(\epsilon)$ in $\mathbb{C}^{k}$ and the vector $v_2(\epsilon)$ in $\mathbb{C}^{N-k}$
by setting
$$V(\epsilon)^*\xi=\left( \begin{array}{ll} v_1(\epsilon)\\ v_2(\epsilon)\end{array}\right).$$
Note that
\begin{equation}\label{identite}\left\| v_2(\epsilon)\right\|_2=\left\| P_{ E(\epsilon)^{\bot}} \xi\right\|_2.\end{equation}
\noindent According to (\ref{sep}), for all large $N$, for any $0<\epsilon < \tilde{\epsilon}_0$,
for all $i\notin \{n(N)+1, \ldots, n(N)+k\}$, $\left| \lambda_i(M_N(\epsilon))- \rho\right| >
\frac{\delta_0}{2} >0$ and therefore
$\rho$ is not an eigenvalue of $\Lambda_2(\epsilon)$.
Then
$$ v_2(\epsilon)= \left(\Lambda_2(\epsilon)- \rho I_{N-k}\right)^{-1} [V^*(\epsilon)R(\epsilon)]_{(N-k)\times 1}$$
\noindent where $[V^*(\epsilon)R(\epsilon)]_{(N-k)\times 1}$ denotes the vector obtained
from $V^*(\epsilon)R(\epsilon)$ after removing the first $k$ components.
Hence \begin{equation}\label{v2}\left\| v_2(\epsilon) \right\|_2 \leq \frac{2}{\delta_0} \left\| R(\epsilon) \right\|_2 .\end{equation}
Now, we have
\begin{eqnarray}
\left\| R(\epsilon)\right\|_2 &= &\left\| \left( M_N(\epsilon)-M_N
+M_N -\rho I_N\right)\xi\right\|_2 \nonumber\\
&\leq & \left\| M_N(\epsilon)-M_N \right\|+ \left| \lambda_{n(N)+l}(M_N)-\rho\right| \label{normeR}
\end{eqnarray}
\noindent Lemma \ref{evProj} readily follows from (\ref{identite}), (\ref{v2}), (\ref{normeR}) and (\ref{nvelledif}). $\Box$\\
~~
Let us define the continuous function $\tau$ on ${\cal O}^{(a)}$ respectively
${\cal O}^{(m)}\cap ]0;+\infty[$
by setting
$$\tau(x)=\left\{\begin{array}{ll} H'(x) \mbox{~if~} M_N=M_N^W,\\
- \frac{{\cal Z}'(\frac{1}{x})}{x {\cal Z}(\frac{1}{x}) } \mbox{~if~} M_N=M_N^S.
\end{array}\right.$$
Assume that $\theta_j $ in $ \Theta_o$ is such that $k_j\neq 1$.
Let us denote by $V_1(j),\ldots, V_{k_j}(j)$, an orthonormal system of eigenvectors of $A_N$ associated with $\theta_j$.
There exists an $N \times N$ unitary matrix $U$ whose $k_j$ first columns are $ V_1(j),\ldots, V_{k_j}(j)$ and a $(N -k_j)\times (N-k_j)$ Hermitian matrix $D$ such that $$A_N=U \begin{pmatrix}\theta_j I_{k_j}& (0) \\ (0)& D \end{pmatrix} U^*.$$
Let us fix $\epsilon_0$ such that $0< \epsilon_0 < \frac{1}{k_j} \min_{s=1}^{J}dist(\theta_s, \mbox{supp~}\nu \cup_{i\neq s}\theta_i )$ and $[\theta_j; \theta_j +k_j\epsilon_0]\subset {\cal O}$. For any $0<\epsilon < \epsilon_0$,
let us consider $$M_N(\epsilon)=\left\{\begin{array}{ll}X_N + A_N(\epsilon)\mbox{~if~} X_N=X_N^W,\\
A_N(\epsilon)^{\frac{1}{2}}X_N A_N(\epsilon)^{\frac{1}{2}} \mbox{~if~} X_N=X_N^S,\end{array}\right.$$
where $$A_N(\epsilon)=U
\begin{pmatrix}\mbox{diag}(\theta_j + k_j \epsilon, \ldots, \theta_j+ 2 \epsilon, \theta_j + \epsilon)& (0)\\(0)&D
\end{pmatrix}U^*.$$
\noindent Of course for any $0<\epsilon < \epsilon_0$, the limiting spectral distribution of $A_N(\epsilon)$, when $N$ goes to infinity, is the same as the limiting spectral distribution of $A_N$.
Moreover, for all large $N$, the descending ranks $n_{j-1}+1, \ldots , n_{j-1}+k_j$ of $\theta _j$ among the eigenvalues of $A_N$
are the ranks of $\theta_j + k_j \epsilon, \ldots, \theta_j+ 2 \epsilon,\theta_j + \epsilon$ among the eigenvalues of $A_N(\epsilon)$.
For each $m$ in $\{1,\ldots,k_j\}$, $V_m(j)$ is an eigenvector of $A_N(\epsilon)$ associated with the
eigenvalue $\theta_j+ (k_j- m+1)\epsilon$ which is of multiplicity one. Note that, since $A_N$ satisfies Assumption A, there exists a constant $C^{'}$ such that for any $0\leq \epsilon < \epsilon_0$,
$sup_N \left\| A_N(\epsilon) \right\| \leq C^{'}$. It is easy to see that
$$\left\| M_N(\epsilon)-M_N\right\| \leq \left\{\begin{array}{ll} \epsilon ~ k_j \mbox{~if~} X_N=X_N^W,\\
(C^{'})^{\frac{1}{2}}\frac{k_j}{\sqrt{\theta_j}} \left\| X_N \right\|~ \epsilon \mbox{~if~} X_N=X_N^S.\end{array}\right.$$
According to Theorem 5.11 in \cite{BaiSil06}, $\left\| X_N^S\right\| =c(1+ \frac{1}{\sqrt{c}})^2+o_{a.s,N}(1)$.
Thus in both cases, there exists some constant $C_1$ such that a.s for all large N,
for any $0< \epsilon< \epsilon_0$,
\begin{equation}\label{normediff} \left\| M_N(\epsilon)-M_N\right\| \leq C_1 \epsilon.\end{equation}
Let $\xi$ be a normalized eigenvector of $M_N$ relative to $\lambda_{n_{j-1}+q}(M_N)$ for some $q$ in $\{1,\ldots, k_j\}$ . Let
$0<\tilde{\epsilon}_0<\epsilon_0$ be chosen such that $C_1 {\tilde{\epsilon}_0}<\frac{\delta_0}{4}$ where $\delta_0$ is defined in Theorem \ref{ThmASCV}.
Using Theorem \ref{ThmASCV} and (\ref{normediff}), according to Lemma \ref{evProj}, almost surely for all large $N$,
for any $0<\epsilon < \tilde{\epsilon}_0$, the set $\{
\lambda_{n_{j-1}+n}(M_N(\epsilon)), n\in \{1,\ldots, k_j\}\}$ is distinct from the set $\{ \lambda_i(M_N(\epsilon)), i\notin \{n_{j-1}+1, \ldots, n_{j-1}+k_j\}\}$ and
\begin{equation}\label{least2}\left\| P_{ E(\epsilon)^{\bot}}\xi\right\|_2 \leq \frac{2}{\delta_0} \left\{C_1 \epsilon + \left| \lambda_{n_{j-1}+q}(M_N)-\rho_{\theta_j}
\right|\right\},\end{equation}
where $E(\epsilon)$ denotes the vector subspace generated by the eigenvectors relative to the eigenvalues $\lambda_{n_{j-1}+n}(M_N(\epsilon))$, $n=1,\ldots,k_j$.
Define \begin{equation}\label{eta}\iota(\epsilon)= \frac{2 (2k_j+\tau(\theta_j))}{\delta_0} C_1 \epsilon + \max_{m=1}^{k_j} \left| \tau(\theta_j + m\epsilon)-\tau(\theta_j)\right|.\end{equation}
\noindent For any $\zeta >0$, choose and fix $0<\epsilon= \epsilon_1<\tilde{\epsilon}_0$ such that \begin{equation} \label{iota}0<\iota(\epsilon_1)< \frac{\zeta}{2}\end{equation}
(using the continuity of the function $\tau$ at the point $\theta_j$). \\
~~
\noindent Now, each $\theta_j+l \epsilon_1$, $l\in \{1,\ldots, k_j\}$, is a spike of $A_N(\epsilon_1)$ with multiplicity one.
According to Theorem \ref{ThmASCV}, for $n\in \{1,\ldots, k_j\}$,
$\lambda_{n_{j-1}+n}(M_N(\epsilon_1))$ asymptotically separates from the rest of the spectrum and converges almost surely towards
$H(\theta_j +(k_j-n+1) \epsilon_1)$; moreover, if $\xi_n(\epsilon_1,j)$ denotes a normalized eigenvector associated to $\lambda_{n_{j-1}+n}(M_N(\epsilon_1))$,
Proposition \ref{cvev}
implies that
$$\left| \langle V_m(j),\xi_n(\epsilon_1,j)\rangle \right|^2 \stackrel{a.s}{\rightarrow} \delta_{m,n}\tau(\theta_j+(k_j-m+1)\epsilon_1).$$
According to Lemma \ref{Proj}, there exists a random variable $a_N(\epsilon_1) \geq 0$, converging almost surely to zero when $N$ goes to infinity,
such that, almost surely, for all large $N$,
\begin{eqnarray*}\left| \left\| P_{\mbox{Ker~}(\theta_j I_N-A_N)} \xi \right\|_2^2-\tau(\theta_j)\right| \leq& (2k_j+\tau(\theta_j)) \left\|P_{E(\epsilon_1)^{\bot}} \xi \right\|_2 +a_N(\epsilon_1)\nonumber \\
&+\max_{m=1}^{k_j} \left| \tau(\theta_j + m\epsilon_1)-\tau(\theta_j)\right|.\end{eqnarray*}
The last inequality, (\ref{least2}) and (\ref{eta}) readily yield that, almost surely, for all large $N$,
\begin{eqnarray*}\label{least}\left| \left\| P_{\mbox{Ker~}(\theta_j I_N-A_N)} \xi \right\|_2^2-\tau(\theta_j)\right| &\leq& \frac{2}{\delta_0}(2k_j+\tau(\theta_j)) \left| \lambda_{n_{j-1}+q}(M_N)-\rho_{\theta_j}
\right| \\ && +a_N(\epsilon_1) +\iota(\epsilon_1).\end{eqnarray*}
\noindent Therefore, (\ref{iota}), the almost surely convergence of $ \lambda_{n_{j-1}+q}(M_N)$ towards $\rho_{\theta_j}$ and of $a_N(\epsilon_1)$
towards zero imply that,
almost surely, for all large $N$,
$$\left|\left\| P_{\mbox{Ker~}(\theta_j I_N-A_N)} \xi \right\|^2_2 -\tau(\theta_j) \right|
\leq
\zeta $$
\noindent and the proof is complete.\hspace*{8cm} $\Box$
\section{Proof of Proposition \ref{cvev}}
Since the proof of Proposition \ref{cvev} is exactly the same for the deformed Wigner model and the sample covariance matrix, in order to present a unified approach of this proof, we adopt in this section the notations of the first column of the array (\ref{correspondance}) standing for both the corresponding elements of the second column (deformed Wigner matrix case) and the third column (sample covariance matrix case). We postpone in a later subsection the technical results that need a specific study for each model in order to not lose the thread of this common proof.
\subsection{Restriction to the asymptotic behavior of some expectation $\mathbb {E}{\rm Tr} \left[h(M_N) f(A_N)\right]$}
The aim of this first step of the proof of Proposition \ref{cvev} is to reduce the study of the asymptotic behaviour of $\sum_{n=1}^{k_j}\left\| P_{\mbox{Ker~}(\theta_l I_N-A_N)} \xi_n(j) \right\|^2_2$ to the one of the expectation $\mathbb {E}{\rm Tr} \left[h(M_N) f(A_N)\right]$ for some functions $f$ and $h$ respectively concentrated on a neighborhood of $\theta_l$ and $\rho_{\theta_j}$. We will use the convergence results on eigenvalues described in Theorem \ref{cvev} above and concentration inequalities presented in the Appendix.\\
P. Biane already suggested in \cite{PB} to evaluate the moduli of the Hermitian inner products of the eigenvectors on test functions.
Indeed, for any smooth function $h$ and $f$ on $\mathbb{R}$, denoted by $u_1,\ldots, u_N$ (resp. $w_1,\ldots, w_N$), the eigenvectors associated with $\lambda_1(A_N), \ldots, \lambda_N(A_N)$ (resp. $\lambda_1(M_N), \ldots, \lambda_N(M_N)$), one can easily check that
$${\rm Tr} \left[h(M_N) f(A_N)\right] =\sum_{k,i} h(\lambda_k(M_N)) f(\lambda_i(A_N)) \vert \langle u_i,w_k \rangle \vert^2.$$
Thus, since $\theta_l$ on one hand and the $\lambda_{n_{j-1}+n}(M_N)$, $n=1,\ldots,k_j$, on the other hand, asymptotically separate from the rest of the spectrum of respectively $A_N$ and $M_N$,
a fit choice of $h$ and $f$ will allow the study of the restrictive sum $\sum_{n=1}^{k_j}\left\| P_{\mbox{Ker~}(\theta_l I_N-A_N)} \xi_n(j) \right\|^2_2$.\\
\noindent Let us fix $$0< \eta < \frac{1}{2} \min_{s=1}^{J}dist(\theta_s, \mbox{supp~}\nu \cup_{i\neq s}\theta_i )$$ and
for any $l=1,\ldots,J$,
choose $f_{\eta,l}$ in ${\cal C}^\infty (\mathbb{R}, \mathbb{R})$ with support in
$[\theta_l -\eta,\theta_l+\eta]$ such that $f_{\eta,l}(\theta_l)=1$ and $0 \leq f_{\eta,l} \leq 1$.\\
~\\
For any $\theta_i \in {\cal O}^{(a)}$, according to Remark \ref{oainclu}, $\theta_i$ belongs to $\overline{\Omega_{\sigma,\nu}}$ and according to Proposition \ref{homeo}, $F_{\sigma,\nu}^{(a)}(\rho_{\theta_i})=\theta_i$.
For any $\theta_i \in {\cal O}^{(m)}\cap ]0;+\infty[$, according to Remark \ref{nonnul}, $\rho_{\theta_i}= {\cal Z}(\frac{1}{\theta_i})>0$
and according to Remark \ref{CScomp}, $g_{\tau_{c,\nu}}(\rho_{\theta_i}) =\frac{1}{\theta_i}$ so that
$\frac{1}{F_{c,\nu}^{(m)}(\frac{1}{\rho_{\theta_i}})} =\theta_i$.\\
According to Theorem \ref{ThmASCV}, there exists
$\delta_0>0$ such that almost surely for all large N, for all $\theta_j$ in $\Theta_o$,
$$\lambda_{n_{j-1}}(M_N) > \rho _{\theta _j} + \delta_0 \, \mbox{~and~} \, \lambda_{n_{j-1}+k_j +1}(M_N) < \rho _{\theta _j} - \delta_0 .$$
\noindent Let us fix $$0< \delta < \frac{1}{3}\min \{\delta_0, \mbox{~dist~}(\rho_{\theta_s}, {\cal S}), \vert \rho_{\theta_s}- \rho_{\theta_{t}}\vert, s\neq t,(\theta_s,\theta_t) \in \Theta_o^2\},$$
where $${\cal S}= \left\{ \begin{array}{ll} \mbox{supp~}(\mu\boxplus \nu) \mbox{~if ~} M_N =M_N^W\\
\mbox{supp~}(\tau_{c,\nu})\cup \{0\}\mbox{~if ~} M_N =M_N^S \end{array} \right. .$$
\noindent For any $j$ such that $\theta_j \in \Theta_o$, choose $h_{\delta,j}$ in ${\cal C}^\infty (\mathbb{R}, \mathbb{R})$ with support in
$[\rho_{\theta_j} -\delta,\rho_{\theta_j}+\delta]$ such that $h_{\delta,j} \equiv 1$ on $[\rho_{\theta_j} -\frac{\delta}{2},\rho_{\theta_j}+\frac{\delta}{2}]$ and $0 \leq h_{\delta,j}\leq 1$.
\begin{lemme} \label{vecttrace}
When $N$ goes to infinity $$Tr \left[h_{\delta,j}(M_N) f_{\eta,l}(A_N)\right]-\sum_{n=1}^{k_j}\left\| P_{\mbox{Ker~}(\theta_l I_N-A_N)} \xi_n(j) \right\|^2_2 \stackrel{a.s}{\rightarrow}0.$$
\end{lemme}
\noindent {\bf Proof:} According to Theorem \ref{ThmASCV},
there exists some set $\Omega$ of probability one
such that on $\Omega$, for all large $N$, $\forall i=1,\ldots,k_j,$
$$
\vert \lambda_{n_{j-1}+i}(M_N) -\rho _{\theta _j}\vert < \frac{\delta}{2} $$
$$ \lambda_{n_{j-1}}(M_N)\geq \rho_{\theta_j} + \delta, \, \lambda_{n_{j-1}+ k_j+1}(M_N) \leq \rho_{\theta_j} - \delta .$$
Using also the assumption (\ref{univconv}) on the $\beta_i(N)$'s, we have that on $\Omega$, for all large $N$, $$\sum_{n=1}^{k_j}\left\| P_{\mbox{Ker~}(\theta_l I_N-A_N)} \xi_n(j) \right\|^2_2 = Tr \left[h_{\delta,j}(M_N) f_{\eta,l}(A_N)\right].$$
Hence,
Lemma \ref{vecttrace} follows. $\Box$\\
Now, according to Lemma \ref{Herbst}, Remark \ref{multiple} and Lemma \ref{Lipschitz}, the random variables $F^W_N={\rm Tr} \left[ h_{\delta,j} (M_N^W) f_{\eta,l}(A_N)\right]$
and $F^S_N={\rm Tr} \left[ h_{\delta,j} (M_N^S) f_{\eta,l}(A_N)\right]$
satisfy respectively the following concentration inequalities
$$\forall \epsilon> 0, \, \mathbb{P}\left( \vert F^W_N-\mathbb {E}(F^W_N) \vert > \epsilon \right) \leq K_1 \exp\left(-\frac{\epsilon\sqrt{ N}}{K_2 \sqrt{ C_{PI} k_l } \Vert h_{\delta,j} \Vert_{Lip}}\right).$$
$$\forall \epsilon> 0, \, \mathbb{P}\left( \vert F^S_N-\mathbb {E}(F^S_N) \vert > \epsilon \right) \leq K_1 \exp\left(-\frac{\epsilon \sqrt{ p}}{K_2 \sqrt{C C_{PI} k_l } \Vert \tilde{h}_{\delta,j} \Vert_{Lip}}\right).$$
\noindent (The constants have been introduced in Lemma \ref{Herbst} and Lemma \ref{Lipschitz}).
By Borel-Cantelli Lemma, we can readily deduce the following lemma.
\begin{lemme}\label{conc}
$${\rm Tr} \left[ h_{\delta,j} (M_N) f_{\eta,l}(A_N)\right]- \mathbb{E}\left[ {\rm Tr} \left[ h_{\delta,j} (M_N) f_{\eta,l}(A_N)\right] \right]\stackrel{a.s}{\rightarrow}0.$$
\end{lemme}
Lemma \ref{vecttrace} and Lemma \ref{conc} allow us to conclude this first step of the proof by the following result.
\begin{proposition}\label{premetap}
For any $\theta_j$ in $\Theta_o$ and any $\theta_l$ in $\Theta$, when $N$ goes to infinity
$$\sum_{n=1}^{k_j}\left\| P_{\mbox{Ker~}(\theta_l I_N-A_N)} \xi_n(j) \right\|^2_2 -\mathbb{E}\left[ {\rm Tr} \left[ h_{\delta,j} (M_N) f_{\eta,l}(A_N)\right] \right]\stackrel{a.s}{\rightarrow}0.$$
\end{proposition}
\subsection{Making use of estimations of the resolvent}
The basic idea of this second step of the proof of Proposition \ref{cvev} is to approximate the function $h_{\delta,j}$ by its convolution by the Poisson Kernel in order to exhibit the resolvent of the deformed model and then use sharp estimations on this resolvent.
\begin{lemme}\label{approxcauchy}
For any continuous function $h$ with compact support and any bounded continuous function $\phi$,
$$\mathbb{E}\left[ {\rm Tr} \left[ h (M_N) \phi (A_N)\right] \right]=- \lim_{y\rightarrow 0^{+}}\frac{1}{\pi} \Im \int \mathbb{E} \left( {\rm Tr} \left[G_{N}(t+iy)\phi (A_N)\right] \right)h(t) dt $$
\noindent where $G_N(z) =(zI_N -M_N)^{-1}$.
\end{lemme}
\noindent {\bf Proof}
Let us denote by $P_y$ the Poisson kernel $$P_y(x) =\frac{y}{\pi (x^2 +y ^2)}, \, x\in \mathbb{R}, \, y>0.$$
We have
\begin{eqnarray*}h(x)&= & \lim_{y\rightarrow 0^{+}} h * P_y(x)\\
&=&\lim_{y\rightarrow 0^{+}}\frac{1}{\pi} \int \frac{y h(t)}{(x-t)^2 +y^2} dt \\
&=& \lim_{y\rightarrow 0^{+}}\frac{1}{\pi} \int \Im\frac{h(t)}{x-iy-t} dt.
\end{eqnarray*}
Thus, for any fixed $N$, $$h(M_N) = - \lim_{y\rightarrow 0^{+}}\frac{1}{\pi} \int \Im G_{N}(t+iy) h(t) dt, $$
and since $\Vert h * P_y \Vert_\infty \leq \Vert h \Vert_\infty$ we have
$\Vert \int \Im G_{N}(t+iy) h(t) dt\Vert \leq \Vert h \Vert_\infty$.
Then, the result readily follows by dominated convergence Theorem and Fubini's Theorem. $\Box$\\
\noindent
Let $U$ be a unitary matrix and $$D={\rm diag}(\gamma_1, \ldots ,\gamma_N )$$
such that
$$A_N= U^*DU$$
Let $G$ stand for $G_N$ and $g$ stand for $g_{\mu_{\mbox{\tiny{LSD}}}}$.
Consider $\tilde{ G }= UGU^*$.
For any continuous function $\phi$, \begin{equation}\label{GtildeA}{\rm Tr} \left[G_{M_N}(z)\phi (A_N)\right]= \sum_{k=1}^N \phi (\gamma_k)\tilde{G}_{kk}(z).\end{equation}
The following result is fundamental in our approach.
\begin{proposition}\label{estimfonda}
There is a polynomial $P$ with nonnegative coefficients, a sequence $a_N$ of nonnegative
real numbers converging to zero when $N$ goes to infinity
and some nonnegative integer number $\alpha$,
such that for any $k$ in $\{1, \ldots,N\}$, for all $z\in \mathbb{C}\setminus \mathbb{R}$,
\begin{equation}
\mathbb {E} (\tilde G_{kk}(z)) = \tilde{\Phi}_k(z)
+\Delta_{k,N}(z),
\end{equation}
with $$\left| \Delta_{k,N} (z)\right| \leq (1+\vert z\vert)^\alpha P(\vert \Im z \vert^{-1})a_N,$$
where $$\tilde{\Phi}_k(z)=\left\{ \begin{array}{ll}
\frac{1}{z-\sigma^2 g(z)-\gamma_k}=\frac{1}{F^{(a)}_{\sigma,\nu}(z)-\gamma_k}\, \hspace*{1cm} \mbox{~if~} \, M_N=M_N^W,\\
\frac{1}{z-\gamma_k(1-c+czg(z))}=\frac{1}{z ( 1 -\gamma_k F^{(m)}_{c,\nu}(\frac{1}{z}))}\, \mbox{~if~} \, M_N=M_N^S.\end{array} \right.$$
\end{proposition}
Note that $\tilde{\Phi}_k(z)$ is well defined for any $z\in \mathbb{C}\setminus \mathbb{R}$ since
\begin{equation}\label{majim}\vert \Im[z-\sigma^2 g(z)-\gamma_k]\vert = \vert \Im(z)\vert \left[1+ \sigma^2\int \frac{1}{\vert z-x\vert^2}d\mu_{\sigma}\boxplus \nu(x)\right]\geq \vert \Im (z)\vert >0,\end{equation}
\begin{equation}\label{majim2}\vert \Im[z-\gamma_k(1-c+czg(z))]\vert = \vert \Im(z)\vert \left[1+ c\gamma_k\int \frac{x}{\vert z-x\vert^2}d (\mu_{\mbox{\tiny{MP}}, {c}}\boxtimes \nu)(x)\right]\geq \vert \Im (z)\vert >0.\end{equation}
According to (\ref{estimfondakW}) and Proposition \ref{estimresolvwish},
there exists a polynomial $P$ with nonnegative coefficients and a sequence $a_N$ of nonnegative
real numbers converging to zero when $N$ goes to infinity such that, for any $k$ in $\{1, \ldots,N\}$, for any $z\in \mathbb{C}\setminus \mathbb{R}$ \begin{equation} \label{estimfondaN}
\mathbb {E} (\tilde G_{kk}(z)) = \tilde{\Phi}_{N,k}(z)
+R_{k,N}(z),
\end{equation}
with $$\vert R_{k,N}(z) \vert \leq (1+\vert z\vert)^2 P(\vert \Im z \vert^{-1})a_N,$$
where $$\tilde{\Phi}_{N,k}(z)=\left\{ \begin{array}{ll}
\frac{1}{z-\sigma^2 g_N^W(z)-\gamma_k}, \hspace*{1cm} \mbox{~if~} \, M_N=M_N^W,\\
\frac{1}{z-\gamma_k(1-\frac{N}{p}+\frac{N}{p}zg_N^S(z))}=\, \mbox{~if~} \, M_N=M_N^S.\end{array} \right.$$
In order to deduce Proposition \ref{estimfonda}, we will need the following description of the convergence of $g_N(z)$ towards $g(z)$.
\begin{proposition}\label{estimdif}
There exists a polynomial $R$ with nonnegative coefficients, a sequence $a_N$ of positive numbers converging towards zero and some nonnegative integer number $\alpha$ such that, for all $z \in \mathbb{C}\setminus \mathbb{R}$,
\begin{equation}\label{dif}\vert {g}_{N}(z) -g(z)\vert \leq (1+ \vert z \vert)^\alpha R(\vert \Im z \vert^{-1}) a_N.\end{equation}
\end{proposition}
\noindent{\bf Proof}:
1) The deformed Wigner model case.\\
Denote by $\tilde{g}_{N}$ the Stieltjes transform of $\mu_\sigma \boxplus \mu_{A_N}$.
Since we have already proved in \cite{CDFF10} that there exists a polynomial $S$ with nonnegative coefficients such that
for all $z \in \mathbb{C}\setminus \mathbb{R}$, \begin{equation}\vert \tilde{g}_{N}(z) -g_N(z)\vert \leq \frac{S(\vert \Im z \vert^{-1})}{N},\end{equation}
the result will readily follow if we prove that
there exists a polynomial $T$ with nonnegative coefficients and a sequence $b_N$ of positive numbers converging towards zero such that, for all $z \in \mathbb{C}^+$,
\begin{equation}\label{dif2}\vert \tilde{g}_{N}(z) -g(z)\vert \leq T(\vert \Im z \vert^{-1}) b_N.\end{equation}
The proof of (\ref{dif2}) follows the lines of Section 4 in \cite{CDFF10}.
For a fixed $z\in \mathbb {C}^+$, according to Proposition \ref{homeo}, we have the subordination equations:
\begin{equation} \label{eqtilde} \tilde g_N(z) = g_{\mu _{A_N}}(F^{(a)}_{\sigma,\mu _{A_N}}(z))=g_{\mu _{A_N}}(z - \sigma ^2\tilde g_N(z)),\end{equation}
\begin{equation} \label{eq} g(z) = g_{\nu}(F^{(a)}_{\sigma,\nu}(z))=g_{\nu}(z - \sigma ^2 g(z)).\end{equation}
Moreover, using Lemma \ref{cvAN} and $\Im (z - \sigma ^2g(z)) \geq \Im z $ ,
we deduce from (\ref{eq}) that
\begin{equation} \label{appsub} g(z) = g_{\mu _{A_N}}(z - \sigma ^2g(z)) + \Delta_N(z) \end{equation}
\noindent with
$$\vert \Delta_N (z)\vert \leq v_N(1) P_1(|\Im z|^{-1}),$$
\noindent where $P_1$ is a polynomial with nonnegative coefficients and $v_N(1)$ is a sequence of positive numbers converging towards zero. \\
Since $z - \sigma ^2g(z)\in \mathbb {C}^+$, $z'\in \mathbb {C}$ is
well-defined by the formula :
$$z':=H_{\sigma , \mu _{A_N}}(z - \sigma ^2g(z)),$$
where $H_{\sigma , \mu _{A_N}}$ is defined by (\ref{defdeH}) replacing $\nu$ by $\mu_{A_N}$.
One has
\begin{eqnarray*}
\vert z'-z \vert&=&\vert-\sigma ^2(g(z)-g_{\mu _{A_N}}(z - \sigma ^2g(z)))\vert \\
&\leq &\sigma^2 v_N(1) P_1(|\Im z|^{-1}).
\end{eqnarray*}
\begin{itemize}
\item
If $$\frac{|\Im z|}{2} \leq \sigma^2 v_N(1) P_1(|\Im z|^{-1}),$$
or equivalently
\begin{equation} \label{1=O(1/N)}
1\leq {2\sigma^2|\Im z|^{-1}P_1(|\Im z|^{-1})}v_N(1),
\end{equation}
\begin{eqnarray*}
|g(z)-\tilde g_N(z)|&\leq &\frac{2}{|\Im z|}\\
&\leq &{4\sigma^2|\Im z|^{-2}P_1(|\Im z|^{-1})}v_N(1).
\end{eqnarray*}
\item If $$\frac{|\Im z|}{2}> \sigma^2 v_N(1) P_1(|\Im z|^{-1}),$$
one has :
$$ |\Im z'-\Im z|\leq |z'-z|\leq \frac{|\Im z|}{2}$$
which implies $\Im z'\geq \frac{\Im z}{2}$ and
therefore $z'\in \mathbb {C}^+$.
Hence, according to (\ref{imageinverse}), it follows that
$z-\sigma ^2g(z)\in \Omega_{\sigma , \mu _{A_N}}$ (where $ \Omega_{\sigma , \mu _{A_N}}$ is defined by (\ref{Ome}) replacing
$\nu$ by $\mu _{A_N}$)
so that $F^{(a)}_{\sigma , \mu _{A_N}}(z')=z-\sigma ^2g(z)$.
Thus, the approximative equation \eqref{appsub} may be rewritten
$$g(z) =g_{\mu _{A_N}}(F^{(a)}_{\sigma , \mu _{A_N}}(z')) + \Delta_N(z)$$
and then, using the subordination equation \eqref{eqtilde}
$$
g(z)=\tilde g_N(z')+
\Delta_N(z).$$
Moreover,
\begin{eqnarray*}\vert \tilde{g}_N(z')-\tilde{g}_N(z)\vert&=& \vert (z-z')\int_{\mathbb{R}}\frac{d(\mu_\sigma \boxplus \mu_{A_N})(x)}{(z'-x)(z-x)}\vert\\
&\leq & 2\sigma^2 v_N(1) \vert \Im z \vert^{-2} P_1(|\Im z|^{-1}).\end{eqnarray*}
Hence
\begin{eqnarray*}
|g(z)-\tilde g_N(z)|&\leq &|g(z)-\tilde g_N(z')|+|\tilde g_N(z')-\tilde g_N(z)|\\ & \leq & (2\sigma^2 \vert \Im z \vert^{-2}+1)v_N(1) P_1(|\Im z|^{-1})
\end{eqnarray*}
\end{itemize}
Finally we get that for all $z \in \mathbb{C}^+$,
$$|g(z)-\tilde g_N(z)|\leq (4\sigma^2 \vert \Im z \vert^{-2}+1)v_N(1) P_1(|\Im z|^{-1})$$
\noindent so that (\ref{dif}) is satisfied in the deformed Wigner model setting with $a_N=v_N(1)$ and $R(x)=(4\sigma^2 x^{2}+1) P_1(x)$, $\alpha=0$.\\
\noindent 2) The sample covariance matrix setting\\
Let $z$ be in $\mathbb{C}^+$.
Note that it is obviously equivalent to prove such an estimation for $\left|\underline{g}_N(z)- g_{\tau_{c,\nu}}(z)\right|$ where $$\underline{g}_N(z) =\frac{1-\frac{N}{p}}{z} + \frac{N}{p} g_N(z)$$
is the expected value of the Stieltjes transform of the spectral measure of $\underline{M}_N^S=\frac{1}{p} B_N^* A_N B_N$.
In the following any $P_i$ will denote a polynomial with nonnegative coefficients and $a_N(i)$ will denote a sequence of nonnegative numbers converging towards zero when $N$ goes to infinity.
Letting the sum running over $k$ in (\ref{submat}) and dividing by $N$ we have
$${g}_N(z) = \int \frac{d\mu_{A_N}(t)}{z(1- t \underline{g}_N(z))} + \Delta_N(z)$$
\noindent where
$$\vert \Delta_N(z) \vert \leq (1+\vert z\vert)^2 P_1(\vert \Im z \vert^{-1})a_N(1).$$
It readily follows that
$$
\underline{g}_N(z) = \frac{\underline{g}_N(z)}{z} {\cal Z}_{\frac{N}{p},\mu_{A_N}}(\underline{g}_N(z)) + \frac{N}{p}\Delta_N(z)$$
\noindent where ${\cal Z}_{\frac{N}{p},\mu_{A_N}}$ is defined by (\ref{defzcnu}) replacing $\nu$ by $\mu_{A_N}$ and $c$ by
$\frac{N}{p}$.
\noindent Then using Lemma \ref{MajinversegN} we deduce that
\begin{equation}\label{zzprime}{\cal Z}_{ \frac{N}{p},\mu_{A_N}}(\underline{g}_N(z)) =z+ R_N(z)\end{equation}
\noindent where for all large $N$ $$\vert R_N(z) \vert \leq (1+\vert z\vert)^5 P_2(\vert \Im z \vert^{-1})a_N(1)$$
\noindent On the other hand, using (\ref{hypA2}) and Lemma \ref{MajinversegN} we have \begin{equation}\label{zzprime2} {\cal Z}_{\frac{N}{p},\mu_{A_N}}(\underline{g}_N(z)) ={\cal Z}(\underline{g}_N(z)) + Q_N(z)\end{equation}
where $$\vert Q_N(z) \vert \leq (1+\vert z\vert)^q P_3(\vert \Im z \vert^{-1})a_N(2)$$ for some nonnegative integer number $q$.
We readily deduce from (\ref{zzprime}) and (\ref{zzprime2}) that
$${\cal Z}(\underline{g}_N(z)) =z + T_N(z)$$
where $$\vert T_N(z) \vert \leq (1+\vert z\vert)^{\alpha} P_4(\vert \Im z \vert^{-1})a_N(3)$$ for some nonnegative integer number $\alpha$.
\noindent Set $$z^{'}={\cal Z}(\underline{g}_N(z)).$$
\begin{itemize}
\item
If $$\frac{|\Im z|}{2} \leq (1+\vert z\vert)^{\alpha} P_4(\vert \Im z \vert^{-1})a_N(3),$$
or equivalently
$$ 1\leq 2|\Im z|^{-1}(1+\vert z\vert)^{\alpha} P_4(\vert \Im z \vert^{-1})a_N(3),$$
\noindent then
\begin{eqnarray*}
|g_{\tau_{c,\nu}}(z)-\underline{g}_N(z)\vert &\leq &\frac{2}{|\Im z|}\\
&\leq &4|\Im z|^{-2}(1+\vert z\vert)^{\alpha} P_4(\vert \Im z \vert^{-1})a_N(3).
\end{eqnarray*}
\item If $$\frac{|\Im z|}{2}> (1+\vert z\vert)^{\alpha} P_4(\vert \Im z \vert^{-1})a_N(3),$$
one has :
$$ |\Im z'-\Im z|\leq |z'-z|< \frac{|\Im z|}{2}$$
which implies $\Im z'\geq \frac{\Im z}{2}$ and
therefore $z'\in \mathbb{C}^+$.
Note that $\underline{g}_N(z)$ satisfied the equation
$$Z =\frac{1}{z^{'}-c\int \frac{t}{1-tZ}d\nu(t)}$$ and since $g_{\tau_{c,\nu}}(z^{'})$ is the unique solution in
$\mathbb{C}^{-}$ of the latter equation, we can deduce that
$g_{\tau_{c,\nu}}(z^{'})=\underline{g}_N(z)$.
Hence
\begin{eqnarray*}\left|\underline{g}_N(z)- g_{\tau_{c,\nu}}(z)\right|&=& \left|g_{\tau_{c,\nu}}(z^{'})-g_{\tau_{c,\nu}}(z)\right|\\
&\leq & \vert \Im z \vert^{-1} \vert \Im z'\vert^{-1} \vert z - z'\vert \\
&\leq& 2\vert \Im z \vert^{-2}(1+\vert z\vert)^{\alpha} P_4(\vert \Im z \vert^{-1})a_N(3).
\end{eqnarray*}
\end{itemize}
Finally we get that for all $z \in \mathbb{C}^+$,
$$\left|\underline{g}_N(z)- g_{\tau_{c,\nu}}(z)\right|\leq 4\vert \Im z \vert^{-2}(1+\vert z\vert)^{\alpha} P_4(\vert \Im z \vert^{-1})a_N(3)$$
\noindent so that (\ref{dif}) is satisfied in the sample covariance matrix setting .\\
$\Box$\\
Now, Proposition \ref{estimfonda} readily follows from (\ref{estimfondaN}) and Proposition \ref{estimdif}
using (\ref{image}), (\ref{majim2}) and (\ref{imageW}) and (\ref{majim}). $\Box$\\
Thus, for any $l=1,\ldots,J$, (\ref{GtildeA}) and Proposition \ref{estimfonda} yield that
for all large $N$, for all $z \in \mathbb{C} \setminus \mathbb{R}$, \begin{equation}\label{ecrit} \mathbb{E} \left( {\rm Tr} \left[G_{M_N}(z)f_{\eta,l} (A_N)\right] \right)
= \phi_l (z) + \Delta_N(z)\end{equation}
where $\phi_l$ is the following analytic function on $\mathbb{C} \setminus \mathbb{R}$:
$${\phi}_l(z)=\left\{ \begin{array}{ll}
\frac{k_l}{z-\sigma^2 g(z)-\theta_l}=\frac{k_l}{F^{(a)}_{\sigma,\nu}(z)-\theta_l}\, \hspace*{1cm} \mbox{~if~} \, M_N=M_N^W,\\
\frac{k_l}{z-\theta_l(1-c+czg(z))}=\frac{k_l}{z ( 1 -\theta_l F^{(m)}_{c,\nu}(\frac{1}{z}))}\, \mbox{~if~} \, M_N=M_N^S,\end{array} \right.$$
\noindent and
$$ \vert \Delta_N(z) \vert \leq a_N (1 + \vert z \vert )^\alpha P(\frac{1}{\vert \Im z \vert})$$
\noindent for some polynomial $P$ with nonnegative coefficients, some sequence $a_N$ of positive
real numbers converging to zero when $N$ goes to infinity and some nonnegative integer number $\alpha$.
\noindent According to Lemma \ref{HT}, we have
$$\limsup_{y \rightarrow 0^+} ~ (a_N)^{-1}\vert \int h_{\delta,j}(t) \Delta_N(t+iy)dt \vert <+\infty$$
so that \begin{equation}\label{reste}\lim_{N\rightarrow + \infty}\limsup_{y \rightarrow 0^+} \vert \int h_{\delta,j}(t) \Delta_N(t+iy)dt \vert =0.\end{equation}
\begin{lemme}\label{pourholomorphe}
Let $\theta_j $ be in $ \Theta_o$.
\begin{itemize}
\item[(1)] Set $\rho_{\theta_j}=H(\theta_j)$.
If $\theta_l \in \Theta\setminus\{\theta_j\}$, the map
$(x,y) \mapsto F^{(a)}_{\sigma,\nu}(x+iy) -\theta_l$ does not vanish on $[\rho_{\theta_j}-2\delta; \rho_{\theta_j}+2\delta]\times \mathbb{R}$.\\
The only vanishing point in $[\rho_{\theta_j}-2\delta; \rho_{\theta_j}+2\delta]\times \mathbb{R}$ of the map $(x,y) \mapsto F^{(a)}_{\sigma,\nu}(x+iy) -\theta_j$ is $(\rho_{\theta_j},0)$.
\item[(2)] Set $\rho_{\theta_j}={\cal Z}(\frac{1}{\theta_j})$. If $\theta_l \in \Theta\setminus\{\theta_j\}$, the map $(x,y) \mapsto (x+iy) (1- \theta_l g_{\tau_{c,\nu}}(x+iy))$ does not vanish on $[\rho_{\theta_j}-2\delta; \rho_{\theta_j}+2\delta]\times \mathbb{R}$.\\
The only vanishing point in $[\rho_{\theta_j}-2\delta; \rho_{\theta_j}+2\delta]\times \mathbb{R}$ of the map $(x,y) \mapsto (x+iy) (1- \theta_j g_{\tau_{c,\nu}}(x+iy))$ is $(\rho_{\theta_j},0)$.
\end{itemize}
\end{lemme}
\noindent {\bf Proof:}
Note that if $y\neq 0$, for any $x$, the imaginary part of $F^{(a)}_{\sigma,\nu}(x+iy) -\theta_l$ and $1- \theta_l g_{\tau_{c,\nu}}(x+iy)$ is
nonnull
so that we will focus on the case $y=0$.\\
\noindent{ Proof of (1):}
\begin{itemize}
\item Assume $\theta_l \notin \Theta_o$.
First, if $H'(\theta_l)<0$, according to (\ref{imagereelle}),
$\theta_l $ does not belong to $F^{(a)}_{\sigma,\nu}(\mathbb{R})$ so that the conclusion of Lemma \ref{pourholomorphe} (1) is true.\\
Now assume that $H'(\theta_l)=0$. According to (\ref{imagereelle}), $\theta_l\in \partial \Omega_{\sigma,\nu}=F^{(a)}_{\sigma,\nu}(\mathbb{R})$,
and,
by Proposition \ref{homeo}, $F^{(a)}_{\sigma,\nu}(x) -\theta_l=0$ implies $x=H(\theta_l)$.
For any $u \in {\cal O}^{(a)}$,
we have $H(\theta_l)\neq H(u)$. Indeed, for any $u \in {\cal O}^{(a)}$, there exists $u_1, u_2$ such that $[u_1;u_2] \subset {\cal O}^{(a)}$ and $u_1< u < u_2$. Since $H$ is globally nondecreasing on
$\{v \in ^c \mbox{supp~}(\nu), \, H'(v)\geq 0\}$
(see Remark \ref{Hcroit}), we have if $\theta_l <u$,
$H(\theta_l) \leq H(u_1) < H(u)$
and if $\theta_l >u$, $H(u) < H(u_2) \leq H(\theta_l)$.
It follows, according to (\ref{car1}), that $H(\theta_l)$ belongs to supp$(\mu\boxplus \sigma)$ and therefore cannot belong to $[\rho_{\theta_j}-2\delta; \rho_{\theta_j}+2\delta]$ so that the conclusion of Lemma \ref{pourholomorphe} (1) is true.
\item
Let us consider now, $\theta_l \in \Theta_o$. By Remark \ref{oainclu} and Proposition \ref{homeo}, $F^{(a)}_{\sigma,\nu}(x) -\theta_l=0$ implies $x=H(\theta_l)=\rho_{\theta_l}$. If $l\neq j$, $\rho_{\theta_l}$ does not belong to $[\rho_{\theta_j}-2\delta; \rho_{\theta_j}+2\delta]$ and the proof of Lemma \ref{pourholomorphe} (1) is complete.
\end{itemize}
\noindent{Proof of (2):}\\
~~\\
First, note that $0 \notin [\rho_{\theta_j}-2\delta; \rho_{\theta_j}+2\delta].$
\begin{itemize}
\item Assume $\theta_l \notin \Theta_o$.
First,
if ${\cal Z}^{'} (\frac{1}{\theta_l}) >0$, according to Remark \ref{pasdansimagedeg}, $\frac{1}{\theta_l}$ does not belong to $g_{\tau_{c,\nu}}(\mathbb{R}\setminus\{0\})$ so that the conclusion of Lemma \ref{pourholomorphe} (2) is true.\\
Now assume that ${\cal Z}^{'} (\frac{1}{\theta_l}) =0$.
By Remark \ref{compositionlimite}, $1- \theta_l g_{\tau_{c,\nu}}(x)=0$ with $x$ nonnull implies $x={\cal Z}(\frac{1}{\theta_l}).$
In particular, if ${\cal Z}(\frac{1}{\theta_l})=0$, the conclusion of Lemma \ref{pourholomorphe} (2) is true since $0 \notin
[\rho_{\theta_j}-2\delta; \rho_{\theta_j}+2\delta].$ Hence, in the following, we will deal
with $\theta_l$ such that ${\cal Z}(\frac{1}{\theta_l})\neq 0$.
For any $u \in {\cal O}^{(m)}$,
we have ${\cal Z}(\frac{1}{\theta_l})
\neq {\cal Z}(\frac{1}{u})$. Indeed, for any $u \in {\cal O}^{(m)}$, there exists $u_1, u_2$ such that $[u_1;u_2] \subset {\cal O}^{(m)}$ and $u_1< u < u_2$. Since $x \mapsto {\cal Z}(\frac{1}{x})$ is globally nondecreasing on $\{v\neq 0, \, v \in ^c \mbox{supp~}(\nu), \, {\cal Z}^{'} (\frac{1}{v}) \leq 0\}$
(see Remark \ref{croissance globaleS}), we have if $\theta_l <u$,
${\cal Z}(\frac{1}{\theta_l})\leq {\cal Z}(\frac{1}{u_1})< {\cal Z}(\frac{1}{u}),$
and if $\theta_l >u$, ${\cal Z}(\frac{1}{u})< {\cal Z}(\frac{1}{u_2})\leq {\cal Z}(\frac{1}{\theta_l})$.
It follows, according to (\ref{car2}), that
${\cal Z}(\frac{1}{\theta_l}) $ belongs to supp $(\tau_{c,\nu})$ and therefore cannot belong to $[\rho_{\theta_j}-2\delta; \rho_{\theta_j}+2\delta]$ so that the conclusion of Lemma \ref{pourholomorphe} (2) is true.
\item Now,
let us consider $\theta_l \in \Theta_o$. By Remark \ref{compositionlimite}, $1- \theta_l g_{\tau_{c,\nu}}(x)=0$ with $x$ nonnull implies $x={\cal Z}(\frac{1}{\theta_l})=\rho_{\theta_l}.$ If $l\neq j$, $\rho_{\theta_l}$ does not belong to $[\rho_{\theta_j}-2\delta; \rho_{\theta_j}+2\delta]$ and the proof of Lemma \ref{pourholomorphe} (2) is complete. $\Box$
\end{itemize}
Let $\theta_j$ be in $\Theta_o$. According to Lemma \ref{pourholomorphe}, $\phi_l$ is an analytic function on $]\rho_{\theta_j}-2\delta; \rho_{\theta_j}+2\delta[\times \mathbb{R}$ for $l\neq j$ and $\phi_j$ is an analytic function on $]\rho_{\theta_j}-2\delta; \rho_{\theta_j}+2\delta[\times \mathbb{R}\setminus \{(\rho_{\theta_j},0)\}$.
Moreover, for any $l$, $\overline{\phi_l(z)}=\phi_l(\overline{z})$.
We have
\begin{eqnarray*}\frac{1}{\pi}\int \Im \phi_l(t+iy) h_{\delta,j}(t) dt&=&
\frac{1}{2i\pi}\int_{\rho_{\theta_j} -\delta}^{ \rho_{\theta_j} -\frac{\delta}{2}}h_{\delta,j}(t)\left[\phi_l(t+iy)
-\phi_l(t-iy)\right] dt\\
&&+\frac{1}{2i\pi}\int^{\rho_{\theta_j}+\delta}_{ \rho_{\theta_j} +\frac{\delta}{2}}h_{\delta,j}(t)\left[\phi_l(t+iy)
-\phi_l(t-iy)\right] dt\\
&&+\frac{1}{2i\pi}\int_{\rho_{\theta_j} -\frac{\delta}{2}}^{ \rho_{\theta_j} +\frac{\delta}{2}}\left[\phi_l(t+iy)
-\phi_l(t-iy)\right] dt\\
&=&\Delta_1+ \Delta_2 +\Delta_3.
\end{eqnarray*}
We immediately get that $\lim_{y \rightarrow 0^+} \Delta_1=0$ and $\lim_{y \rightarrow 0^+} \Delta_2=0$.
Now, \begin{eqnarray*}
\Delta_3&=&\frac{1}{2i\pi}\int_{\rho_{\theta_j} -\frac{\delta}{2}}^{ \rho_{\theta_j} +\frac{\delta}{2}}\left[\phi_l(t+iy)
-\phi_l(t-iy)\right] dt\\
&=& \frac{1}{2i\pi}\int_{\gamma_{j,y,\delta}}\phi_l(z)dz\\
&&-\frac{1}{2\pi}\int_{-y}^{ y}\phi_l(\rho_{\theta_j} -\frac{\delta}{2}+iu)du \\
&&+\frac{1}{2\pi}\int_{-y}^{ y}\phi_l(\rho_{\theta_j} +\frac{\delta}{2}+iu)du,\\&=&\Delta_{3,1}+ \Delta_{3,2} +\Delta_{3,3}
\end{eqnarray*}
where $\gamma_{j,y,\delta}$ is the clockwise oriented rectangular with corners $\rho_{\theta_j} -\frac{\delta}{2}-iy$,
$\rho_{\theta_j} -\frac{\delta}{2}+iy$, $\rho_{\theta_j} +\frac{\delta}{2}+iy$ and $\rho_{\theta_j} +\frac{\delta}{2}-iy$.
We immediately get that $\lim_{y \rightarrow 0^+}\Delta_{3,2}=0$ and $\lim_{y \rightarrow 0^+} \Delta_{3,3}=0$.
Moreover, for all $y$, $$\Delta_{3,1}=\frac{1}{2i\pi}\int_{\gamma_{j,y,\delta}}\phi_l(z)dz =-\mbox{Res~}(\phi_l, \rho_{\theta_j})$$
with $$\mbox{Res~}(\phi_l, \rho_{\theta_j})=0 \mbox{~if~}
l \neq j,$$
\noindent and
$$\mbox{Res~}(\phi_j, \rho_{\theta_j})=\left\{\begin{array}{ll}
\frac{k_j}{F^{(a)'}_{\sigma,\nu}(\rho_{\theta_j})}\, \hspace*{1cm} \mbox{~if~} \, M_N=M_N^W\\
=\frac{k_j \rho_{\theta_j}}{\theta_j F^{(m)'}_{c,\nu}(\frac{1}{\rho_{\theta_j}})}\, \mbox{~if~} \, M_N=M_N^S.\end{array} \right.$$
\noindent Thus
\begin{equation}\label{residu}\lim_{y \rightarrow 0^+} \frac{1}{\pi} \int h_{\delta, j}(t) \Im \phi_l(t+iy)dt
=0 \mbox{~if~} l\neq j,\end{equation}
and \begin{equation}\label{residu2}\lim_{y \rightarrow 0^{+}} \frac{1}{\pi} \int h_{\delta, j}(t)\Im \phi_j(t+iy) dt
=\left\{\begin{array}{ll}
-k_j H'(\theta_j)\, \hspace*{1cm} \mbox{~if~} \, M_N=M_N^W\\
k_j \frac{{\cal Z}^{'}(\frac{1}{\theta_j})}{\theta_j {\cal Z}(\frac{1}{\theta_j}) } \, \hspace*{0.5cm}\mbox{~if~} \, M_N=M_N^S.\end{array} \right. \end{equation}
Finally from (\ref{ecrit}), (\ref{reste}) (\ref{residu2}) and (\ref{residu}) we deduce that
$$\lim_{N\rightarrow +\infty} \lim_{y\rightarrow 0^{+}}\frac{1}{\pi} \Im \int \mathbb{E} \left( {\rm Tr} \left[G_{M_N}(t+iy)f_{\eta,l}(A_N)\right] \right)h_{\delta, j}(t) dt=0 \mbox{~if~} l\neq j,$$
and \\
$\lim_{N\rightarrow +\infty} \lim_{y\rightarrow 0^+}\frac{1}{\pi} \Im \int \mathbb{E} \left( {\rm Tr} \left[G_{M_N}(t+iy)f_{\eta,j}(A_N)\right] \right)h_{\delta, j}(t) dt $
$$=\left\{\begin{array}{ll}
-k_j H'(\theta_j)\, \hspace*{1cm} \mbox{~if~} \, M_N=M_N^W\\
k_j \frac{{\cal Z}^{'}(\frac{1}{\theta_j})}{\theta_j {\cal Z}(\frac{1}{\theta_j}) } \, \hspace*{0.5cm} \mbox{~if~} \, M_N=M_N^S.\end{array} \right. $$
Then, Proposition \ref{cvev} follows by Proposition \ref{premetap} and Lemma \ref{approxcauchy}.$\Box$
\section{Technical results specific to each model}
\begin{lemme}\label{Lipschitz}
~~\\
\begin{itemize}
\item[(1)] For any $N \times N$ Hermitian matrix $X$,\\
~
$\left\{\left(X(i,i)\right)_{1\leq i\leq N}\,\left(\sqrt{2}\Re X(i,j)\right)_{1\leq i<j \leq N} \,\left(\sqrt{2}\Im X(i,j)\right)_{1\leq i< j \leq N}\right\}$ $$ \mapsto {\rm Tr} \left[ h_{\delta,j} (X+A_N) f_{\eta,l}(A_N)\right]$$
\noindent is Lipschitz with constant bounded by $\sqrt{k_l} \Vert h_{\delta,j} \Vert_{Lip}$.
\item[(2)]For any $N \times p$ matrix $B$,
$$\left\{\left(\Re B(i,j),~ \Im B(i,j)\right)_{1\leq i \leq N, 1 \leq j \leq p}\right\} \mapsto {\rm Tr} \left[ h_{\delta,j} (A_N^{\frac{1}{2}}BB^*A_N^{\frac{1}{2}}) f_{\eta,l}(A_N)\right]$$
\noindent is Lipschitz with constant bounded by $\sqrt{k_l} \sqrt{2C} \Vert \tilde{h}_{\delta,j} \Vert_{Lip}$
where $\tilde{h}_{\delta,j} (x) =h_{\delta,j}(x^2)$ and $C=\sup_N \Vert A_N \Vert $.
\end{itemize}
\end{lemme}
\noindent {\bf Proof} Given two $N \times N$ Hermitian matrices $X$ and $X^{'}$,
we have (using Lemma \ref{extlipschitz}) that \\
$\left|{\rm Tr} \left[ h_{\delta,j} (X+A_N) f_{\eta,l}(A_N)\right]-{\rm Tr} \left[ h_{\delta,j} (X^{'}+A_N) f_{\eta,l}(A_N)\right]\right|$
$$\leq \Vert f_{\eta,l}(A_N)\Vert_2 \Vert X-X^{'} \Vert_2 \Vert h_{\delta,j} \Vert_{Lip}$$
and (1) follows since $\Vert f_{\eta,l}(A_N)\Vert_2 =\sqrt{k_l}$.\\
To prove (2) we will make use of a useful observation already made in \cite{GuZei}.
Let us introduce the $(N+p) \times (N+p)$ matrices $${\cal M}_{N+p}(B) =\left( \begin{array}{ll} 0_{p\times p}~~~ B^*A_N^{\frac{1}{2}}\\ A_N^{\frac{1}{2}}B ~~~ 0_{N\times N} \end{array} \right),$$
$${\cal N}_{N+p}=\left( \begin{array}{ll} 0_{p\times p}~~~ 0_{p\times N}\\ 0_{N\times p} ~~~ f_{\eta,l}(A_N)\end{array} \right).$$
It is easy to see that
$${\rm Tr} \left[ h_{\delta,j} (A_N^{\frac{1}{2}}BB^*A_N^{\frac{1}{2}}) f_{\eta,l}(A_N)\right]=
{\rm Tr} \left[ \tilde{h}_{\delta,j} ({\cal M}_{N+p}(B)) {\cal N}_{N+p}\right],$$
where $\tilde{h}_{\delta,j}(x) =h_{\delta,j}(x^2)$. Note that since $h_{\delta,j}$ is a ${\cal C}^\infty$ compactly supported function,
$\tilde{h}_{\delta,j}$ is obviously a Lipschitz function.
Hence \\
$\left| {\rm Tr} \left[ h_{\delta,j} (A_N^{\frac{1}{2}}BB^*A_N^{\frac{1}{2}}) f_{\eta,l}(A_N)\right]
-{\rm Tr} \left[ h_{\delta,j} (A_N^{\frac{1}{2}}B^{'}(B^{'})^*A_N^{\frac{1}{2}}) f_{\eta,l}(A_N)\right]\right|$
\begin{eqnarray}&= &\left| {\rm Tr} \left[ \tilde{h}_{\delta,j} ({\cal M}_{N+p}(B)) {\cal N}_{N+p}\right]- {\rm Tr} \left[ \tilde{h}_{\delta,j} ({\cal M}_{N+p}(B^{'})) {\cal N}_{N+p}\right]\right| \nonumber\\ &\leq &
\left\| {\cal N}_{N+p}\right\|_2 \left\| \tilde{h}_{\delta,j} ({\cal M}_{N+p}(B))-
\tilde{h}_{\delta,j} ({\cal M}_{N+p}(B^{'}))\right\|_2 \nonumber\\
&\leq & \sqrt{k_l}\left\|\tilde{h}_{\delta,j} \right\|_{Lip} \left\| {\cal M}_{N+p}(B)-{\cal M}_{N+p}(B^{'}) \right\|_2.\label{lip1}
\end{eqnarray}
\noindent where we used Lemma \ref{extlipschitz} in the last line.
Now, \begin{eqnarray}\left\| {\cal M}_{N+p}(B)-{\cal M}_{N+p}(B^{'}) \right\|^2_2&=& 2 {\rm Tr} \left[ (B-B^{'})^{*}A_N (B-B^{'}) \right]\nonumber \\ &\leq & 2\left\| A_N \right\| \left\| B-B^{'}\right\|^2_2
\nonumber \\ & \leq & 2 C \left\| B-B^{'}\right\|^2_2.\label{lip2}\end{eqnarray}
\noindent (2) readily follows from (\ref{lip1}) and (\ref{lip2}).
$\Box$\\
We have proved in Lemma 3.3 \cite{CDFF10} that $\forall z \in \mathbb{C}\setminus \mathbb{R}$,
\begin{equation}\label{estimfondakW}
\mathbb {E} (\tilde G^W_{kk}(z)) = \frac{1}{(z-\sigma ^2g^W_N(z)-\gamma_k)}
+R_{k,N}(z),
\end{equation}
with $$\vert R_{k,N}(z) \vert \leq \frac{P(\vert \Im z \vert^{-1})}{N}$$
\noindent for some
polynomial $P$ with nonnegative coefficients.
Note that \begin{equation} \label{imageW} \vert \Im (z-\sigma ^2g^W_N(z)-\gamma_k)\vert \geq \vert \Im z \vert.\end{equation}
We are going to establish the following similar result for the sample covariance matrix setting using many ideas from \cite{BaiSil06}.
\begin{proposition} \label{estimresolvwish} There exists a polynomial $P$ with nonnegative coefficients and a sequence $a_N$ of nonnegative
real numbers converging to zero when $N$ goes to infinity such that, for any $k$ in $\{1, \ldots,N\}$, any $z$ in $\mathbb{C}\setminus \mathbb{R}$,
\begin{equation} \label{submat}
\mathbb {E} (\tilde G^S_{kk}(z)) = \frac{1}{z-\gamma_k(1-\frac{N}{p}+\frac{N}{p}zg^S_N(z))}
+R_{k,N}(z),
\end{equation}
with $$\left|R_{k,N}(z) \right| \leq (1+\vert z\vert)^2 P(\vert \Im z \vert^{-1})a_N.$$
\end{proposition}
\noindent {\bf Proof}
Let $J$ be a $N\times N$ matrix and $u$ be a vector in $\mathbb{C}^N$ such that $J$ and $J +uu^*$ are invertible then
\begin{eqnarray*}
u^* J^{-1} u u^* \left( J+ uu^* \right)^{-1} & = & u^* J^{-1} \left(u u^* +J \right) \left( J+ uu^* \right)^{-1}
- u^* \left( J+ uu^* \right)^{-1}\\&=& u^* J^{-1}-u^* \left( J+ uu^* \right)^{-1}
\end{eqnarray*}
so that \begin{equation}\label{trick}
u^* \left( J+ uu^* \right)^{-1}=\frac{u^* J^{-1}}{1+u^* J^{-1}u}.
\end{equation}
Hence if $u_1,\ldots,u_p$ are $p$ vectors and $X=\sum_{i=1}^p u_i u_i^*$, denoting by $G_X(z)$ the resolvent
$(zI-X)^{-1}$,
(\ref{trick}) yields that for any $i\in \{1,\ldots,p\}$, for any $z\in \mathbb{C}\setminus \mathbb{R}$,
\begin{equation}\label {etape1}u_iG_X(z) = \frac{u_i^* \left(zI-\sum_{l\neq i} u_l u_l^*\right)^{-1}}{1-u_i^* \left(zI-\sum_{l\neq i} u_l u_l^*\right)^{-1}u_i}.\end{equation}
Multiplying (\ref{etape1}) by $u_i$ and summing in $i$ yields
\begin{equation}\label{etape2}
XG_X(z) =\sum_{i=1}^p \frac{u_i u_i^* \left(zI-\sum_{l\neq i} u_l u_l^*\right)^{-1}}{1-u_i^* \left(zI-\sum_{l\neq i} u_l u_l^*\right)^{-1}u_i}.
\end{equation}
From (\ref{etape2}) and the resolvent identity
$$-I+zG_X(z) =XG_X(z),$$ we deduce that
\begin{equation}\label{Gkk}
\left(G_X(z)\right)_{kk} =\frac{1}{z} +
\sum_{i=1}^p \frac{\left[u_i u_i^* \left(zI-\sum_{l\neq i} u_l u_l^*\right)^{-1}\right]_{kk}}{z \left[1-u_i^* \left(zI-\sum_{l\neq i} u_l u_l^*\right)^{-1}u_i\right]}.
\end{equation}
Noticing that \begin{equation}\label{remarque} \tilde{G}^S(z)= G_{\frac{1}{p} D^{\frac{1}{2}}U X_N^S U^* D^{\frac{1}{2}}}\end{equation}
and that $$\frac{1}{p} D^{\frac{1}{2}}U X_N^S U^* D^{\frac{1}{2}} =\sum_{i=1}^p u_i u_i^*$$
where $u_i =\frac{1}{\sqrt{p}} D^{\frac{1}{2}} U x_i$ and $x_i$ is the $i$th column of $B_N$,
we deduce from (\ref{Gkk}) that
\begin{equation}\label{Gkk2}
\tilde{G}^S_{kk}(z)=\frac{1}{z} + \frac{1}{p} \sum_{i=1}^p
\frac{\left[D^{\frac{1}{2}} U x_i x_i^* U^* D^{\frac{1}{2}} \left(zI-\frac{1}{p}\sum_{l\neq i} D^{\frac{1}{2}} U x_l x_l^* U^* D^{\frac{1}{2}}\right)^{-1}\right]_{kk}}{z \left\{1-\frac{1}{p} x_i^* U^* D^{\frac{1}{2}} \left(zI- \frac{1}{p} \sum_{l\neq i} D^{\frac{1}{2}} U x_l x_l^* U^* D^{\frac{1}{2}}\right)^{-1}D^{\frac{1}{2}} U x_i \right\}}.
\end{equation}
Set for $i=1,\ldots,p$, $$y_i = \frac{1}{\sqrt{p}} A_N^{\frac{1}{2}}x_i$$
and $$M_N^{(i)} = \frac{1}{{p}} \sum_{l\neq i} A_N^{\frac{1}{2}}x_l x_l^* A_N^{\frac{1}{2}}=\sum_{l\neq i} y_l y_l^*.$$
Note that the $y_i$'s are i.i.d and that $y_i$ is independent of $M_N^{(i)}$.
Note also that $$M_N^S =M_N^{(i)} + y_i y_i^*.$$
(\ref{Gkk2}) can be rewritten as follows
\begin{equation}\label{Gkk3}
\tilde{G}^S_{kk}(z)=\frac{1}{z} + \frac{1}{p} \sum_{i=1}^p
\frac{\left[D^{\frac{1}{2}} U x_i x_i^* A_N^{\frac{1}{2}} \left(zI-M_N^{(i)}\right)^{-1}U^*\right]_{kk}}{z \left\{1-y_i^* \left(zI- M_N^{(i)}\right)^{-1} y_i \right\}}
\end{equation}
Now applying (\ref{Gkk}) with $u_i= y_i$ and $X= M_N^S$, summing on $k$ and dividing by $N$ we have
\begin{eqnarray}
{\rm tr}_N\left( G_{M_N^S} (z)\right) &=& \frac{1}{z} + \frac{1}{N} \sum_{i=1}^p
\frac{y_i^* \left(zI-M_N^{(i)}\right)^{-1}y_i}{z \left\{1-y_i^* \left(zI- M_N^{(i)}\right)^{-1} y_i \right\}}\nonumber\\
&=&
\frac{1}{z}-\frac{p}{N} \frac{1}{z} + \frac{1}{N} \sum_{i=1}^p
\frac{1}{z \left\{1-y_i^* \left(zI- M_N^{(i)}\right)^{-1} y_i \right\}}.\label{trace}
\end{eqnarray}
Let us define the $p\times p$ matrix
$$\underline{M}_N^S= \frac{1}{p}B_N^* A_N B_N.$$
Since
\begin{equation}\label{baroupas}
{\rm tr}_p\left( G_{\underline{M}_N^S} (z)\right)=\frac{1-\frac{N}{p}}{z} +\frac{N}{p}
{\rm tr}_N\left( G_{M_N^S} (z)\right),\end{equation}
we deduce from (\ref{trace}) that
\begin{equation}\label{tracebar}
{\rm tr}_N\left( G_{\underline{M}_N^S} (z)\right) = \frac{1}{p} \sum_{i=1}^p
\frac{1}{z \left\{1-y_i^* \left(zI- M_N^{(i)}\right)^{-1} y_i \right\}}.
\end{equation}
Following the ideas of Section 6.4.1 of \cite{BaiSil06}, we are going to establish the following preliminary lemma.
\begin{lemme} \label{normeL2}
There exists a constant $K>0$ and a sequence of nonnegative numbers $a_N$ converging to zero when $N$ goes to infinity such that for each $i=1, \ldots,p$, $\forall z \in \mathbb{C}\setminus \mathbb{R}$,
$$\left\| {\rm tr}_p\left( G_{\underline{M}_N^S} (z)\right) - \frac{1}{z \left\{1-y_i^* \left(zI- M_N^{(i)}\right)^{-1} y_i \right\}} \right\|_{L^2} \leq K \frac{\vert z \vert }{\vert \Im z \vert^3}a_N.$$
\end{lemme}
\noindent We have from (\ref{tracebar})\\
${\rm tr}_p\left( G_{\underline{M}_N^S} (z)\right) - \frac{1}{z \left\{1-y_i^* \left(zI- M_N^{(i)}\right)^{-1} y_i \right\}} $
$$= \frac{1}{p} \sum_{l\neq i}
\frac{y_l^* \left(zI- M_N^{(l)}\right)^{-1} y_l - y_i^* \left(zI- M_N^{(i)}\right)^{-1} y_i }{z \left\{1-y_l^* \left(zI- M_N^{(l)}\right)^{-1} y_l \right\}\left\{1-y_i^* \left(zI- M_N^{(i)}\right)^{-1} y_i \right\}}.$$
\begin{lemme}\label{maj}
For any $N\times N$ positive semidefinite matrix $H$, any vector $v$ in $\mathbb{C}^N$ and any $z$ in $
\mathbb{C}\setminus \mathbb{R}$,
$$\frac{1}{\vert z \left\{1-v^* (zI- H)^{-1} v \right\}\vert } \leq \frac{1}{\vert \Im z \vert }.$$
\end{lemme}
\noindent {\bf Proof:}\begin{eqnarray*}
\Im \left\{ z v^* (zI- H)^{-1} v \right\}&=& \frac{1}{2i} \left\{
z v^* (zI- H)^{-1} v -\bar{z} v^* (\bar{z}I- H)^{-1} v \right\}\\
&=&
\frac{1}{2i} v^*\left\{
(I- \frac{1}{z}H)^{-1} -(I- \frac{1}{\bar{z}} H)^{-1} \right\}v\\
&=&
-\frac{\Im z}{\vert z \vert^2} v^*
(I- \frac{1}{z}H)^{-1}H(I- \frac{1}{\bar{z}} H)^{-1} v.
\end{eqnarray*}
Hence \begin{eqnarray*}
\vert \Im \left[ z \left\{1-v^* (zI- H)^{-1} v \right\} \right] \vert &=&
\vert \Im z \vert \left\{ 1 + \frac{1\!\!{\sf I}}{\vert z \vert^2}v^*
(I- \frac{1}{z}H)^{-1}H(I- \frac{1}{\bar{z}} H)^{-1} v \right\}\\
& \geq & \vert \Im z \vert
\end{eqnarray*}
and Lemma \ref{maj} follows.$\Box$\\
\noindent According to Lemma \ref{maj}, for any $l$ and $i$ in $\{1, \ldots,p\}$,
$$\frac{1}{\left|z \left\{1-y_l^* \left(zI- M_N^{(l)}\right)^{-1} y_l \right\}\left\{1-y_i^* \left(zI- M_N^{(i)}\right)^{-1} y_i \right\}\right| }\leq \frac{\vert z \vert}{\vert \Im z \vert^2 }.$$
Hence \\
$\left\| {\rm tr}_p\left( G_{\underline{M}_N^S} (z)\right) - \frac{1}{z \left\{1-y_i^* \left(zI- M_N^{(i)}\right)^{-1} y_i \right\}} \right\|_{L^2}$
$$\leq \frac{\vert z \vert}{\vert \Im z \vert^2 } \frac{1}{p} \sum_{l\neq i} \left\|
y_l^* \left(zI- M_N^{(l)}\right)^{-1} y_l - y_i^* \left(zI- M_N^{(i)}\right)^{-1} y_i \right\|_{L^2}.$$
We have\\
$
y_l^* \left(zI- M_N^{(l)}\right)^{-1} y_l - y_i^* \left(zI- M_N^{(i)}\right)^{-1} y_i$
\begin{eqnarray*}&=&
y_l^* \left(zI- M_N^{(l)}\right)^{-1} y_l - \frac{1}{p}
{\rm Tr} \left[\left(zI- M_N^{(l)}\right)^{-1} A_N \right]\\ & &
+\frac{1}{p}
{\rm Tr} \left[\left(zI- M_N^{(l)}\right)^{-1} A_N \right] - \frac{1}{p}
{\rm Tr} \left[\left(zI- M_N^{(i)}\right)^{-1} A_N \right]\\
&& +\frac{1}{p}
{\rm Tr} \left[\left(zI- M_N^{(i)}\right)^{-1} A_N \right]
-y_i^* \left(zI- M_N^{(i)}\right)^{-1} y_i\\
&=& \Delta_l + \Delta_{l,i} - \Delta_i.
\end{eqnarray*}
We have
\begin{eqnarray*}
\vert \Delta_{l,i} \vert &\leq & \left| \frac{1}{p}
{\rm Tr} \left[\left\{\left(zI- M_N^{(l)}\right)^{-1}-\left(zI- M_N^S \right)^{-1} \right\}A_N \right] \right| \\
& & +\left| \frac{1}{p}
{\rm Tr} \left[\left\{\left(zI- M_N^S\right)^{-1}-\left(zI- M_N^{(i)} \right)^{-1} \right\}A_N \right] \right|\\
&\leq & \frac{2 \Vert A_N \Vert}{\vert \Im z \vert p}\leq \frac{2 C}{\vert \Im z \vert p}
\end{eqnarray*}
where we used Lemma 6.9 \cite{BaiSil06}) in the last line. \\
Now we have for any $l=1,\ldots,p$,
$$y_l^* \left(zI- M_N^{(l)}\right)^{-1} y_l= \frac{1}{p} x_l^* A_N^{\frac{1}{2}}\left(zI- M_N^{(l)}\right)^{-1}
A_N^{\frac{1}{2}}x_l$$
so that according to Proposition \ref{QF} in the Appendix,
for any $l=1,\ldots,p$, $$\Vert \Delta_{l} \Vert_{L^2} \leq \sqrt{K} \sqrt{\frac{N}{p}} \frac{1}{\sqrt{p}}
\frac{ \Vert A_N \Vert}{\vert \Im z \vert }
\leq \sqrt{\frac{N}{p}} \frac{1}{\sqrt{p}}
\frac{ C}{\vert \Im z \vert }.$$
It follows that
$$\left\| {\rm tr}_p\left( G_{\underline{M}_N^S} (z)\right) - \frac{1}{z \left\{1-y_i^* \left(zI- M_N^{(i)}\right)^{-1} y_i \right\}} \right\|_{L^2} \leq \frac{2C \vert z \vert }{\vert \Im z \vert^3}
\left( \frac{1}{p}+ \sqrt{\frac{N}{p}} \frac{1}{\sqrt{p}} \right)$$
and the proof of Lemma \ref{normeL2} is complete. $\Box$
\begin{lemme}\label{phi}
There exists a constant $K>0$ such that, for any $k=1, \ldots, N$, for any $i=1,\ldots,p$ and any $z$ in $\mathbb{C}\setminus \mathbb{R}$,
$$\left\|
\left[D^{\frac{1}{2}} U x_i x_i^* A_N^{\frac{1}{2}} \left(zI-M_N^{(i)}\right)^{-1}U^*\right]_{kk}
\right\|_{L^2} \leq \frac{K}{\vert \Im z \vert}.$$
\end{lemme}
\noindent {\bf Proof:}
Note that
\begin{eqnarray*}
\left[D^{\frac{1}{2}} U x_i x_i^* A_N^{\frac{1}{2}} \left(zI-M_N^{(i)}\right)^{-1}U^*\right]_{kk}&=&
{\rm Tr} D^{\frac{1}{2}} U x_i x_i^* A_N^{\frac{1}{2}} \left(zI-M_N^{(i)}\right)^{-1}U^*E_{kk}\\
&=& x_i^* A_N^{\frac{1}{2}} \left(zI-M_N^{(i)}\right)^{-1}U^*E_{kk}D^{\frac{1}{2}} U x_i.
\end{eqnarray*}
Thus, according to Proposition \ref{QF} in the Appendix,\\
$
\left\|
\left[D^{\frac{1}{2}} U x_i x_i^* A_N^{\frac{1}{2}} \left(zI-M_N^{(i)}\right)^{-1}U^*\right]_{kk}
-{\rm Tr} A_N^{\frac{1}{2}} \left(zI-M_N^{(i)}\right)^{-1}U^*E_{kk}D^{\frac{1}{2}} U
\right\|_{L^2}$ $$\leq K
\left[ {\rm Tr} A_N^{\frac{1}{2}} \left(zI-M_N^{(i)}\right)^{-1}U^*E_{kk}D E_{kk} U \left(\bar{z}I-M_N^{(i)}\right)^{-1} A_N^{\frac{1}{2}}\right]^{\frac{1}{2}} \leq \frac{C}{\vert \Im z\vert}.
$$
Since moreover
$$\left| {\rm Tr} A_N^{\frac{1}{2}} \left(zI-M_N^{(i)}\right)^{-1}U^*E_{kk}D^{\frac{1}{2}} U \right| \leq \frac{C}{\vert \Im z\vert},$$
we deduce that $$\left\|
\left[D^{\frac{1}{2}} U x_i x_i^* A_N^{\frac{1}{2}} \left(zI-M_N^{(i)}\right)^{-1}U^*\right]_{kk}
\right\|_{L^2} \leq \frac{2C}{\vert \Im z \vert}.$$
\hspace*{10cm}$\Box$\\
We will need this last lemma concerning the variance of
${\rm tr}_p\left( G_{\underline{M}_N^S} (z)\right)$.
\begin{lemme}\label{variancebar} There exists some polynomial $P$ with nonnegative coefficients such that, $\forall z \in \mathbb{C}\setminus \mathbb{R}$,
$$\left\| {\rm tr}_p\left( G_{\underline{M}_N^S} (z)\right) - \mathbb{E}\left( {\rm tr}_p\left( G_{\underline{M}_N^S} (z)\right)\right) \right\|_{L^2} \leq \frac{1}{p^2} \left( \vert z\vert +1\right)^2 P\left(\vert \Im z \vert^{-1}\right).$$
\end{lemme}
\noindent{\bf Proof:}
Let us define $\Psi: \mathbb {R} ^{2(p\times N)} \rightarrow {\mathcal M }_{N\times p}(\mathbb {C})$ by
$$\Psi:~~\{x_{ij},y_{ij},i=1,\ldots,N,j=1,\ldots,p\}\rightarrow
\sum_{i=1,\ldots,N}\sum_{j=1,\ldots,p} \left( x_{ij} +\sqrt{-1} y_{ij} \right) E_{ij}.$$
Let $F$ be a smooth complex function on ${ M}_{N\times p}(\mathbb {C})$ and
define the complex function $f$ on $\mathbb {R}^{2(p\times N)}$ by setting $f=F\circ \Psi$.
Then,$$\Vert {\rm grad} f(u)\Vert = \sup_{V\in { M }_{N\times p}(\mathbb {C}), Tr VV^*=1}
\left| \frac{d}{dt} F(\Psi(u)+tV)\vert _{t=0}\right| .$$
We have $B_N=\Psi(\Re ((B_N)_{ij}),\Im ((B_N)_{ij}),1\leq i\leq N,1\leq j\leq p)$
where the distribution of $\{\sqrt{2}\Re ((B_N)_{ij}),\sqrt{2}\Im ((B_N)_{ij}),1\leq i\leq N,1\leq j\leq p\}$
satisfies a Poincar\'e inequality with constant $C_{PI}$.\\
Hence consider $F:~B \rightarrow {\rm tr} _N(zI_N -A_N^{\frac{1}{2}} \frac{B B^*}{p}A_N^{\frac{1}{2}} )^{-1}$.\\
Let $ V\in { M }_{N\times p}(\mathbb {C})$ such that $ Tr VV^*=1$.
\begin{eqnarray}
\frac{d}{dt} F(B+tV)\vert _{t=0}&=&\frac{1}{Np} {\rm Tr}(G_{M_N^S}(z) A_N^{\frac{1}{2}} V B^* A_N^{\frac{1}{2}} G_{M_N^S}(z)) \nonumber \\
&&+ \frac{1}{Np} {\rm Tr}(G_{M_N^S}(z) A_N^{\frac{1}{2}} B V^* A_N^{\frac{1}{2}} G_{M_N^S}(z)). \label{derive}
\end{eqnarray}
By Cauchy-Schwartz inequality, we have \\
$
\left| \frac{1}{Np} {\rm Tr}(G_{M_N^S}(z) A_N^{\frac{1}{2}} V B_N^* A_N^{\frac{1}{2}} G_{M_N^S}(z)) \right| $
\begin{eqnarray*}
&\leq &\frac{1}{N \sqrt{p}}
\left[{\rm Tr} \frac{B_N^*A_N^{\frac{1}{2}} \left[G_{M_N^S}(z)\right]^2 A_N \left[G_{M_N^S}(\bar{z})\right]^2 A_N^{\frac{1}{2}} B_N }{p}\right] ^{\frac{1}{2}}
({\rm Tr} VV^* )^{\frac{1}{2}}\\
&= & \frac{1}{\sqrt{Np}}\left[{\rm tr}_N M_N^S \left[G_{M_N^S}(z)\right]^2 A_N \left[G_{M_N^S}(\bar{z})\right]^2
\right] ^{\frac{1}{2}}.
\end{eqnarray*}
Since by the resolvent identity
$$M_N^S G_{M_N^S}(z)=-I_N +z G_{M_N^S},$$ we have \\
${\rm tr}_N M_N^S \left[G_{M_N^S}(z)\right]^2 A_N \left[G_{M_N^S}(\bar{z})\right]^2$ $$ =
-{\rm tr}_N G_{M_N^S}(z) A_N \left[G_{M_N^S}(\bar{z})\right]^2 +z {\rm tr}_N \left[G_{M_N^S}(z)\right]^2 A_N \left[G_{M_N^S}(\bar{z})\right]^2,$$
we can deduce that
$${\rm tr}_N M_N^S \left[G_{M_N^S}(z)\right]^2 A_N \left[G_{M_N^S}(\bar{z})\right]^2 \leq \frac{\Vert A_N \Vert}{\vert \Im z \vert^3} +\frac{\vert z \vert \Vert A_N \Vert}{\vert \Im z \vert^4}\leq C(\vert z \vert +1) P\left( \vert \Im z \vert^{-1} \right),$$
\noindent where $P$ is a polynomial with non negative coefficients.
Hence
$$\left| \frac{1}{Np} {\rm Tr}(G_{M_N^S}(z) A_N^{\frac{1}{2}} V B_N^* A_N^{\frac{1}{2}} G_{M_N^S}(z)) \right| \leq
\frac{1}{\sqrt{Np}}C(\vert z \vert +1) P\left( \vert \Im z \vert^{-1} \right).$$
Since a similar upper bound can be obtained in the same way for the second term on the right hand side of (\ref{derive})
we deduce that
$$\mathbb{E}\left( \left( \sup_{V\in { M }_{N\times p}(\mathbb {C}), Tr VV^*=1}
\left| \frac{d}{dt} F(B_N+tV)\vert _{t=0}\right|\right)^2 \right)^{\frac{1}{2}} \leq \frac{2C}{\sqrt{Np}}(\vert z \vert +1) P\left( \vert \Im z \vert^{-1} \right).$$
Therefore, Poincar\'e inequality yields
$$\left\| {\rm tr}_N\left( G_{{M}_N^S} (z)\right) - \mathbb{E}\left( {\rm tr}_N\left( G_{{M}_N^S} (z)\right)\right) \right\|_{L^2}\leq
\frac{1}{{Np}}(\vert z \vert +1)^2 Q\left(\vert \Im z \vert^{-1} \right)$$
\noindent where $Q$ is a polynomial with non negative coefficients.
Now, since
$${\rm tr}_p\left( G_{\underline{M}_N^S} (z)\right) - \mathbb{E}\left( {\rm tr}_p\left( G_{\underline{M}_N^S} (z)\right)\right)
=\frac{N}{p} \left[{\rm tr}_N\left( G_{{M}_N^S} (z)\right) - \mathbb{E}\left( {\rm tr}_N\left( G_{{M}_N^S} (z)\right)\right) \right],$$
Lemma \ref{variancebar} follows. $\Box$\\
\noindent Using (\ref{Gkk3}), we have for any $z$ in $\mathbb{C}\setminus \mathbb{R}$,
\begin{eqnarray*}
\mathbb{E}\left( \tilde{G}^S_{kk}(z)\right) &=&\frac{1}{z} +
\frac{1}{p} \sum_{i=1}^p \mathbb{E} \left(\Phi_{i,k} \right) \mathbb{E} \left({\rm tr}_p G_{\underline{M}_N^S} (z)\right) \\&& +\frac{1}{p} \sum_{i=1}^p \mathbb{E} \left[\Phi_{i,k} \left\{{\rm tr}_p G_{\underline{M}_N^S} (z)
- \mathbb{E} \left({\rm tr}_p G_{\underline{M}_N^S} (z)\right)\right\}\right]\\
&& +\frac{1}{p} \sum_{i=1}^p \mathbb{E} \left[\Phi_{i,k} \left\{ \frac{1}{z \left\{1-y_i^* \left(zI- M_N^{(i)}\right)^{-1} y_i \right\}}
-{\rm tr}_p G_{\underline{M}_N^S} (z)
\right\}\right]
\end{eqnarray*}
where $$\Phi_{i,k}=
\left[D^{\frac{1}{2}} U x_i x_i^* A_N^{\frac{1}{2}} \left(zI-M_N^{(i)}\right)^{-1}U^*\right]_{kk}.$$
By Cauchy-Schwartz inequality, using Lemmas \ref{phi} and \ref{normeL2}, we easily have that there exists a constant $K$
and a sequence of nonnegative numbers $a_N$ converging towards zero when $N$ goes to infinity such that for any $k=1,\ldots,N$,
$$\left| \frac{1}{p} \sum_{i=1}^p \mathbb{E} \left[\Phi_{i,k} \left\{ \frac{1}{z \left\{1-y_i^* \left(zI- M_N^{(i)}\right)^{-1} y_i \right\}}
-{\rm tr}_p G_{\underline{M}_N^S} (z)
\right\}\right]\right| \leq \frac{K \vert z \vert}{\vert \Im z \vert^4} a_N.$$
By Cauchy-Schwartz inequality, using Lemmas \ref{phi} and \ref{variancebar}, we also have that there exists a polynomial $P$ with
nonnegative coefficients such that
$$\left| \frac{1}{p} \sum_{i=1}^p \mathbb{E} \left[\Phi_{i,k} \left\{{\rm tr}_p G_{\underline{M}_N^S} (z)
- \mathbb{E} \left({\rm tr}_p G_{\underline{M}_N^S} (z)\right)\right\}\right]\right|\leq \frac{1}{p^2}(\vert z \vert +1)^2
P(\vert \Im z \vert^{-1}).$$
Thus
\begin{equation}\label{eqinter} \mathbb{E}\left( \tilde{G}^S_{kk}(z)\right) =\frac{1}{z} +
\frac{1}{p} \sum_{i=1}^p \mathbb{E} \left(\Phi_{i,k} \right) \mathbb{E} \left({\rm tr}_p G_{\underline{M}_N^S} (z)\right) + \Delta_N(k)\end{equation}
where there exists a polynomial $Q$ with
nonnegative coefficients and a sequence of nonnegative numbers $b_N$ converging towards zero when $N$ goes to infinity
such that, for any $k=1,\ldots,N$,
$$\vert \Delta_N(k) \vert \leq (\vert z \vert +1)^2
Q(\vert \Im z \vert^{-1}) b_N.$$
\noindent Now, one can easily see that
\begin{eqnarray}
\mathbb{E} \left(\Phi_{i,k} \right) &=& \gamma_k \mathbb{E} ( [U
(zI-M_N^{(i)})^{-1}U^*]_{kk} ) \nonumber \\
&=& \gamma_k \mathbb{E} ( [
(zI-\sum_{l\neq i} u_l u_l^*)^{-1}]_{kk} ) \label{eqphi}
\end{eqnarray}
where $$u_i=\frac{1}{\sqrt{p}}D^{\frac{1}{2}}Ux_i.$$
\begin{lemme}\label{difftilde}
There exists a polynomial $P$ with
nonnegative coefficients such that for any $i=1, \ldots,p$, any $k=1,\ldots,N$, any
$ z \in \mathbb{C}\setminus \mathbb{R}$,
$$\left|\mathbb{E}\left( \tilde{G}^S_{kk}(z)\right)- \mathbb{E} ( [
(zI-\sum_{l\neq i} u_l u_l^*)^{-1}]_{kk}) \right| \leq \frac{1}{p}(\vert z \vert +1)
P(\vert \Im z \vert^{-1}).$$
\end{lemme}
\noindent {\bf Proof:} Remember that according to (\ref{remarque}),
$$\mathbb{E}( \tilde{G}^S_{kk}(z))=\mathbb{E} ( [
(zI-\sum_{l=1}^p u_l u_l^*)^{-1}]_{kk} ).
$$
By the formula (3.3.4) in \cite{BaiSil06},
we have $$[
(zI-\sum_{l=1}^p u_l u_l^*)^{-1}]_{kk}=[
(zI-\sum_{l\neq i} u_l u_l^*)^{-1}]_{kk}+ \frac{\psi_{i,k} }{1-u_i^*\left(zI-\sum_{l\neq i} u_l u_l^*\right)^{-1}u_i}$$
where $$\psi_{i,k}= \left[
\left(zI-\sum_{l\neq i} u_l u_l^*\right)^{-1}u_i u_i^*\left(zI-\sum_{l\neq i} u_l u_l^*\right)^{-1}\right]_{kk} .$$
\noindent Noticing that $$\psi_{i,k}= \frac{1}{p} x_i^* U^* D^{\frac{1}{2}} (zI-\sum_{l\neq i} u_l u_l^*)^{-1}
E_{kk} (zI-\sum_{l\neq i} u_l u_l^*)^{-1} D^{\frac{1}{2}} U x_i,$$
we have by Proposition \ref{QF} that\\
$
\left\| \psi_{i,k}- \frac{1}{p} {\rm Tr} D \left(zI-\sum_{l\neq i} u_l u_l^*\right)^{-1}
E_{kk} \left(zI-\sum_{l\neq i} u_l u_l^*\right)^{-1} \right\|_{L^2}$
$$\leq \frac{K}{p}
\left\{ {\rm Tr} D (zI-\sum_{l\neq i} u_l u_l^*)^{-1}
E_{kk} (zI-\sum_{l\neq i} u_l u_l^*)^{-1}
(\bar{z}I-\sum_{l\neq i} u_l u_l^*)^{-1}
E_{kk} (\bar{z}I-\sum_{l\neq i} u_l u_l^*)^{-1} D \right\}^{\frac{1}{2}}$$
$\leq\frac{CK}{p \vert \Im z \vert^{2}}.
$\\
~
\noindent Using also Lemma \ref{maj}, we readily have that \\
$\left|\mathbb{E}\left( \tilde{G}^S_{kk}(z)\right)- \mathbb{E} \left( \left[
\left(zI-\sum_{l\neq i} u_l u_l^*\right)^{-1}\right]_{kk} \right) \right|$
$$\leq \frac{\vert z \vert}{p \vert \Im z \vert} \left\{ \frac{CK}{ \vert \Im z \vert^{2}} +
\mathbb{E} \left( \left| {\rm Tr} D(zI-\sum_{l\neq i} u_l u_l^*)^{-1} E_{kk} (zI-\sum_{l\neq i} u_l u_l^*)^{-1} \right| \right) \right\}.$$
Since $$\left| {\rm Tr} D(zI-\sum_{l\neq i} u_l u_l^*)^{-1} E_{kk} (zI-\sum_{l\neq i} u_l u_l^*)^{-1} \right| \leq \frac{C}{ \vert \Im z \vert^{2}},$$
Lemma \ref{difftilde} readily follows. $\Box$\\
\noindent Hence (\ref{eqphi}) and Lemma \ref{difftilde} yield
$$\mathbb{E}\left( \Phi_{i,k}\right) = \gamma_k \mathbb{E}\left( \tilde{G}^S_{kk}(z)\right) +
\tau_{i,k}$$
with $\vert \tau_{i,k}\vert \leq \frac{1}{p} \left( \vert z \vert +1 \right)
P(\vert \Im z \vert^{-1}) $ and thus, using equation (\ref{eqinter}), there exists a polynomial $Q$ with
nonnegative coefficients and a sequence of nonnegative numbers $a_N$ converging towards zero when $N$ goes to infinity such that for any $k=1,\ldots,N$ ,
$$\mathbb{E}\left( \tilde{G}^S_{kk}(z)\right)= \frac{1}{z} + \gamma_k\mathbb{E}\left( \tilde{G}^S_{kk}(z)\right)
\mathbb{E}\left( {\rm tr}_p G_{\underline{M}_N^S}(z) \right) + \xi_k$$
with $$\vert \xi_k\vert \leq \left( \vert z \vert +1 \right)^2
Q(\vert \Im z \vert^{-1})a_N.$$
Thus \begin{equation}\label{eqinterm}
\left\{z- \gamma_k z \mathbb{E}\left( {\rm tr}_p G_{\underline{M}_N^S}(z) \right) \right\}
\mathbb{E}\left( \tilde{G}^S_{kk}(z)\right)=1 + z \xi_k.
\end{equation}
Using the resolvent identity $$ zG_{\underline{M}_N^S}(z)=I_p + \underline{M}_N^SG_{\underline{M}_N^S}(z)$$
we can easily see that
$$\Im \left[ z {\rm tr}_p G_{\underline{M}_N^S}(z) \right]=-(\Im z) {\rm tr}_p G_{\underline{M}_N^S}(z)^* \underline{M}_N^SG_{\underline{M}_N^S}(z).$$
Hence \\
$
\vert \Im \left\{z- \gamma_k z \mathbb{E}\left( {\rm tr}_p G_{\underline{M}_N^S}(z) \right) \right\}\vert$
\begin{equation} \label{image} = \vert \Im z \vert \left\{ 1+ \gamma_k{\rm tr}_p G_{\underline{M}_N^S}(z)^* \underline{M}_N^SG_{\underline{M}_N^S}(z) \right\} \geq \vert \Im z \vert.\end{equation}
Thus (\ref{eqinterm}) yields that
$$\mathbb{E}\left( \tilde{G}_{kk}^S(z)\right)= \frac{1}{z- \gamma_k z \mathbb{E}\left( {\rm tr}_p G_{\underline{M}_N^S}(z)\right)} + \xi^{'}_k$$
with $$\vert \xi^{'}_k\vert \leq \frac{\vert z\vert }{\vert \Im z \vert}\left( \vert z \vert +1 \right)^2
Q(\vert \Im z \vert^{-1})a_N.$$
Proposition \ref{estimresolvwish} readily follows since (see(\ref{baroupas})) we have
$$\mathbb{E}\left( {\rm tr}_p G_{\underline{M}_N^S}(z)\right)=\frac{N}{p} \mathbb{E}\left( {\rm tr}_N G_{{M}_N^S}(z)\right)
+ \frac{1-\frac{N}{p}}{z}.$$ $\Box$\\
To prove Proposition \ref{estimdif} in the previous section, we need the following description, when the matrix $A_N$ and the measure $\nu$ satisfied Assumption A in the Introduction, of the convergence of $g_{\mu_{A_N}}(z)$ towards $g_\nu(z)$
and of the convergence of ${\cal Z}_{\frac{N}{p},\mu_{A_N}}(z)$ towards ${\cal Z}(z)$ (dealing in the last case with measures on $[0;+\infty[$) where ${\cal Z}_{\frac{N}{p},\mu_{A_N}}(z)$ is defined by (\ref{defzcnu}) replacing $\nu$ by $\mu_{A_N}$ and $c$ by $\frac{N}{p}$.
\begin{lemme}\label{cvAN} Under Assumption A,
there exists polynomials $P_1$ and $P_2$ with nonnegative coefficients and sequences $v_N(1)$ and $v_N(2)$ of positive numbers converging towards zero such that for all $z \in \mathbb{C}\setminus \mathbb{R}$,
\begin{equation}\label{hypA}\vert g_{\mu_{A_N}}(z) -g_\nu(z)\vert \leq P_1(\vert \Im z \vert^{-1}) v_N(1),\end{equation}
\begin{equation}\label{hypA2}\vert {\cal Z}_{\frac{N}{p},\mu_{A_N}}(z) -{\cal Z}(z)\vert \leq (\vert z \vert +1)^2 P_2(\vert \Im z \vert^{-1}) v_N(2),\end{equation}
\end{lemme}
\noindent{\bf Proof}: Let us introduce $$\hat{\nu}_N= \frac{1}{N-r}\sum_{j=1}^{N-r} \delta_{\beta_j(N)}.$$
Let us fix $\epsilon >0$. According to the assumption (\ref{univconv}), for $N$ large all the $\beta_j(N)$
are in the set $\{x, d(x,\mbox{supp~} \nu)<{\epsilon} \}$. Moreover, $\{x, d(x,\mbox{supp~} \nu)< {\epsilon}\}$ may be covered by a finite number $n_\epsilon$ of disjoint intervals $I_i(\epsilon)$ with diameter smaller than $\epsilon$, of the form $]a_i(\epsilon);b_i (\epsilon)]$ where $a_i(\epsilon)$ and $b_i (\epsilon)$ are two continuity points of the distribution function of $\nu$.
Note that for any $i=1,\ldots, n_\epsilon,$
when $N$ goes to infinity, $$\hat{\nu}_N (I_i(\epsilon)) \rightarrow \nu(I_i(\epsilon)).$$
Since $\vert \frac{1}{N}\sum_{i=1}^J \frac{1}{z-\theta_i}\vert \leq \frac{r}{N}\vert\Im z \vert^{-1}$,
and $\vert \left[\frac{1}{N}-\frac{1}{N-r}\right]\sum_{j=1}^{N-r} \frac{1}{z-\beta_j(N)}\vert \leq \frac{r}{N}\vert\Im z \vert^{-1}$,
we focus on the difference $g_{\hat{\nu}_N}(z) -g_\nu(z)$.
Similarly, since $\vert \frac{1}{N}\sum_{i=1}^J \frac{\theta_i}{1-\theta_i z}\vert \leq \frac{r}{N}\vert\Im z \vert^{-1}$,
and $\vert \left[\frac{1}{N}-\frac{1}{N-r}\right]\sum_{j=1}^{N-r} \frac{\beta_j(N)}{1-\beta_j(N)z}\vert \leq \frac{r}{N}\vert\Im z \vert^{-1}$, we focus on the difference
${\cal Z}_{\frac{N}{p},\hat{\nu}_N}(z) -{\cal Z}(z)$
where ${\cal Z}_{\frac{N}{p},\hat{\nu}_N}$ is defined by (\ref{defzcnu}) replacing $\nu$ by $\hat{\nu}_N$ and $c$ by $\frac{N}{p}$ .
\begin{eqnarray*}
g_{\hat{\nu}_N}(z) -g_\nu(z)&=& \sum_{i=1}^{n_\epsilon} \left\{ \frac{1}{N-r}\sum_{\beta_j(N) \in I_i(\epsilon)}\frac{1}{z-\beta_j(N)} - \int_{I_i(\epsilon)} \frac{1}{z-x} d\nu(x)\right\}\\
&=& \sum_{i,\nu(I_i(\epsilon))=0} \frac{1}{N-r}\sum_{\beta_j(N) \in I_i(\epsilon)}\frac{1}{z-\beta_j(N)}\\
&+&\sum_{i,\nu(I_i(\epsilon))>0} \frac{1}{N-r}\sum_{\beta_j(N) \in I_i(\epsilon)}\frac{1}{\nu(I_i(\epsilon))}
\int_{I_i(\epsilon)} \left( \frac{1}{z-\beta_j(N)} - \frac{1}{z-x}\right) d\nu(x)\\
&+& \sum_{i,\nu(I_i(\epsilon))>0} \left( \frac{\hat{\nu}_N (I_i(\epsilon))} {\nu(I_i(\epsilon))} - 1 \right)
\int_{I_i(\epsilon)} \frac{1}{z-x}d\nu(x)\\
&=& \Delta_1+\Delta_2+\Delta_3.
\end{eqnarray*}
where
$$\vert \Delta_1\vert \leq \sum_{i,\nu(I_i(\epsilon))=0} \hat{\nu}_N (I_i(\epsilon)) \vert \Im z \vert^{-1},$$
$$\vert \Delta_2\vert \leq \epsilon \sum_{i,\nu(I_i(\epsilon))>0} \hat{\nu}_N (I_i(\epsilon)) \vert \Im z \vert^{-2}\leq \epsilon \vert \Im z \vert^{-2},$$
$$\vert \Delta_3\vert \leq \sum_{i,\nu(I_i(\epsilon))>0} \vert \hat{\nu}_N (I_i(\epsilon))- {\nu} (I_i(\epsilon)) \vert \vert \Im z \vert^{-1}.$$
Hence $$\vert g_{\hat{\nu}_N}(z) -g_\nu(z)\vert \leq \left(\vert \Im z \vert^{-2}+ \vert \Im z \vert^{-1}\right)\left(\epsilon +\sum_{i=1}^{n_\epsilon} \vert \hat{\nu}_N (I_i(\epsilon))- {\nu} (I_i(\epsilon)) \vert\right)$$
and then $$\limsup_{N\rightarrow + \infty} \sup_{z\in \mathbb{C}\setminus \mathbb{R}} \left\{\left(\vert \Im z \vert^{-2}+\vert \Im z \vert^{-1}\right)^{-1}
\vert g_{\hat{\nu}_N}(z) -g_\nu(z)\vert\right\} \leq \epsilon.$$
Since this is true for any $\epsilon >0$, we get that
$$\lim_{N\rightarrow + \infty} \sup_{z\in \mathbb{C}\setminus \mathbb{R}} \left\{\left(\vert \Im z \vert^{-2}+ \vert \Im z \vert^{-1}\right)^{-1}
\vert g_{\hat{\nu}_N}(z) -g_\nu(z)\vert\right\}=0$$
which yields (\ref{hypA}).\\
Now, since moreover $\vert {\cal Z}_{\frac{N}{p},\hat{\nu}_N}(z)- {\cal Z}_{c,\hat{\nu}_N}(z)\vert \leq \vert \frac{N}{p}-c\vert
\vert \Im z \vert^{-1}$ we will study ${\cal Z}_{c,\hat{\nu}_N}(z)-{\cal Z}(z)$.
Similarly,
\begin{eqnarray*}
\frac{1}{c}[{\cal Z}_{c,\hat{\nu}_N}(z) -{\cal Z}(z)]&=& \sum_{i=1}^{n_\epsilon} \left\{ \frac{1}{N-r}\sum_{\beta_j(N) \in I_i(\epsilon)}\frac{\beta_j(N)}{1-\beta_j(N)z} - \int_{I_i(\epsilon)} \frac{x}{1-xz} d\nu(x)\right\}\\
&=& \sum_{i,\nu(I_i(\epsilon))=0} \frac{1}{N-r}\sum_{\beta_j(N) \in I_i(\epsilon)}\frac{\beta_j(N)}{1-\beta_j(N)z}\\
&+&\sum_{i,\nu(I_i(\epsilon))>0} \frac{1}{N-r}\sum_{\beta_j(N) \in I_i(\epsilon)}\frac{1}{\nu(I_i(\epsilon))}
\int_{I_i(\epsilon)} \left( \frac{\beta_j(N)}{1-\beta_j(N)z} - \frac{x}{1-xz}\right) d\nu(x)\\
&+& \sum_{i,\nu(I_i(\epsilon))>0} \left( \frac{\hat{\nu}_N (I_i(\epsilon))} {\nu(I_i(\epsilon))} - 1 \right)
\int_{I_i(\epsilon)} \frac{x}{1-xz}d\nu(x)\\
&=& \Delta_1+\Delta_2+\Delta_3.
\end{eqnarray*}
where
$$\vert \Delta_1\vert \leq \sum_{i,\nu(I_i(\epsilon))=0} \hat{\nu}_N (I_i(\epsilon)) \vert \Im z \vert^{-1},$$
\begin{eqnarray*}\vert \Delta_2\vert & \leq &\sum_{i,\nu(I_i(\epsilon))>0} \left\{ \frac{1}{N-r}\sum_{\beta_j(N) \in I_i(\epsilon)}\frac{1}{\nu(I_i(\epsilon))}
\int_{I_i(\epsilon)} \frac{\vert \beta_j(N)-x\vert }{\vert z\vert ^2(\vert \frac{1}{z}-\beta_j(N) \vert \vert \frac{1}{z}-x\vert } d\nu(x)\right\}\\&
\leq &
\epsilon \sum_{i,\nu(I_i(\epsilon))>0} \hat{\nu}_N (I_i(\epsilon)) \vert z\vert ^{-2} \left|\Im (\frac{1}{z}) \right|^{-2}\\
&\leq &\epsilon \vert z\vert ^2 \vert \Im z \vert^{-2},\end{eqnarray*}
$$\vert \Delta_3\vert \leq \sum_{i,\nu(I_i(\epsilon))>0} \vert \hat{\nu}_N (I_i(\epsilon))- {\nu} (I_i(\epsilon)) \vert \vert \Im z \vert^{-1}.$$
Hence $$\vert \frac{1}{c}[{\cal Z}_{c,\hat{\nu}_N}(z) -{\cal Z}(z)]\vert \leq \left( \vert z \vert^2 \vert \Im z \vert^{-2}+\vert \Im z \vert^{-1}\right)\left(\epsilon +\sum_{i=1}^{n_\epsilon} \vert \hat{\nu}_N (I_i(\epsilon))- {\nu} (I_i(\epsilon)) \vert\right)$$
and then $$\limsup_{N\rightarrow + \infty} \sup_{z\in \mathbb{C}\setminus \mathbb{R}} \left\{\left(\vert z \vert^2 \vert \Im z \vert^{-2}+\vert \Im z \vert^{-1}\right)^{-1}
\vert \frac{1}{c}[{\cal Z}_{c,\hat{\nu}_N}(z) -{\cal Z}(z)]\vert\right\} \leq \epsilon.$$
Since this is true for any $\epsilon >0$, we get that
$$\lim_{N\rightarrow + \infty} \sup_{z\in \mathbb{C}\setminus \mathbb{R}} \left\{\left( \vert z \vert^2 \vert \Im z \vert^{-2}+ \vert \Im z \vert^{-1}\right)^{-1}
\vert {\cal Z}_{c,\hat{\nu}_N}(z) -{\cal Z}(z) \vert\right\}=0$$
and (\ref{hypA2}) follows.
$\Box$\\
In the sample covariance matrix setting we will need the following upper bound of $\frac{1}{\Im (\underline{g}_N^S(z))}$
where $$\underline{g}_N^S(z)=\mathbb{E}\left( {\rm tr}_p G_{\underline{M}_N^S}(z)\right)=\frac{N}{p} \mathbb{E}\left( {\rm tr}_N G_{{M}_N^S}(z)\right)
+ \frac{1-\frac{N}{p}}{z},$$
with $$\underline{M}_N^S=\frac{1}{p} B_N^* A_N B_N.$$
\begin{lemme}\label{MajinversegN} There exists a constant $\underline{C}$ such that for any $z$ in $\mathbb{C}\setminus \mathbb{R}$,
$$\left|\frac{1}{\Im (\underline{g}_N^S(z))}\right| \leq \underline{C}(1+ \vert z \vert )^2 \vert \Im z \vert^{-1}$$
\end{lemme}
\noindent {\bf Proof}: Note that $$\vert \Im (\underline{g}_N^S(z)) \vert = \vert \Im z \vert \mathbb{E} \left[ \int_{\mathbb{R}} \frac{d \mu_{\underline{M}_N}(x)}{\vert z-x\vert^2} \right].$$
Now, for any $x$ in the spectrum of $\underline{M}_N^S$,
$$\vert z-x\vert^2 \leq 2 (\vert z \vert^2 + \Vert \underline{M}_N^S \Vert^2) \leq 2 (\vert z \vert^2 + \Vert A_N \Vert^2
\Vert \frac{1}{p} B_N^* B_N \Vert^2 )$$
\noindent so that, with $C= \sup_N \Vert A_N \Vert$,
$$\mathbb{E} \left[ \int_{\mathbb{R}} \frac{d \mu_
{\underline{M}_N}(x)}{\vert z-x\vert^2} \right] \geq \mathbb{E} \left[ \frac{1}{ 2 (\vert z \vert^2 + C^2
\Vert \frac{1}{p} B_N^* B_N \Vert^2 )} \right].$$
According to Theorem 5.11 in \cite{BaiSil06}, $\left\| \frac{1}{p} B_N^* B_N \right\| =c(1+ \frac{1}{\sqrt{c}})^2+o_{a.s,N}(1)$ so that by the dominated convergence Theorem we can deduce that for all large $N$,
$$\mathbb{E} \left[ \int_{\mathbb{R}} \frac{d \mu_
{\underline{M}_N}(x)}{\vert z-x\vert^2} \right] \geq \frac{1}{ 2 (\vert z \vert^2 + C_1
)} $$ \noindent where $C_1 > [C c(1+ \frac{1}{\sqrt{c}})^2]^2.$ Therefore
$$\left|\frac{1}{\Im (\underline{g}_N^S(z))}\right| \leq \frac{2 (\vert z \vert^2 + C_1
)}{\vert \Im z\vert }.\, $$
so that Lemma \ref{MajinversegN} readily follows. $\Box$
\section{Appendix}
\subsection{Poincar\'e inequality and concentration inequalities}
We first derive in this section concentration inequalities based on the Poincar\'e inequality.
We refer the reader to the book \cite{Tou}.
A probability measure $\mu$ on $\mathbb{R}$ is said to satisfy the Poincar\' e inequality with constant $C_{PI}$ if
for any
${\cal C}^1$ function $f: \mathbb {R}\rightarrow \mathbb {C}$ such that $f$ and
$f' $ are in $L^2(\mu)$,
$$\mathbf{V}(f)\leq C_{PI}\int \vert f' \vert^2 d\mu ,$$
\noindent with $\mathbf{V}(f) = \int \vert
f-\int f d\mu \vert^2 d\mu$. \\
We refer the reader to \cite{BobGot99} for a characterization
of the measures on $\mathbb{R}$ which satisfy a Poincar\'e inequality.
\begin{remarque}\label{multiple} If the law of a random variable $X$ satisfies the Poincar\'e inequality with constant $C_{PI}$ then, for any fixed $\alpha \neq 0$, the law of $\alpha X$ satisfies the Poincar\'e inequality with constant $\alpha^2 C_{PI}$.\\
If a probability measure $\mu$ on $\mathbb{R}$ satisfies the Poincar\'e inequality with constant $C_{PI}$ then the product measure $\mu^{\otimes M}$ on $\mathbb{R}^M$ satisfies the Poincar\'e inequality with constant $C_{PI}$ in the sense that for any differentiable function $F$ such that $F$ and its gradient $\nabla F$ are in $L^2(\mu^{\otimes M})$,
$$\mathbf{V}(f)\leq C_{PI} \int \Vert \nabla F \Vert_2 ^2 d\mu^{\otimes M}$$
\noindent with $\mathbf{V}(f) = \int \vert
f-\int f d\mu^{\otimes M} \vert^2 d\mu^{\otimes M}$ (see Theorem 2.5 in \cite{GuZe03}) .
\end{remarque}
An important consequence of the Poincar\'e inequality is the following concentration result.
\begin{lemme}\label{Herbst}{ Lemma 4.4.3 and Exercise 4.4.5 in \cite{AGZ09} or Chapter 3 in \cite{Ledoux01}. }
Let $\mathbb{P}$ be a probability measure on $\mathbb{R^M}$ which satisfies a Poincar\'e inequality with constant $C_{PI}$. Then there exists $K_1>0$ and $K_2>0$ such that, for any Lipschitz function $F$ on $\mathbb{R}^M$ with Lipschitz constant $\vert F \vert_{Lip}$,
$$\forall \epsilon> 0, \, \mathbb{P}\left( \vert F-\mathbb {E}_{\mathbb{P}}(F) \vert > \epsilon \right) \leq K_1 \exp\left(-\frac{\epsilon}{K_2 \sqrt{C_{PI}} \vert F \vert_{Lip}}\right).$$
\end{lemme}
\subsection{Technical tools}
We need the following result on the extension of Lipschitz functions on $\mathbb{R}$ to the Hermitian matrices.
\begin{lemme}\label{extlipschitz}(see \cite{D})
Let $f$ be a real $C_{\cal L}$-Lipschitz function on $\mathbb{R}$. Then its extension on the $N\times N$
Hermitian matrices is $C_{\cal L}$-Lipschitz with respect to the norm $\Vert M\Vert_2=\{Tr (MM^*)\}^{\frac{1}{2}}$.
\end{lemme}
\noindent {\bf Proof:} Let $A$ and $B$ be $N\times N$
Hermitian matrices. Let us consider their spectral decompositions
$$A=\sum_i \lambda_i(A)P^{(A)}_i$$ \noindent and
$$B=\sum_i \lambda_i(B)P^{(B)}_i.$$ \noindent We have
\begin{eqnarray*}
\Vert f(B) -f(A) \Vert^2_2 &=& Tr \left( \sum_i f(\lambda_i(A))P^{(A)}_i -\sum_i f(\lambda_i(B))P^{(B)}_i \right)^2\\
&=&Tr \left( \sum_{i} f(\lambda_i(A))^2P^{(A)}_i +\sum_{j} f(\lambda_j(B))^2P^{(B)}_j \right)
\\ & &-2 \sum_{i,j} f(\lambda_i(A))f(\lambda_j(B))Tr (P^{(A)}_i P^{(B)}_j)\\
&=&Tr \left( \sum_{ij} (f(\lambda_i(A))^2P^{(A)}_i P^{(B)}_j+\sum_{i,j} (f(\lambda_j(B))^2P^{(A)}_i P^{(B)}_j \right)
\\ & & -2 \sum_{i,j} f(\lambda_i(A))f(\lambda_j(B))Tr (P^{(A)}_i P^{(B)}_j)\\
&=& \sum_{i,j} (f(\lambda_i(A))-f(\lambda_j(B))^2Tr( P^{(A)}_i P^{(B)}_j).
\end{eqnarray*}
Now, since $Tr( P^{(A)}_i P^{(B)}_j)\geq 0$, we can deduce that
$$\Vert f(B) -f(A) \Vert_2^2 \leq \sum_{i,j} C_{\cal L}^2(\lambda_i(A)-\lambda_j(B))^2Tr( P^{(A)}_i P^{(B)}_j)= C_{\cal L}^2 \Vert B-A \Vert_2^2. \Box$$
We recall here some useful properties of the resolvent (see \cite{KKP96,CD07}).
\begin{lemme} \label{lem0}
For a $N \times N$ Hermitian or symmetric matrix $M$,
for any $z \in \mathbb {C}\setminus {\rm Spect}(M)$,
we denote by $G(z) := (zI_N-M)^{-1}$ the resolvent of $M$.\\
Let $z \in \mathbb {C}\setminus \mathbb {R}$,
\begin{itemize}
\item[(i)] $\Vert G(z) \Vert \leq |\Im z|^{-1}$ where $\Vert . \Vert$ denotes the operator norm.
\item[(ii)] $\vert G(z)_{ij} \vert \leq |\Im z|^{-1}$ for all $i,j = 1, \ldots , N$.
\item[(iii)] Let $z \in \mathbb {C}$ such that $|z| > \Vert M \Vert$; we have
$$\Vert G(z) \Vert \leq \frac{1}{|z| - \Vert M \Vert}.$$
\end{itemize}
\end{lemme}
We recall here the following classical result due to
Weyl.
\begin{lemme}{(cf. Theorem 4.3.7 of \cite{HJ})} \label{Weyl}
Let B and C be two $N \times N$ Hermitian matrices. For any pair of integers
$j,k$ such that $1 \leq j,k\leq N$ and $j+k \leq N+1$, we have
$$\lambda_{j+k-1} (B+C) \leq \lambda_{j}(B) + \lambda_{k}(C).$$
For any pair of integers
$j,k$ such that $1 \leq j,k\leq N$ and $j+k \geq N+1$, we have
$$ \lambda_j(B) + \lambda_k(C) \leq \lambda_{j+k-N} (B+C).$$
\end{lemme}
The following result on quadratic forms is of basic use in the sample covariance matrix setting. Note that, a complex random variable $x$ will be said {\it standardized} if $\mathbb{E}(x)=0$ and $\mathbb{E}(\vert x \vert^2)=1$.
\begin{proposition}\label{QF}(Lemma 2.7 \cite{BaiSil98})
Let $B=(b_{ij})$ be a $N \times N$ matrix and
$Y_N$ be a vector of size $N$ which contains i.i.d
standardized entries with bounded fourth
moment. Then there is a constant $K>0$ such that $$\mathbb E\vert Y_N ^* B Y_N - {\rm{Tr}} B\vert^2 \leq K \mathbb {\rm Tr} (BB^*).$$
\end{proposition}
The following technical lemma is fundamental in this paper.
We refer the reader to the Appendix of \cite{CD07}
where it is proved using the ideas of \cite{HaaThor05}.
\begin{lemme}\label{HT} Let $h$ be an analytic function on $\mathbb {C}\setminus \mathbb {R}$ which satisfies
\begin{equation*}\label{nestimgdif}
\vert h(z)\vert \leq (\vert z\vert +K)^\alpha P(\vert \Im z\vert ^{-1})
\end{equation*} and $\varphi$ be in $\cal C^\infty (\mathbb {R}, \mathbb {R})$ with compact support. Then,
$$\limsup _{y\rightarrow 0^+}\vert \int _\mathbb {R}\varphi (x)h(x+iy)dx\vert < + \infty.$$
\end{lemme}
\noindent {\bf Acknowledgments:}
I am grateful to Charles Bordenave
for useful discussions. I would like to thank the anonymous referees for their careful reading and their pertinent comments which led to an overall improvement of the paper.
\def$'${$'$}
|
\section{Introduction}
Studies on particles of a few nanometers in size have been attracting
quite a lot of interest of late. For a physicist the attraction is
mainly due to the emergent properties that a particle or a collection of
them shows when the particle size is made very small. According to
N\'{e}el tiny particles of an antiferromagnetic material should exhibit
magnetic properties such as superparamagnetism and weak ferromagnetism
\cite{Low Temp. Phys.}. If the surface to volume ratio, which varies as
the reciprocal of particle size, of an antiferromagnetic particle
is made sufficiently large then it can have a nonzero net magnetic
moment because of an imperfect cancellation of elementary moments
pointing in different directions near the surface of the particle.
Recently nanoparticles of antiferromagnetic materials have gained quite
a lot of attention mainly because they show some surprising and unusual
behavior unobservable in ferro or ferrimagnetic nanoparticles
\cite{Kodama 1997, Morup 2004}. Among different antiferromagnetic
nanoparticles, NiO is a comparatively more interesting and well studied
system \cite {Kodama 1997, Richardson and Milligan, Richardson 1991,
Makhlouf 1997}. Long back, Richardson and Milligan \cite { Richardson
and Milligan} reported magnetization measurements on NiO nanoparticles
of different sizes and this has been followed by many more reports
\cite{Kodama 1997, Richardson 1991, Makhlouf 1997}, mostly by other
workers. We also reported some work on NiO nanoparticles \cite{SDT 2005,
SDT 2006} where we found that, at low temperatures, this system shows
spin glass behavior. We also showed that the low temperature behavior of
NiO nanoparticles is not superparamagnetic contrary to popular belief
and expectation. After our work some other workers have also reached similar
conclusions on NiO nanoparticles independently \cite{Winkler, Vijay}.
The magnetization of antiferromagnetic nanoparticles is expected to be
described by a modified Langevin function \cite{Makhlouf 1997, Kilcoyne,
Makhlouf, Seehra 2000}. However, fitting the magnetization data of bare
antiferromagnetic NiO nanoparticles as a function of magnetic field to
the modified Langevin function results in unphysical fit parameters.
For instance, the estimated particle magnetic moment turns out to be
about 2000~$\mu_{\textnormal{B}}$ for 5.3~nm particles \cite {Makhlouf 1997}. Such values
for the particle magnetic moments are much larger than the expected value of about a few hundred Bohr magnetons from uncompensated spins on the surface of particles. Another method for the estimation of
particle magnetic moment of NiO nanoparticles has also been used
\cite{Kodama 1997}. Here the value of saturation magnetization at low
temperature is divided by the estimated number of particles to get the
particle magnetic moment. However the moment estimated
by this method has also been found to be larger than the expected value. A multisublattice model has been proposed to explain this large value of
particle magnetic moment for NiO nanoparticles though there is no experimental support for this model yet \cite{Kodama 1997}.
The magnetization of nanoparticles of magnetic materials is expected to be
only a function of the applied magnetic field $B$ and temperature $T$
and should scale with $(\frac{B}{T})$ above the bifurcation temperature ($T_{\textnormal{bf}}$) between low field cooled and zero field cooled magnetization. The magnetization of NiO
nanoparticles is not found to show this scaling \cite{Makhlouf 1997}.
These observations motivated us to revisit the antiferromagnetic NiO
nanoparticles system once again. This work is an attempt to find the
reason for getting unphysical numbers for the particle magnetic moment of
NiO nanoparticles when the traditional methods are used to analyze the
magnetization data above $T_{\textnormal{bf}}$. The value of the bifurcation temperature $T_{\textnormal{bf}}$ for the present system is about 295~K in 100~G applied magnetic field \cite{SDT 2006}.
Here we present the magnetization measurements on 5~nm bare NiO
particles as a function of applied magnetic field at
sufficiently high temperatures, but well below
the N\'{e}el temperature ($T_{\textnormal{N}}$) of the system which is known to be about
523~K \cite{Smart}. We analyzed the data using the modified
Langevin function without considering any distribution in
particle magnetic moment and then repeated the analysis taking into account a distribution in
the particle moment. We were pleasantly surprised by the results as it turned out that we can account for the anomalous observations reported on this system by the earlier workers.
NiO nanoparticles were prepared by a sol-gel method by reacting in
aqueous solution nickel nitrate and sodium hydroxide at pH = 12 at
room temperature as described elsewhere \cite{Richardson and
Milligan, Makhlouf 1997, Richardson 1991}. We used nickel (II)
nitrate hexahydrate (99.999\%), sodium hydroxide pellets
(99.99\%), both from Aldrich, and triple distilled water to make
nickel hydroxide. The sample of nickel oxide nanoparticles was
prepared by heating the nickel hydroxide at 523~K for 3 hours in
flowing helium gas (99.995\%). The sample was characterized by
x-ray diffraction and transmission electron microscopy. The
average crystallite size as well as the particle size were found to be
about 5~nm. The details of sample synthesis and structural
characterization have been reported in one of our earlier works \cite{SDT
2006}. All the magnetic measurements were done with a commercial SQUID
magnetometer (Quantum Design, MPMS XL5).
\begin{figure}[thb]
\begin{center}
\includegraphics[angle=0,width=1.\columnwidth]{fig1.eps}
\caption{(Color online) Magnetization as a function of applied magnetic field for 5~nm NiO particles at different temperatures. Solid lines show fits to Equation(\ref{eq:lang-modi}). The inset shows a magnified view of the data along with the fits at lower magnetic fields which makes it clear that the fit quality is poor there.}
\label{fig:mh-5nm}
\end{center}
\end{figure}
We measured the magnetization $M$ of the 5~nm NiO particles as a
function of external applied magnetic field $B$ at different
temperatures $T$ ($T_{\textnormal{bf}} < T < T_{\textnormal{N}}$). These measurements are shown in Figure
\ref{fig:mh-5nm}. Small particles of magnetic materials
are expected to be superparamagnetic at sufficiently high temperature ($T > T_{\textnormal{bf}}$)\cite{Low Temp. Phys.}.
Magnetization $M$ of a superparamagnetic material as a function of magnetic field $B$ and temperature $T$ is described by \cite{Bean}
\begin{equation}
\label{eq:lang}
M = M_0 L(x),
\end{equation}
where $L(x) = [\coth{(x)} - (\frac{1}{x})]$ is the Langevin
function and $x = \frac{\mu_{\textnormal{p}} B}{k_{\textnormal{B}} T}$. Here $M_0$ is the
saturation magnetization, $\mu_{\textnormal{p}}$ is the particle magnetic moment
and $k_{\textnormal{B}}$ is the Boltzmann constant. However, magnetization of small
particles of antiferromagnetic materials, even though they are
superparamagnetic cannot be described by
Equation~(\ref{eq:lang}). Rather, they are well described by an
altered form known as the
modified Langevin function \cite{Kilcoyne, Makhlouf,Seehra 2000},
which contains an extra linear term in $B$ i.e.
\begin{equation}
\label{eq:lang-modi}
M = M_0 L(x) + \chi_{\textnormal{a}} B,
\end{equation}
where $\chi_{\textnormal{a}}$ is the susceptibility of randomly oriented
antiferromagnetic particle cores. We fitted the data shown in Figure~\ref{fig:mh-5nm} to the above equation and the resulting fits are shown as solid lines in the figure. Values of fit
parameters $M_{0}$, $\mu_{\textnormal{p}}$ and $\chi_{\textnormal{a}}$ obtained are
presented in Table~\ref{table-310-350K-5nm}.
\begin{table}[htb]
\caption{Values of fit parameters $M_{0}$, $\mu_{\textnormal{p}}$ and
$\chi_{\textnormal{a}}$ to Equation (\ref{eq:lang-modi}) and the corresponding values of the goodness of fit parameter
$R^{2}$ for the 5~nm NiO nanoparticles at different temperatures.}
\label{table-310-350K-5nm} \vspace{0.2in}
\begin{tabular}{|c|c|c|c|c|}
\hline
$T$&$M_{\textnormal{0}}$&$\mu_{\textnormal{p}}$&$\chi_{\textnormal{a}}$ ($10^{-6}$&$R^{2}$\\
(K)&(emu/g)&($\mu_{\textnormal{B}}$)&emu/g Oe)&\\
\hline
310&1.30&1967&34&0.9993\\
320&1.18&1841&32&0.9994\\
330&1.05&1825&31&0.9995\\
340&0.94&1734&29&0.9996\\
350&0.82&1621&27&0.9997\\
\hline
\end{tabular}
\end{table}
From Figure \ref{fig:mh-5nm} we see that the fits
to Equation (\ref{eq:lang-modi}) more or less pass
through the measured data points which means the fit quality is quite
good. A statistical measure of the goodness
of a fit is the coefficient of determination $R^{2}$. The closer
$R^2$ is to unity the better the fit. Values of the coefficient of
determination $R^{2}$ for the fits
are shown in Table \ref{table-310-350K-5nm} along with the fit
parameters. The $R^2$ values are greater than 0.999 in all cases and this
once again confirms the good quality of the fits.
From Table
\ref{table-310-350K-5nm} we see that the particle magnetic moment
$\mu_{\textnormal{p}}$ is about two thousand Bohr magnetons and it decreases
with increasing temperature. Somewhat similar results have been
reported by others on NiO nanoparticles \cite{Makhlouf 1997} as well as
on NiO nanorods \cite{Seehra 2004}. However there is a glaring inconsistency between
the numbers we get for $M_0$ and $\mu_{\textnormal{p}}$. In
Equation (\ref{eq:lang-modi}) the fit parameters $M_0$ and $\mu_{\textnormal{p}}$ should be
related as $M_0 = N \mu_{\textnormal{p}}$ where $N$ is the number of particles per unit mass
of the sample. If we use this relation to estimate the average particle
magnetic moment $\mu_{\textnormal{p}}$ it turns out to be about 60~$\mu_{\textnormal{B}}$ at 320~K which
is very small compared to the value of $\mu_{\textnormal{p}}$ presented in Table
\ref{table-310-350K-5nm}. This fact invalidates the fit of experimental data to Equation~(\ref{eq:lang-modi}). In the next paragraph we shall give another argument against the numbers presented in Table~\ref{table-310-350K-5nm}.
N\'{e}el had discussed various ways by which a magnetic moment can appear on an NiO particles due to incomplete compensation of atomic moments in different sublattices \cite{Low Temp. Phys., Richardson 1991}. According to N\'{e}el the
particle magnetic moment of an NiO nanoparticle can be written as
\begin{equation}
\label{eq:neil-moment}
\mu_{\textnormal{p}} = p~\mu_{\textnormal{A}}~\mu_{\textnormal{B}}.
\end{equation}
Here $p$ is a number which depends on the size, crystal structure and form or shape of the particle and $\mu_{\textnormal{A}}$ is the magnetic moment of Ni$^{2+}$ ion which is known to be 3.2~$\mu_{\textnormal{B}}$ \cite{Kittel}.
N\'{e}el proposed a total of five different possibilities for the
origin of magnetic moment on an NiO particle of approximately cubic shape. Out of the five, two
possibilities give the values of particle magnetic moment to be
zero corresponding to $p=0$. The other three
possible values of $p$ are about $n^{\frac{1}{3}}$, $n^{\frac{1}{2}}$
and $n^{\frac{2}{3}}$ where $n$ is the number of Ni$^{2+}$ ions in the
particle. Now, there can be many other possible set of values of $p$ corresponding to shapes such as spheres, tetrahedrons or, more realistically, irregular shapes. But it seems safe to assume that the upper limit of $p$ will be about $n^{\frac{2}{3}}$ and the average will be considerably less.
Making use of the fact that the crystal structure of NiO is
face centered cubic with lattice constant 4.176~$\AA$, the value
of $n$ for a 5~nm diameter spherical particle turns out to be about 3592. Using the information presented above along with Equation~(\ref{eq:neil-moment})
we get the particle magnetic moment to be about 49~$\mu_{\textnormal{B}}$,
191~$\mu_{\textnormal{B}}$ and 750~$\mu_{\textnormal{B}}$ corresponding to $p$ values of
$n^{\frac{1}{3}}$, $n^{\frac{1}{2}}$ and $n^{\frac{2}{3}}$
respectively.
Thus according to N\'{e}el's picture the maximum possible
value of particle magnetic moment for a 5~nm NiO particle would be about
750~$\mu_{\textnormal{B}}$ which is much smaller than the values shown in Table \ref{table-310-350K-5nm}.
While using Equation (\ref{eq:lang-modi}) to fit the magnetization
data we made a tacit assumption that all the particles have the same magnetic
moment. But, this is not true. A sample of NiO
nanoparticles has a distribution of particle magnetic moments not
only due to distribution in size and shape but also due to the way in which the imbalance of spins arises in the particle as pointed out by N\'{e}el. We hazard the guess that disregarding the distribution in particle moments is perhaps the reason for the unphysical fit parameters obtained in Table
\ref{table-310-350K-5nm}. Now we would like to carry out this fitting taking into account a moment distribution. Making our path easier is a precedent in this kind of analysis set by Silva et. al. in analyzing the magnetization of ferritin, a biological antiferromagnetic nanoparticle system \cite{Silva}.
In a sample that has a distribution in particle magnetic
moments the low field magnetization is governed by the particles
with larger magnetic moments. The contribution of particles of lower
magnetic moments to the magnetization becomes important only at higher applied
fields where the high field forces the moments to align with the field. The
effect of a distribution in the particle magnetic moment on the
magnetization of the system will show up in the Langevin or modified Langevin fit. To see this dependence let us take a look at the comparison of the measured magnetization to the modified Langevin fit at low fields in the inset of Figure
\ref{fig:mh-5nm}. It is clear that the fits are no good at
low fields while from the main panel one can see that the situation is much better
at higher fields. This low field misfit is an indication of the role
a distribution in the particle magnetic moment plays on the
magnetization of a system. In this case the modified Langevin function has clearly underestimated the contribution of the larger moments.
Following Silva et. al. we assume that in a sample of nanoparticles the distribution in particle magnetic
moment, $\mu$, can be described by a log normal distribution function of the form \cite{Silva}
\begin{equation}
\label{eq:lognormal-distribution}
f(\mu) = \frac{1}{\mu s \sqrt{2 \pi}} \exp -\frac{[{\ln (\frac{\mu}{n})]}^2}{2 s^2},
\end{equation}
where $n$ and $s$ are parameters that characterize the distribution.
The mean particle magnetic moment $\mu_{\textnormal{mean}}$ is equal to $n
\sqrt{e^{s^2}}$. Now, Equation (\ref{eq:lang-modi}) takes on the new form
\begin{equation}
\label{eq:lang-modi-lognormal} M (B, T) = N \int_{0}^{\infty} \mu
L(x) f(\mu) d\mu + \chi_a B,
\end{equation}
where $N$ is the total number of particles contributing to
magnetization and $L(x)$ is the Langevin function with $x = \frac{\mu B}{k_{\textnormal{B}} T}$. We used this
equation to fit the magnetization data of Figure
\ref{fig:mh-5nm} and the fit parameters thus obtained along
with the values of coefficient of determination $R^2$ are shown in
Table \ref{table-310-350K-5nm-lognormal}.
\begin{table}
\caption{Values of fit parameters $N$, $s$, $n$ and $\chi_{a}$ to
Equation (\ref{eq:lang-modi-lognormal}) and the values of $R^{2}$
for the 5~nm NiO particles at different temperatures. The mean magnetic moment of a particle estimated from the fit parameters are also shown.}
\label{table-310-350K-5nm-lognormal} \vspace{0.2in}
\begin{tabular}{|c|c|c|c|c|c|c|}
\hline
$T$&$N$ ($10^{17}$&$s$&$n$&$\chi_a (10^{-6}$& $\mu_{\textnormal{mean}}$ &$R^{2}$\\
(K)&/g)&&$(\mu_\textnormal{B})$&emu/g~Oe)& $(\mu_\textnormal{B})$ &\\
\hline
310&8.7&1.34&116.5&21.4&287&0.999997\\
320&7.3&1.30&130.3&21.4&305& 0.999998\\
330&6.5&1.27&136.8&20.9&308&0.999997\\
340&6.9&1.28&118.9&19.8&271&0.999998\\
350&6.8&1.26&114.4&18.9&254&0.999998\\
\hline
\end{tabular}
\end{table}
We see in Table \ref{table-310-350K-5nm-lognormal} that the
values of the coefficient of determination $R^2$ are surprisingly
greater than 0.99999 in all the cases. These values of $R^2$ are much better than the values of the same shown in Table
\ref{table-310-350K-5nm}. This means that the quality of the fits
using Equation (\ref{eq:lang-modi-lognormal}) is much better than
that using Equation (\ref{eq:lang-modi}). In Figure
\ref{fig:mh-5nm-modi-lang-fit-distri} we show the new fits which clearly look good. The inset clearly shows that the
fits are very good for lower values of applied magnetic fields
also where the effect of distribution in the particle magnetic
moment on the magnetization of the system is important as already
discussed. We can now conclude that the fits of the magnetization data to Equation
(\ref{eq:lang-modi-lognormal}) are much better than that to
Equation (\ref{eq:lang-modi}).
\begin{figure}[thb]
\begin{center}
\includegraphics[angle=0,width=1.0\columnwidth]{fig2.eps}
\caption{(Color online) Magnetization as a function of applied magnetic field for 5~nm NiO particles at different temperatures. Solid lines show the fits of data to Equation~(\ref{eq:lang-modi-lognormal}). The inset shows the fits at low magnetic fields.}
\label{fig:mh-5nm-modi-lang-fit-distri}
\end{center}
\end{figure}
From Table \ref{table-310-350K-5nm-lognormal} we find that the mean particle magnetic moments, which are equal to $n\sqrt{e^{s^2}}$, turns out to be about a few hundred Bohr magnetons, rather than a few thousand Bohr magnetons as found in the earlier fit, and are also well below the maximum possible value of about 750~$\mu_{\textnormal{B}}$ according to the models proposed by N\'{e}el for the origin of magnetic moment in NiO particles.
In conclusion, we reported magnetization as a function of applied
magnetic field for 5~nm NiO particles at different temperatures above
the bifurcation temperature $T_{\textnormal{bf}}$. Fitting the magnetization data to the modified
Langevin function without considering any distribution in the particle
magnetic moments yields very large and unphysical values for the
particle magnetic moment. However we got reasonable values for the particle
magnetic moment if the modified Langevin function is used to fit the
magnetization data taking into account a distribution in the particle
magnetic moment. The distribution in the particle magnetic moment arises
from a distribution in particle size and form.
This work clearly shows that the non consideration of
a distribution in particle magnetic moment could be the reason for the
anomalously large values of magnetic moment of NiO nanoparticles
reported in the literature.
|
\section{Introduction}
Nigel Kalton was one of the greatest mathematicians of the last
40~years, although he did his best to conceal this fact. An outsider
wouldn't recognise the mathematical giant that he was in this modest
person who was always friendly and good-humoured and who was more than
willing to share his ideas with everyone.
Nigel published more than 260 papers (including several books) not
only in Banach and quasi-Banach space theory, but also in so diverse
fields such as game theory, continued fractions, harmonic analysis,
operator semigroups and convex geometry.
Every single of these papers contains a deep contribution by Nigel,
often taking care of the most difficult case that his coauthors would
have to leave open without his help. He had a wide interest in
mathematics, and his problem solving abilities were legendary. For
instance, once after a colloquium talk on continued fractions he got
hooked on the subject and redeveloped the theory for himself over one
weekend, eventually solving the problem exposed in the talk.
I met Nigel for the first time at the conference on Banach spaces in
Mons in 1987. It so happened that on the day after the conference we
were both waiting for the same train to Paris, but not for the same
coach: he told me that he always rides the first class, adding, ``I'm
snobbish.'' Of course he couldn't be more wrong in his
self-assessment! Some years later he solved a big problem in $M$-ideal
theory (see Section~\ref{sec2}), a problem, where we, a group of fresh
Ph.D.s in Berlin, couldn't get anywhere. He emailed me a file with his
solution, and this was the beginning of our collaboration, in which
more often than not I felt like a pedestrian next to a racing-car.
In June 2010, after a talk at the conference in Valencia with a
somewhat set-theoretical flavour, I reminded him of the quote from
Star Trek, ``It's mathematics, but not as we know it.'' I knew that
this would strike his sense of humour; I did not know that this would
be the last time I saw him.
In the next few sections I will try to survey some of Nigel's
contributions to Banach space theory. I will restrict myself to
problems of an isometric nature, but even this narrower area is still
so rich that omissions and misconceptions will be
inevitable. Certainly, the only way to do justice to Nigel's genius
would be to not only paraphrase the main results, but to expound all
the ideas contained in his papers. I have to leave this to an abler
mathematician.
In the following, papers by Nigel will be cited in the form \kcite{Kal}
and other papers in the form \cite{LinPel2}. The bibliography will
first list Nigel's
cited papers chronologically, and then the other papers
alphabetically.
\section{$M$-ideals} \label{sec2}
An \textit{$M$-ideal} $V$ of a Banach space $E$ is a closed subspace such that
the dual admits an $\ell_1$-direct decomposition $E^*= V^\bot \oplus_1
W$ for some closed subspace $W\subset E^*$. (In other words, $V^\bot$
is an $L$-summand of $E^*$.) The notion of an $M$-ideal was
introduced by E.M.~Alfsen and E.G.~Effros \cite{AlEf}; for a detailed study
one may consult \cite{HWW}. Of course, $c_0$ is an $M$-ideal in
$\ell_\infty$, and Dixmier proved back in 1951 that $K(H)$, the space
of compact operators on a Hilbert space $H$, is an $M$-ideal in
$L(H)$, the space of bounded operators. By the end of the 1980s more
examples of Banach spaces $X$ for which $K(X)$ is an $M$-ideal in
$L(X)$ were known; basically, these examples were $\ell_p$-sums
(${p>1}$) or $c_0$-sums of finite-dimensional spaces and certain of
their subspaces and quotients. On the other hand, several necessary
conditions were known: $X$ has to be an $M$-embedded space (i.e., $X$
is an $M$-ideal when canonically embedded into $X^{**}$), which
implies for example that $X^*$ has the RNP, and $X$ must have
the metric compact approximation property; this was proved by
P.~Harmand and \AA.~Lima \cite{HaLi}.
For subspaces of $X\subset \ell_p$ the converse
was proved by C.-M.~Cho and W.B.~Johnson \cite{ChJo1}: the metric compact
approximation property is sufficient for $K(X)$ to be an $M$-ideal in
$L(X)$. (Later Nigel obtained this in complete generality, see
Corollary~\ref{cor2.1.a}.)
Let us recall what this approximation property means.
A Banach space $X$ has the \textit{metric compact approximation property} if
there is a net of compact operators $K_i\dopu X\to X$ of norm ${\le1}$
that converges pointwise to the identity: $K_ix\to x$ for all $x\in
X$. Actually, an even stronger approximation property holds if $K(X)$
is an $M$-ideal in $L(X)$: one can achieve that $\limsup \|\Id
-2K_i\|\le1$ and both $K_i\to \Id$ and $K_i^*\to \Id$ pointwise
(\textit{shrinking unconditional metric compact approximation
property}). Here the \textit{shrinking} bit derives from the fact
that also $K_i^*x^*\to x^*$ for all $x^*$, like for the projections
associated with a shrinking basis, and the \textit{unconditionality}
is hidden in
the norm condition $\limsup \|\Id -2K_i\|\le1$; see the beginning of
Section~\ref{sec4}.
Although the $M$-ideal problem for $K(X)$ was intensively studied in
the 1980s, there were no conditions on $X$ known that were necessary and
sufficient for $K(X)$ to be an $M$-ideal in $L(X)$; the best result by
then was given by W.~Werner \cite{WW1}:
$K(X)$ is an $M$-ideal in $L(X)$ if and
only if $X$ has the metric compact approximation property by means of
a net $(K_i)$ satisfying
$$
\limsup \|SK_i + T(\Id-K_i)\|\le \max\{\|S\|,\|T\|\} \qquad\forall
S,T\in L(X).
$$
This was a big achievement, but still the condition is so complicated
that one cannot check easily that it is fulfilled for $X=\ell_2$.
This was the moment when Nigel got interested in the problem, probably
after some eventually successful bugging by Gilles Godefroy. (It
should be added that in the 1980s people working on $M$-ideals of
compact operators often used techniques from Nigel's much quoted paper
\kcite{Kal}.\footnote{It got quoted 47~times according to the Mathematical
Reviews database;
but Nigel once told me that it was accepted for publication only at
the third attempt.})
In his paper \kcite{Kal-M} he offered an entirely new approach based on what he
called property~$(M)$ and property~$(M^*)$. Here are the definitions.
A Banach space $X$ has \textit{property~$(M)$} if whenever $(x_i)$ is a
bounded weakly null net and $x,y\in X$ satisfy $\|x\|=\|y\|$, then
$$
\limsup \|x_i+x\| = \limsup \|x_i+y\|, \eqno(1)
$$
and $X$ has \textit{property~$(M^*)$} if whenever $(x_i^*)$ is a
bounded weak$^*$ null net and $x^*,y^*\in X^*$ satisfy $\|x^*\|=\|y^*\|$, then
$$
\limsup \|x_i^*+x^*\| = \limsup \|x_i^*+y^*\|. \eqno(2)
$$
These notions can be recast in the language of types on Banach spaces;
see Section~\ref{sec6} below.
It is easy to see that $(M^*)$ implies $(M)$; for the converse see
the remarks following Theorem~\ref{theo2.2}.
Nigel's theorem is as follows.
\begin{theo}\label{theo2.1}
The following assertions about a Banach space $X$ are equivalent:
\begaeq
\item
$K(X)$ is an $M$-ideal in $L(X)$.
\item
$X$ has
property~$(M)$, does not contain a copy of $\ell_1$ and has the
unconditional metric compact approximation property, i.e., there is a
net of compact operators satisfying $K_ix\to x$ for all $x$ and
$\limsup\|\Id-2K_i\|\le1$.
\item
$X$ has property~$(M^*)$ and has the
unconditional metric compact approximation property.
\endaeq
\end{theo}
Actually, in his paper Nigel only deals with the case of separable
spaces and the sequential versions of $(M)$ and $(M^*)$, but the
extension to the general case doesn't offer any difficulties; it can
be found in \cite[page~299]{HWW} for example. One should remark that
the sequential
property~$(M)$ is trivially satisfied in Schur spaces (where by
definition weakly convergent sequences are norm convergent), but these
are excluded by the requirement that $\ell_1$ does not embed into $X$
in Theorem~\ref{theo2.1}.
Let me point out that it is trivial to verify that the $\ell_p$-spaces
for $1<p<\infty$ and in particular Hilbert spaces satisfy Nigel's
conditions; to see that $\ell_p$ has $(M)$ just note that for $\ell_p$
we have
$$
\limsup \|x+x_i\|^p = \|x\|^p + \limsup\|x_i\|^p
\eqno(3)
$$
so that (1) becomes obvious.
Using property~$(M)$ Nigel has been able to solve a number of open
problems in the theory of $M$-ideals. For example, he proved a general
version of the theorem of Cho and Johnson mentioned above:
\begin{cor}\label{cor2.1.a}
If $K(X)$ is an $M$-ideal in $L(X)$ and $E\subset X$ has the
compact metric approximation property, then $K(E)$ is an $M$-ideal
in $L(E)$ as well.
\end{cor}
He also showed that Orlicz sequence spaces and more generally modular
sequence spaces can be renormed to have property~$(M)$; if in addition
such a space $X$ has a separable dual, then $X$ can be renormed so
that $K(X)$ becomes an $M$-ideal in $L(X)$. In the other direction, a
separable non-atomic order continuous Banach lattice can be renormed
to have property~$(M)$ if and only if it is lattice isomorphic
to~$L_2$.
Another interesting corollary is that spaces $X$ with property~$(M)$ contain
subspaces isomorphic to $\ell_p$; more precisely, there exists $1\le
p<\infty$ so that $\ell_p$ embeds almost isometrically into $X$ (cf.\
Section~\ref{sec3} for this concept) or $c_0$ embeds almost
isometrically into $X$. The proof relies on a deep theorem due to
J.-L.~Krivine \cite{Kri}; a particular consequence is the theorem, originally
obtained by J.~Lindenstrauss and L.~Tzafriri, that an
ifinite-dimensional subspace of an Orlicz sequence space $h_M$
contains a copy of some $\ell_p$ or of $c_0$.
Concerning $L_p=L_p[0,1]$ it was known by 1990 that $K(L_p)$ is not an
$M$-ideal in $L(L_p)$ for $p\neq2$, and conversely that an $M$-ideal
in $L(L_p)$ is necessarily a two-sided ideal for $1<p<\infty$; indeed
this is so since $L_p$ and its dual are uniformly convex \cite{ChJo2}. But in
our joint paper \kcite{KalW1}
it was shown that, for $p\neq1,2,\infty$, there are
no nontrivial $M$-ideals in $L(L_p)$ whatsoever (nontrivial meaning
different from $\{0\}$ and the whole space), and if $1<p,q<\infty$
then $K(\ell_p(\ell_q^n))$ is the only nontrivial $M$-ideal in
$L(\ell_p(\ell_q^n))$. In both these results we had to assume complex
scalars; the proofs use arguments involving hermitian operators.
In the paper \kcite{AndCazKal}, with coauthors G.~Androulakis and
C.D.~Cazacu, Nigel
takes his construction of spaces with property~$(M)$ still further in
that he considers Fenchel-Orlicz spaces \cite{Turett}. The definition of
these spaces is similar to that of Orlicz sequence spaces, but they
are built on a Young function on $\R^n$ rather than $\R$ and consist of
vector-valued sequences. Now Nigel and his coauthors proved that Fenchel-Orlicz
spaces can be renormed to have propert~$(M)$ and that many interesting
Banach spaces have a represention as a Fenchel-Orlicz space. This is
in particular so for the ``twisted sums'' $Z_p$, $1<p<\infty$, from
\kcite{KaltonPeck}. These are ``extreme'' counterexamples to the
three-space problem for $\ell_p$: $Z_p$ is not isomorphic to $\ell_p$,
yet contains a subspace $Y_p$ isomorphic to $\ell_p$ such that
$Z_p/Y_p$ is isomorphic to $\ell_p$ as well. In the language of
homological algebra, $Z_p$ is a nontrivial twisted sum of $\ell_p$
with itself, i.e., there is a short exact sequence $0\to \ell_p \to
Z_p\to \ell_p \to 0$ that does not split. Nigel has contributed a lot
to twisted sums, but this is another story.
In a subsequent publication \kcite{Kal-Wer} Nigel pursued
an idea mentioned at the end of \kcite{Kal-M}, namely to decide whether the
unconditionality assumption, i.e., $\limsup \|\Id-2K_i\|\le1$, in
Theorem~\ref{theo2.1} is actually
needed. It turns out that this is not so.
\begin{theo}\label{theo2.2}
For a separable Banach space $X$, $K(X)$ is an $M$-ideal in $L(X)$
if and only if $X$ has
property~$(M)$, does not contain a copy of $\ell_1$ and has the
metric compact approximation property.
\end{theo}
The proof of the if-part consists of two steps: first to show that $X$
has property~$(M^*)$ as well, which is much more difficult than the
implication $(M^*)$ $\Rightarrow$ $(M)$, and then to construct, using
property~$(M^*)$, from a compact approximation of the identity
satisfying $\limsup\|K_n\|\le1$ another compact approximation of the identity
satisfying $\limsup\|\Id-2L_n\|\le1$. This is done by a skipped
blocking decomposition argument. Meanwhile other and simpler arguments
for the second step have been given by \AA.~Lima \cite{Lima1995},
E.~Oja \cite{Oja2000} and
O.~Nygaard and M.~P\~oldvere \cite{NygPold}.
This theorem was proved while I was a visitor at the University of
Missouri in 1993. Let me commit myself to some personal recollections
at this stage. The day I arrived, Nigel asked me what I was working
on. One of the questions had to do with Banach spaces $X$ for which
$K(X\oplus_p X)$ is an $M$-ideal in $L(X\oplus_p X)$, like
$X=\ell_p$. The conjecture was that such a space should be similar to
$\ell_p$, more precisely such an $X$ should embed almost isometrically
into an $\ell_p$-sum of finite-dimesional spaces. In fact, in the
paper \cite{Dirk4} I had previously formulated the bold conjecture that all
spaces for which $K(X)$ is an $M$-ideal in $L(X)$ are stable in the
sense of Krivine and Maurey \cite{KriMau} and that
one should be able to deduce from that that $K(X\oplus_p X)$ is an
$M$-ideal in $L(X\oplus_p X)$ for some~$p$. The first half was
disproved by Nigel in \kcite{Kal-M} whereas he did prove the second part to be
correct. Almost on the spot he suggested an idea how to tackle the
problem. Of course, it took me some time to digest it, and after a
week or so I understood what he had in mind. I then suggested to use
an ultrapower argument at some stage of the proof, upon which Nigel
said, ``Oh, I missed that point!'' -- only to come up with a much
better idea that eventually solved the problem. I also pointed out a
relation to work by Bill Johnson and Morry Zippin, and Nigel asked, ``Can
they do it without the approximation property?'', which was the case,
and he added, ``Then we can do without the approximation property
too!'' I'll describe the outcome in the next section.
\section{Almost isometric embeddings}\label{sec3}
Let us start with some vocabulary. We say that a (separable) Banach
space $X$ has property~$(m_p)$ if, whenever $x_n\to 0$ weakly,
$$
\limsup \|x+x_n\|^p = \|x\|^p + \limsup\|x_n\|^p \qquad\forall x\in X
$$
if $p<\infty$, resp.
$$
\limsup \|x+x_n\| = \max\{ \|x\|, \limsup\|x_n\| \} \qquad\forall x\in X
$$
for $p=\infty$. It is clear that $\ell_p$ has $(m_p)$ for $p<\infty$
(cf.\ (3) above) and that $c_0$ has $(m_\infty)$. If $K(X\oplus_p X)$
is an $M$-ideal, then $X$ has $(m_p)$, as proved by Nigel \kcite{Kal-M}.
The Johnson-Zippin space $C_p$ is an $\ell_p$-sum of a sequence of
finite-dimen\-sional spaces $E_1,E_2,\dots$ that are dense in all
finite-dimensional spaces with respect to the Banach-Mazur distance. A
Banach space $X$ embeds almost isometrically into $Y$ if for each
$\eps>0$ there is a subspace $X_\eps\subset Y$ such that
$d(X,X_\eps)\le 1+\eps$, $d(X,X_\eps)$ denoting the Banach-Mazur
distance. We use a similar definition for $X$ to be almost isometric
to a quotient of $Y$. Note that any two versions of $C_p$ (built on
different $E_k$) embed almost isometrically into each other.
In \kcite{Kal-Wer}, the following result is proved.
\begin{theo}\label{theo3.1}
Suppose $X$ is a separable Banach space not containing $\ell_1$. Let
$1<p<\infty$. Then $X$ has $(m_p)$ if and only if $X$ embeds almost
isometrically into $C_p$. Likewise, $X$ has $(m_\infty)$ if and only
if $X$ embeds almost isometrically into $c_0$.
\end{theo}
The proof uses again a skipped blocking decomposition
technique.
Because of the duality of the property $(m_p)$, one obtains the
following corollary.
\begin{cor}\label{cor3.2}
Let $1<p<\infty$. If $X\subset C_p$, then $X$ is almost isometric to a
quotient of $C_p$, and if $X=C_p/Z$, then $X$ is almost isometric to a
subspace of $C_p$.
\end{cor}
This is an almost isometric refinement of a result due to Johnson and
Zippin who proved the corresponding isomorphic result \cite{JohZip2}. On the
other hand, isomorphic versions of Theorem~\ref{theo3.1} using tree
conditions were later obtained by Nigel for $p=\infty$ \kcite{Kal-QJM01}
\label{page5}and E.~Odell and
Th.~Schlumprecht for $p<\infty $ \cite{OdeSch}.
In the context of $L_p$-spaces more can be proved. First of all, if
$X\subset L_p=L_p[0,1]$ has property~$(M)$, then it has $(m_r)$ for
some $r$; if $1<p\le2$, then $p\le r\le 2$, and if $2<p<\infty$, then
$r=2$ or $r=p$. We now have \kcite{Kal-Wer}:
\begin{theo}\label{theo3.3}
Suppose $1<p<\infty$, $p\neq2$, and let $X\subset L_p$ be
infinite-dimensional. Then the following are equivalent:
\begaeq
\item
$B_X$, the unit ball of $X$, is compact in $L_1$ (i.e., with respect
to the topology inherited from $L_1$).
\item
$X$ has property $(m_p)$.
\item
$X$ embeds almost isometrically into $\ell_p$.
\endaeq
If $p>2$, then \mbox{\rm{(i)--(iii)}} are equivalent to each of the
following:
\begaeq
\item[\rm(iv)]
$X$ is isomorphic to a subspace of $\ell_p$.
\item[\rm(v)]
$X$ does not contain a copy of $\ell_2$.
\endaeq
\end{theo}
Our proof was (iii) $\Rightarrow$ (ii) $\Rightarrow$ (i) $\Rightarrow$
(iii) and (iii) $\Rightarrow$ (iv) $\Rightarrow$ (v)
$\Rightarrow$ (i) for $p>2$, where (iii) $\Rightarrow$ (iv)
$\Rightarrow$ (v) is trivial, as is (iii) $\Rightarrow$ (ii).
Originally the equivalence of (iv) and
(v) for $p>2$ is due to Bill Johnson and Ted Odell \cite{JohOde}, see
also \cite{John3}.
Concrete examples of subspaces of $L_p$ with $(m_p)$ are the Bergman
spaces consisting of all analytic functions on the unit disc
$\D=\{z\in\C\dopu |z|<1\}$ for
which $\int_\D |f(x+iy)|^p\,dx\,dy<\infty$. Likewise, the ``little''
Bloch space has $(m_\infty)$. In \kcite{Kal-Coll}, Nigel proved that
also the ``little''
Lipschitz space has $(m_\infty)$, thus showing that it is an $M$-ideal
in its bidual, a problem left open in \cite{BerWer}.
Results concerning isometric embeddings of subspaces of $L_p$ into
$\ell_p$ were proved by F.~Delbaen, H.~Jarchow and A.~\pel\
\cite{DelJarPel}; as is
often the case in isometric considerations in $L_p$, one has to
distinguish whether or not $p$ is an even integer.
In a conversation in 1998, Nigel once suggested to prove a result
similar to Theorem~\ref{theo3.1} for property~$(M)$. His idea was to
show that such spaces embed into Fenchel-Orlicz spaces. I am sure that
he knew an outline of the argument, but as far as I know the proof has
never been written down.
There is also an isomorphic version of Theorem~\ref{theo3.1} for
$p=\infty$, devised by Nigel and his coauthors G.~Godefroy and
G.~Lancien \kcite{GodKalLan}.
The main result of their paper is that the Banach space
$c_0$ is determined by its metric, that is:
\begin{theo}\label{theo3.4}
If a Banach space $X$ is Lipschitz isomorphic to $c_0$, that is, if
there is a bijective map $T\dopu X\to c_0$ with $T$ and $T^{-1}$
Lipschitz, then $X$ is linearly isomorphic to $c_0$, that is, $T$ can
be chosen linear.
\end{theo}
To prove this, they first show that $X$ embeds isomorphically into
$c_0$, for which an isomorphic version of property~$(m_\infty)$ and of
Theorem~\ref{theo3.1} is needed. To conclude the proof of
Theorem~\ref{theo3.4} one has to appeal to known properties of
subspaces of $c_0$. Theorem~\ref{theo3.4} is one of the most
remarkable achievements in the nonlinear theory of Banach spaces.
In their paper \kcite{GodKalLi1}, Nigel together with G.~Godefroy and D.~Li
addressed the problem of extending results like Theorem~\ref{theo3.3}
to the case of subspaces of~$L_1$.
There are two intrinsic difficulties that do not occur in the case
$p>1$. For one thing, the Haar basis is unconditional in $L_p$ for
$p>1$, but not in $L_1$, and this was an essential ingredient in the
proof of Theorem~\ref{theo3.3}. Also, the $L_1$-topology on the unit ball of
a subspace $X\subset L_p$ is certainly locally convex, whereas the
$L_r$-spaces for $r<1$ are not locally convex and hence the
$L_r$-topology on the unit ball of
a subspace $X\subset L_1$ need not be locally convex. (Here we enter
the world of quasi-Banach spaces, one of Nigel's favourite
areas.)
So in order to be able to study when subspaces of $L_1$ embed into
$\ell_1$, one has to assume some unconditionality.
A separable Banach space has the \textit{unconditional metric
approximation property} (UMAP for short)
\label{page6}if there is a sequence of
finite rank operators such that $F_nx\to x$ for all $x$ and
$\|\Id-2F_n\|\to1$; the latter obviously implies that $\|F_n\|\to 1$
as well. We have already encountered a variant in
Theorem~\ref{theo2.1}; the definition of UMAP is due to Nigel and Pete
Casazza \kcite{CasKal}. We shall have more to say on this in
Section~\ref{sec4}.
In the next theorem, one of the main results from \kcite{GodKalLi1}, $\tau_m$
denotes the topology of convergence in measure, i.e., the topology of
the $F$-space~$L_0$.
\begin{theo}\label{theo3.5}
Let $X$ be a subspace of $L_1$ with the approximation property. The
following statements are equivalent.
\begaeq
\item
$X$ has the UMAP, and $B_X$ is relatively compact for the
topology~$\tau_m$.
\item
$B_X$ is $\tau_m$-compact and $\tau_m$-locally convex.
\item
For any $\eps>0$, there exists a weak\/$^*$ closed subspace $X_\eps$
of $\ell_1$ with Banach-Mazur distance $d(X,X_\eps)\le1+\eps$.
\endaeq
\end{theo}
By a result of Rosenthal, $B_X$ is $\tau_m$-relatively compact if and
only if $X$ fulfills a strong quantitative version of the Schur
property, called the $1$-strong Schur property.
The paper \kcite{GodKalLi1} also contains a very interesting
counterexample as to
possible generalisations of the previous theorem.
\begin{theo}\label{theo3.6}
There exists a subspace $X$ of $L_1$ with the approximation property,
whose unit ball is $\tau_m$-compact but not $\tau_m$-locally
convex. In particular, $X$ fails the UMAP.
\end{theo}
The construction of this space would not have been possible without
Nigel's insight into the nature of $L_p$-spaces for $0\le p<1$---for
most of us a no-go area---in particular the strange world of needle
points and the failure of the Krein-Milman theorem there, cf.\ \cite{Rob76}.
Another type of embedding result is contained in Nigel's work with
Alex Koldobsky \kcite{KalKol}; it concerns subspaces of the quasi-Banach space
$L_p$ for $p<1$. It is known that a Banach space $X$ embeds into some
$L_p$, $p<1$, isomorphically if and only if $X$ embeds into all
$L_r$, $0<r<1$, isomorphically. (The embedding into $L_r$ for $r\le p$
is clear since $L_p$ embeds into those $L_r$ isometrically; the issue
is the range of $r$ between $p$ and~$1$.) Nigel proved in \kcite{Kal85}
that embedding into $L_1$ is equivalent to the
embeddability of $\ell_1(X)$ into $L_p$.
As for the corresponding isometric question, A.~Koldobsky \cite{Kol96}
showed that there is a finite-dimensional Banach space that embeds
isometrically into $L_{1/2}$, but not into $L_1$. Using stable random
variables, Nigel and Alex Koldobsky obtain a vast generalisation.
\begin{theo}\label{theo3.7}
Let $0<p<1$; then there is an infinite-dimensional Banach space $E_p$
that embeds isometrically into $L_p$, but not into any $L_r$ for
${p<r\le 1}$.
\end{theo}
The space $E_p$ has a representation as $\ell_1 \oplus_{N_p} \R$ by
means of some very cleverly chosen absolute norm $N_p$ on $\R^2$, and
the embedding of $E_p$ into $L_p$ is realised by means of a sequence
of independent random variables having a $1$-stable (i.e., Cauchy)
distribution. In addition to this example $E_p$, a second example is
constructed that is isomorphic to a Hilbert space, viz.\ $F_p= \ell_2
\oplus_{\tilde N_p} \ell_2$.
\section{Unconditionality}\label{sec4}
We have already mentioned the notion of unconditional metric
approximation property (UMAP) on page~\pageref{page6}.
The UMAP was introduced in the paper \kcite{CasKal} by Nigel and Pete Casazza.
Let us explain what ``unconditional'' refers to here. Following Nigel
and Pete one can obtain from an approximating sequence $(F_n)$ with
$\|\Id-2F_n\|\to1$, for a given $\eps>0$, another approximating sequence
$(F'_n)$ such that for $A_n=F_n'-F_{n-1}'$ and all~$N$
$$
\biggl\|\sum_{n=1}^N \eps_n A_n \biggr\| \le 1+\eps
$$
whenever $\eps_n=\pm1$; this should be compared to the estimate
$$
\biggl\|\sum_{n=1}^N \eps_n e_n^*(x)e_n \biggr\| \le (1+\eps)\|x\|
$$
for $(1+\eps)$-unconditional bases. Hence the epithet ``unconditional.''
Replacing finite rank operator by compact operators in the
approximating sequence, one arrives at the notion of unconditional metric
compact approximation property (UMCAP). Apart form studying these
properties, \kcite{CasKal} also contains the proof of the following stunning
result concerning the ordinary metric approximation property (MAP).
\begin{theo}\label{theo4.1}
If a separable Banach space has the MAP, then it even has the
commuting MAP, meaning that there are commuting finite rank operators
with $\|T_n\|\le1$ and $T_nx\to x$ for all~$x$.
\end{theo}
For reflexive spaces, the same type of conclusion is proved for the
UMAP in \kcite{CasKal} as well, but in a paper by Nigel and
G.~Godefroy \kcite{GodKal97},
the result was proved in full generality.
\begin{theo}\label{theo4.2}
If a separable Banach space has the UMAP, then it even has the
commuting UMAP.
\end{theo}
The key to the proof is to look at some norm-$1$
approximating sequence $(F_n)$
and to study the limiting projection
$$
Px^{**} = w^*\mbox{-}\lim_{n\to\infty} F_n^{**}x^{**}
$$
in the bidual and to prove that its range is weak$^*$ closed. The
paper \kcite{GodKal97} also contains
an embedding result that is similar in spirit to Theorem~\ref{theo3.1}
and Theorem~\ref{theo3.3}:
\begin{theo}\label{theo4.3}
A Banach space $X$ has the UMAP if and only if, for each $\eps>0$, $X$
embeds isometrically as a $(1+\eps)$-complemented subspace into a space with
a $(1+\eps)$-unconditional Schauder basis.
\end{theo}
Another outgrowth of \kcite{CasKal} is the concept of a $u$-ideal that is
studied in detail in the influential paper \kcite{GKS} by Nigel, G.~Godefroy
and P.~Saphar.
Let $X\subset Y$ be Banach spaces. $X$ is said to be a
\textit{$u$-summand} in $Y$ if there is a projection $P$ from $Y$ onto
$X$ with $\|\Id-2P\|=1$; equivalently, one may decompose $Y=X\oplus
X_s$ in such a way that $\|x+x_s\|=\|x-x_s\|$ whenever $x\in X$,
$x_s\in X_s$.
An easy example of a $u$-decomposition is the decomposition of $f\in
C[-1,1]$ into its even and odd part.
An important special case is when $Y=X^{**}$; for
example, the bidual of $L_1$ admits such a decomposition, which is
even $\ell_1$-direct: $(L_1)^{**}=L_1 \oplus_1 (L_1)_s$, the so-called
Yosida-Hewitt decomposition. (In technical terms, $L_1$ is an example
of an $L$-embedded space; see Chapter~IV in \cite{HWW}.)
Likewise, $X$ is a \textit{$u$-ideal} in $Y$ if $X^\bot$, its
annihilator, is a $u$-summand in~$Y^*$. By definition, every $M$-ideal is a
$u$-ideal, but also every order ideal in a Banach lattice is a
$u$-ideal. When working with complex scalars, it is more appropriate
to replace the condition $\|\Id-2P\|=1$ by $\|\Id-(1+\lambda)P\|=1$
for all scalars of modulus~$1$; correspondingly, one then speaks of
\textit{$h$-summands} and \textit{$h$-ideals}.
In these definitions, $u$ stands for
``unconditional'' and $h$ for ``hermitian.''
The extensive paper \kcite{GKS} contains a wealth of information concerning
$u$-ideals and $h$-ideals. I will mention only a few aspects. First
of all, there are certain similarities to $M$-ideals, for example, the
$u$-projection is uniquely determined, and a $u$-ideal that does not
contain a copy of $c_0$ is a $u$-summand. On the other hand, if $X$ is
a $u$-ideal in its bidual, then the $u$-projection need not be the
canonical projection from the decomposition $X^{***}=X^\bot \oplus
X^*$,
as for $M$-embedded spaces. (An example is $X=L^1$.) Let us say that
$X$ is a strict $u$-ideal (in its bidual) in this case.
One of the
main results in \kcite{GKS} characterises strict $u$-ideals by means of a
quantitative version of Pe{\l}czy\'nski's property~$(u)$. To explain
this, some notation is needed. Let $x^{**}\in X^{**}$ be such that
there is a sequence in $X$ converging to $x^{**}$ in the weak$^*$
topology of $X^{**}$. Define
$$
\kappa_u(x^{**}) =
\inf \biggl\{ \sup_n \Bigl\| \sum_{k=1}^n \eps_k x_k
\Bigr\| \dopu \eps_k=\pm1,\ x_k\in X,\
x^{**}=w^*\mbox{-}\sum_{k=1}^\infty x_k \biggr\}
$$
and let $\kappa_u(X)$ be the smallest constant such that
$\kappa_u(x^{**})\le C \|x^{**}\|$ for all $x^{**}$ in the sequential
closure of $X$ in $(X^{**},w^*)$. $\kappa_h$ is defined in the same
way, replacing $\eps_k=\pm1$ with $|\eps_k|=1$, $\eps_k\in\C$. The
Banach space $X$ has property $(u)$ if $\kappa_u(X)<\infty$.
\begin{theo}\label{theo4.4}
Suppose $X$ does not contain a copy of $\ell_1$. Then $X$ is a strict
$u$-ideal in $X^{**}$ if and only if $\kappa_u(X)=1$.
\end{theo}
As for spaces of operators, Nigel and his collaborators obtain the
following results.
\begin{theo}\label{theo4.5}
Let $X$ be a separable reflexive Banach space. Then $X$ has the UMCAP if and
only if $K(X)$ is a $u$-ideal in $L(X)$.
\end{theo}
Generally, the results for $h$-ideals are more satisfactory and
precise, since one can apply the powerful machinery of hermitian
operators; for example (the notion of complex UMCAP should be
self-explanatory):
\begin{theo}\label{theo4.6}
Let $X$ be a complex Banach space with a separable dual. Then $X$ has
the complex UMCAP if and
only if $K(X)$ is an $h$-ideal in $L(X)$ and $X$ is an $h$-ideal in $X^{**}$.
\end{theo}
\begin{theo}\label{theo4.7}
Let $X$ be a separable complex Banach space with the MCAP.
Then $K(X)$ is an $h$-ideal in its bidual $K(X)^{**}$ if and only if
$X$ is an $M$-ideal in $X^{**}$ and has the complex UMCAP.
\end{theo}
In their paper \kcite{CowKal} (with S.R.~Cowell) Nigel takes up the question of
embedding into a space with an unconditional basis, as in
Theorem~\ref{theo4.3}, but without the assumption of the approximation
property. Such a result was given by W.B.~Johnson and B.~Zheng
\cite{JohZhe}, but now the aim is to find an (almost) isometric
version. The key to this are the asymptotic unconditionality properties $(au)$
and $(au^*)$.
A separable Banach space $X$ has \textit{property $(au)$} if
$$
\lim (\|x+x_d\|-\|x-x_d\|) =0
$$
whenever $(x_d)$ is a bounded weakly null
net; in \kcite{Kal-QJM01} Nigel referred to this property as ``$X$ is
of unconditional type.''
If $X^*$ is separable, it is equivalent to use weakly null
sequences instead, and one obtains a notion that has been known by the
acronym WORTH in the literature. Likewise, $X$ has \textit{property
$(au^*)$} (previously ``$X$ is of shrinking unconditional type'') if
$$
\lim (\|x^*+x_n^*\|-\|x^*-x_n^*\|) =0
$$
whenever $(x_n^*)$ is a
(necessarily bounded) weak$^*$ null sequence; due to the weak$^*$
metrisability of the unit ball there is no need to consider nets
here.
These notions are reminiscent of the properties $(M)$,
$(M^*)$ and $(m_p)$, and
$(au^*)$ has also been considered by \AA.~Lima \cite{Lima1995}
under the name
$(wM^*)$.
In general $(au^*)$ implies $(au)$, and
the converse is true under additional hypotheses.
The main result of \kcite{CowKal} says the following.
\begin{theo}\label{theo4.8}
A separable Banach space $X$ has
property $(au^*)$ if and only if $X$ embeds almost isometrically into
a space $Y$ with a shrinking $1$-uncondi\-tional basis; if $X$ is
reflexive, $Y$ can be taken to be reflexive as well.
\end{theo}
I will briefly mention other contributions by Nigel on the topic of
unconditional bases, Schauder decompositions and expansions. In a
series of papers with P.~Casazza or F.~Albiac and C.~Ler\'anoz,
Nigel investigated
uniqueness of unconditional bases in Banach or quasi-Banach spaces
(\kcite{CasKal96}, \kcite{CasKal98}, \kcite{CasKal99},
\kcite{AlbKalLer03}, \kcite{AlbKalLer04}).
For example, in \kcite{CasKal99} it is proved that although the
Tsirelson space $T$ admits a unique unconditional basis, this is not
so for $c_0(T)$. In the paper \kcite{DefKal} with A.~Defant the question of
existence of an unconditional basis in the space $P(^m E)$ of
$m$-homogeneous on a Banach space $E$ is considered. Sean Dineen had
conjectured that $P(^m E)$ has an unconditional basis if and only if
$E$ is finite-dimensional. This conjecture is vindicated in
\kcite{DefKal};
the proof uses local Banach space theory and some greedy basis theory.
See also Theorem~\ref{theo5.8} below for unconditional expansions in $L_1$.
\section{Operators on $L_1$}\label{sec5}
Nigel looked at operators on function spaces like $L_1$ or indeed
$L_p$ for $p\le1$ in a vast number of papers; I will report on a small
sample of his results.
In \kcite{Kal-Ind78} he devised a very useful representation
theorem for operators on $L_p$, $0\le p\le1$, by means of random
measures. For $p=1$ the result is as follows.
\begin{theo}\label{theo5.1}
If $T\dopu L_1[0,1]\to L_1[0,1]$ is a bounded linear operator, then
there are measures $\mu_x$ on $[0,1]$, with $x\mapsto \mu_x\in M[0,1]$
weak\/$^*$ measurable, such that
$$
(Tf)(x)= \int_0^1 f(s)\,d\mu_x(s) \quad a.e.
$$
Moreover,
$$
\|T\|= \sup_{\lambda(B)>0} \frac1{\lambda(B)} \int_0^1 |\mu_x(B)|\,dx.
$$
\end{theo}
Decomposing the measures $\mu_x$ into their atomic and continuous
parts $\mu_x^a$ and $\mu_x^c$ allows to define the corresponding
operators
$$
(T^af)(x)= \int_0^1 f(s)\,d\mu^a_x(s), \quad
(T^cf)(x)= \int_0^1 f(s)\,d\mu^c_x(s) .
$$
This permits Nigel to derive the following variant of a result due to
P.~Enflo and T.W.~Starbird \cite{EnfStar}.
\begin{theo}\label{theo5.2}
If $T^a\neq0$, there is a Borel set $B$ of positive measure such that
$\rest{T}{L_1(B)}$ is an into-isomorphism whose range is
complemented.
\end{theo}
Let me remark that it is still an open problem whether a complemented
infinite-dimensional
subspace of $L_1$ is isomorphic to $L_1$ or $\ell_1$. However, Enflo
and Starbird have proved that $L_p$ is primary for
$p\ge1$. This means that whenever $L_p$ is isomorphic to a direct sum
$X\oplus Y$, then $X$ or $Y$ is isomorphic to~$L_p$. Using his
representation theorem, Nigel extends this result to $p<1$ and obtains
a new proof for $p=1$. Thus:
\begin{theo}\label{theo5.3}
$L_p$ is primary for $0<p<\infty$.
\end{theo}
The representation theorem~\ref{theo5.1} is also used in the
paper \kcite{KalBeata} of Nigel and Beata Randrianantoanina, where
they show that a
surjective isometry on a real rearrangement invariant space $X$ on
$[0,1]$ different from $L_2$ has the form $(Tf)(s)= a(s)f(\sigma(s))$;
if $X$ is not isomorphic to $L_p$ for any $1\le p\le \infty$, then in
addition $|a|=1$ a.e.\ and $\sigma$ is measure preserving.
In \kcite{GodKalLi2}, Nigel and his coauthors G.~Godefroy and D.~Li take up
the line of reasoning based on Theorem~\ref{theo5.1} to obtain results
of a more isometric flavour, for example a quantitative version of a result
due to D.~Alspach \cite{Als2}.
\begin{theo}\label{theo5.4}
There is a function $\varphi$ with $\lim_{\alpha\to 0^+}
\varphi(\alpha)=1$ such that if $T\dopu L_1\to L_1$ satisfies
$$
\|f\|\le \|Tf\| \le \alpha \|f\|\qquad \forall f\in L_1,
$$
then there is an isometry $J\dopu L_1\to L_1$ such that $\|T-J\|\le
\varphi(\alpha)$.
\end{theo}
In other words, there is always an into isometry close to a given into
near-isometry.
The paper also contains an example, of a similar nature as that of
Theorem~\ref{theo3.6}, to show that this result does not extend to
(near-) isometries from subspaces of $L_1$ to~$L_1$.
Specifically, Nigel and his coauthors investigate certain subspaces of
$L_1$ that they call \textit{small subspaces} and operators that they call
\textit{strong Enflo operators}. The latter means that in the
representation of $T$ by the measures $\mu_x$ one has
$\mu_x(\{x\})\neq0$ on a set of positive measure. A subspace $X$ of
$L_1$ is called small if the mapping $f\mapsto \rest fA$ from $X$ to
$L_1(A)$ is not surjective for any $A\subset [0,1]$ of positive
measure. In other words, the inclusion $B_{L_1(A)} \subset kB_X$ is
false whenever $\lambda(A)>0$ and $k>0$, where we consider
$L_1(A)\subset L_1[0,1]$ in a natural way. The following short-hand
notation is now handy. Say
$$
M \simsubset N
$$
for two subset of $L_1$ if there is a positive constant $k$ such that
$M\subset kN$, i.e., $M$ is absorbed by~$N$.
Hence $X$ is not small if $B_{L_1(A)} \simsubset B_X$.
It is worthwhile to equip $L_1$ with the topology $\tau_m$ of
convergence in measure, defined above Theorem~\ref{theo3.5}.
For a subspace $X\subset
L_1$, let $C_X$ be the closure of $B_X$ in $L_1$ with respect to
$\tau_m$. The subspace $X$ is called \textit{nicely placed} if
$B_X=C_X$, that is, if its unit ball is closed for the topology
$\tau_m$. By a theorem due to A.V.~Buhvalov and G.J.~Lozanovski\u\i\
\cite[page~183]{HWW}, $X$ is nicely placed if and only if the Yosida-Hewitt
projection associated to the decomposition
$$
(L_1)^{**}=L_1 \oplus (L_1)_s
$$
maps $X^{\bot\bot}$ onto $X$. Hence, $X\subset L_1$ is nicely placed
if and only if it is $L$-embedded, i.e., $X^{**}=X\oplus_1
X_s$. Nicely placed subspaces were introduced in \cite{G-bien} and studied
intensively in many papers, see Chapter~IV in \cite{HWW} for a survey.
Returning to the paper \kcite{GodKalLi2}, let us note the following result.
\begin{theo}\label{theo5.5}
A subspace $X$ of $L_1$ is small if and only if no strong Enflo
operator satisfies $T(B_{L_1})\simsubset C_X$. Consequently, a nicely
placed subspace is small if and only if no operator from $L_1$ to $X$
is a strong Enflo operator.
\end{theo}
Since strong Enflo operators have a nonzero atomic part, they fix a
copy of $L_1$ by Theorem~\ref{theo5.2}; therefore a nicely placed
subspace of $L_1$ not containing $L_1$ is small.
Another feature of operators that are not strong Enflo operators is
the equation
$$
\|\Id+T\|=1+\|T\|,
$$
known as the \textit{Daugavet equation}. By the above, this is
fulfilled for all operators valued in a small nicely placed
subspace. The Daugavet equation is one of the technical ingredients in
the proof of the main result of \kcite{GodKalLi2}:
\begin{theo}\label{theo5.6}
Let $X$ and $Y$ be small subspaces of $L_1$ and suppose that there is
an isomorphism $S\dopu L_1/X \to L_1/Y$ with
$\max\{\|S\|,\|S^{-1}\|\}<1+\delta< 1.25$. Then there is an
invertible operator $U\dopu L_1\to L_1$ such that
$\|U\|\,\|U^{-1}\|\le (1+\delta)/(1-25\delta)$ and
$d_{\mathcal{H}}(U(B_X),B_Y)\le 71\delta/(1-25\delta)$, where
$d_{\mathcal{H}}$ denotes the Hausdorff distance.
\end{theo}
If $X$ and $Y$ are additionally assumed to be nicely placed, the
conclusion can be strengthened: If $\|S\|\,\|S^{-1}\|=1+\alpha<2$,
then $U$ above can be chosen to map $X$ onto $Y$ and
$\|U\|\,\|U^{-1}\|\le (1+\alpha)/(1-\alpha )$. A particular
consequence in this setting is: If $L_1/X$ and $L_1/Y$ are isometric,
then so are $X$ and $Y$.
Leaving the field of small subspaces let us turn to \textit{rich
subspaces}. These were introduced by A.~Plichko and M.~Popov
\cite{PliPop}; later
the definition was slightly modified \cite{KadSW2} to accommodate the general
setting of Banach spaces with the Daugavet property introduced in
\cite{KadSSW}.
A Banach space $X$ has the \textit{Daugavet property} if the Daugavet
equation
$$
\|\Id+T\|=1+\|T\|
$$
holds for all compact operators $T\dopu X\to X$; examples include
$L_1[0,1]$, $C[0,1]$, $L_\infty[0,1]$, the disc algebra, $L_1/H^1$ and
many other function spaces, but also more pathological spaces in the
spirit of Theorem~\ref{theo3.6} \cite{KW}. A subspace $Y$ of $X$ is
called rich if
whenever $Y\subset Z\subset X$, then $Z$ has the Daugavet
property. (There are other equivalent reformulations of this, see e.g.\
the survey \cite{Wer-IMB}.)
I did some work on these notions with Vova Kadets and
other coauthors. Upon hearing about some of our results, Nigel
immediately contributed the following theorem, published in our joint
paper \kcite{KKW1}, showing that rich subspaces are indeed the largest possible
proper subspaces of~$L_1$. Recall that $C_X$ is the closure of $B_X$
for the topology of convergence in measure.
\begin{theo}\label{theo5.7}
A subspace $X\subset L_1$ is rich if and only if, for each
$1$-codimensional subspace $H\subset X$, $\frac12 B_{L_1}\subset
C_H$. On the other hand, if $rB_{L_1}\subset C_X$ for some
$r>\frac12$, then $X=L_1$.
\end{theo}
One of the consequences of Nigel's representation theorem
(Theorem~\ref{theo5.1}) is not only the primariness of $L_1$, but more
generally that whenever $L_1$ is isomorpic to an unconditional
Schauder decomposition $X_1 \oplus X_2 \oplus \cdots$, then one of the
$X_k$ is isomorphic to $L_1$. In other words, this result ponders on
possible or impossible representations of $\Id=\sum T_n$ as a
pointwise unconditionally convergent series. In our joint paper \kcite{KKW2}
(with V.~Kadets) we investigate this question further. This paper
introduces a class $\mathcal C$ of operators related to the narrow
operators of Plichko and Popov \cite{PliPop} and to the not sign preserving
operators of H.P.~Rosenthal \cite{Ros-emb}, but I will skip the exact
definition and will only mention that compact operators belong to this
class. Anyway, here is the result.
\begin{theo}\label{theo5.8}
Let $X$ be a Banach space and $T,T_n\dopu L_1\to X$ be bounded
operators such that $Tf=\sum_n T_nf$ unconditionally for each $f\in
L_1$. If each $T_n$ is in $\mathcal C$, then $T$ is in $\mathcal C$.
\end{theo}
Nigel's student R.~Shvydkoy had obtained a similar result for the
narrow operators in the case $X=L_1$ in his Ph.D. thesis.
Theorem~\ref{theo5.8} can be considered as a generalisation of
A.~Pe{\l}czy\'nski's classical theorem that $L_1$ does not embed into
a space with an unconditional basis. It also contains an unpublished
(in fact unwritten, but occasionally quoted) result due to
H.P.~Rosenthal \cite{Ros-stop} as a special case:
$L_1$ does not even sign-embed into
a space with an unconditional basis.
\section{Extensions of operators into $C(K)$-spaces}\label{sec6}
It is a classical fact that whenever $E\subset X$ are Banach spaces
and $T_0\dopu E\to L_\infty[0,1]$ is a bounded linear operator, then
there exists an extension $T\dopu X\to L_\infty[0,1]$ of the same
norm; however, there need not exist a bounded linear extension
whatsoever if $L_\infty[0,1]$ is replaced by $C[0,1]$. By constrast,
Joram Lindenstrauss's memoir \cite{Lin-Mem} presents a detailed study of those
range spaces for which compact operators are extendible; it turns out,
that, for all pairs $E\subset X$, every compact operator $T_0\dopu
E\to F$ admits, for every $\eps>0$, an extension $T\dopu X\to F$ of
norm $\|T\|\le (1+\eps)\|T_0\|$ if and only if $F^*$ is isometric to a
space $L_1(\Omega,\Sigma,\mu)$. In particular this is true for
$F=C[0,1]$.
More recently, the problem of extension of bounded operators into $C(K)$ was
reexamined by Bill~Johnson and Morry Zippin \cite{JohZip95} who obtained
\hbox{$(3+\eps)$}-extensions for the pair $(E,\ell_1)$ provided $E$ is
weak$^*$ (meaning $\sigma(\ell_1,c_0)$-) closed.
In a series of papers that are enormous in wealth of ideas, technical
mastery and also in size (\kcite{Kal-QJM01}, \kcite{Kal-NY07},
\kcite{Kal-Isr07}, \kcite{Kal-Ill08}) Nigel has contributed to this circle
of ideas. I will now describe just a few of his findings. Let us say,
following Nigel, that the pair $(E,X)$, with $E$ a subspace of the
separable space~$X$, has the
\textit{$\lambda$-$C$-extension property} if, given a bounded linear
operator $T_0\dopu
E\to C(K)$ into some separable
$C(K)$-space, there is an extension $T\dopu
X\to C(K)$ of norm $\|T\|\le \lambda\|T_0\|$. (The same problem can be
studied for Lipschitz maps where the Lipschitz constant plays the role
of the norm. This setting is investigated in detail in
\kcite{Kal-NY07}, \kcite{Kal-Isr07}.)
In this language the Johnson-Zippin result says that $(E,\ell_1)$ has
the $(3+\eps)$-$C$-extension property for every $\eps>0$ if $E\subset
\ell_1$ is weak$^*$ closed.
Now, Nigel improves the $(3+\eps)$-bound to $(1+\eps)$, and he also
obtains a converse: If $(E,\ell_1)$ has the $\lambda$-$C$-extension
property for some $\lambda>0$ and additionally
$\ell_1/E$ has an unconditional FDD,
then there is an automorphism $S\dopu \ell_1\to\ell_1$ such that
$S(E)$ is weak$^*$ closed. To prove this, he makes a detailed study of
spaces embedding into $c_0$ and presents the tree characterisation of
such spaces mentioned on page~\pageref{page5}. He also proves that if
$(E,\ell_1)$ has the $(1+\eps)$-$C$-extension
property for all $\eps>0$ (the ``almost isometric $C$-extension
property''), then $E$ is weak$^*$ closed.
In \kcite{Kal-Ill08} Nigel goes on to develop an approach to the $C$-extension
property that is rooted in the theory of types, introduced by
J.L.~Krivine and B.~Maurey \cite{KriMau}. Let $X$ be a separable Banach space. A
\textit{type} generated by a bounded sequence $(x_n)$ (or more
generally a bounded net) is a function of the form
$$
\sigma\dopu X\to \R,\quad \sigma(x)= \lim_{n\to\infty} \|x+x_n\|
$$
(provided all these limits exist); it is called a weakly null type if
$x_n\to0$ weakly. Property~$(M)$ can now be rephrased by saying that
for a weakly null type $\sigma$, $\sigma(x)$ depends only on $\|x\|$,
and $X$ has $(m_p)$ whenever each weakly null type has the form
$\sigma(x)= (\|x\|^p + \lim_n \|x_n\|^p)^{1/p}$. Property~$(au)$ is a
symmetry condition, viz.\ $\sigma(x)=\sigma(-x)$. Now $X$ is said to
have \textit{property~$(L)$} if two weakly null types coincide once they
coincide at~$0$:
$$
\sigma_1(0)= \sigma_2(0) \quad\Rightarrow\quad
\sigma_1(x)= \sigma_2(x) \mbox{ for all }x.
$$
Property $(L^*)$ is defined similarly using weak$^*$ null types in the
dual. Clearly, $\ell_p$ has properties $(L)$ and $(L^*)$ for
$1<p<\infty$, and so do certain renormings of Orlicz sequence spaces
or Fenchel-Orlicz spaces. These renormings are similar in spirit to
the renormings that yield property~$(M)$,
but not identical. Indeed, if $X$ has both properties $(M)$
and $(L)$ and fails to contain $\ell_1$, then it has $(m_p)$ for some
$1<p\le\infty$.
With the help of types, Nigel is able to characterise the almost
isometric $C$-extension property. I won't give the exact formulation
of this result, but will note only one special case.
\begin{theo}\label{theo6.1}
If $X$ is a separable Banach space with $(M^*)$ or $(L^*)$ and if
$E\subset X$, then $(E,X)$ has the almost
isometric $C$-extension property.
\end{theo}
In particular, this applies to (renormings of) the twisted sums
$Z_p$. Nigel also provides a particular renorming of $\ell_2$ to show
that the existence of almost isometric extensions need not guarantee
isometric extensions.
Moreover, Nigel gives the first examples of spaces with the following
universal property: Whenever $E$ embeds into a separable space $X$
isometrically, then the pair $(E,X)$ has the almost
isometric $C$-extension property.
Indeed, all weak$^*$ closed subspaces of $\ell_1$ have this property
as do the spaces of Theorem~\ref{theo3.6}. The proof again uses the
theory of types.
An intriguing study of the corresponding isomorphic property is
contained in \kcite{Kal-Ill08}. Let us say
that the pair $(E,X)$ of separable spaces has the
\textit{$C$-extension property} if, given a bounded linear
operator $T_0\dopu
E\to C(K)$ into some separable $C(K)$-space, there is a bounded linear
extension $T\dopu X\to C(K)$; here it is enough to
consider $K=[0,1]$ by Milutin's theorem. A Banach space $E$ has the
\textit{universal $C$-extension property} if $(E,X)$ has the
$C$-extension property whenever $E$ embeds into $X$
isometrically.
Nigel proves the following result:
\begin{theo}\label{theo6.2}
A separable Banach space $E$ has the universal $C$-extension property
if and only if $E$ is $C$-automorphic.
\end{theo}
The latter means that whenever $E_1\subset C[0,1]$ and $E_2\subset
C[0,1]$ are isomorphic to $E$, then there is an automorphism $S\dopu
C[0,1]\to C[0,1]$ mapping $E_1$ to $E_2$. One can paraphrase this by
saying that there is essentially only one way to embed $E$ into
$C[0,1]$.
It is a classical result due to J.~Lindenstrauss and A.~\pel\ \cite{LinPel2}
that $c_0$ is $C$-automorphic.
Theorem~\ref{theo6.2} enables Nigel to show that $c_0(X)$ is
$C$-automorphic as well if $X$ is (for example $X=\ell_1$),
but $\ell_p$ is not $C$-automorphic for $1<p<\infty$.
Indeed, for a certain superreflexive $Z\supset \ell_p$ with an
unconditional basis, the pair $(\ell_p,Z)$ fails the $C$-extension
property. If, however, $Z\supset \ell_p$ is a UMD-space with an
unconditional basis, then $(\ell_p,Z)$ satisfies the $C$-extension
property; in Nigel's own words, ``the appearance of the
UMD-condition is quite mysterious.''
A technical device to prove
these theorems are homogeneous mappings $\Phi\dopu X^* \to Z^*$ that are
weak$^*$ continuous on bounded sets such that $\Phi(x^*)$ extends
$x^*$ with a bound $\|\Phi(x^*)\|\le \lambda \|x^*\|$, for all
$x^*\in X^*$. Such mappings were introduced by M.~Zippin \cite{Zip03}
and are called Zippin selectors by Nigel.
Based on this notion, the following
technical key result on $X=\ell_p$ or, more generally, an
$\ell_p$-sum of finite-dimensional spaces, $1<p<\infty$, is proved,
where again types play an essential role:
\begin{theo}\label{theo6.3}
Let $X$ be as above.
For a separable superspace $Z\supset X$,
the pair $(X,Z)$ has the
$C$-extension property if and only if $Z$ can be renormed so as to
contain $X$ isometrically and such that
$$
\lim_{n\to\infty} \|z+x_n\| \ge
\lim_{n\to\infty} (\|z\|^p+ \|x_n\|^p)^{1/p}
$$
for all
$z\in Z$ and all weakly null sequences $(x_n)\subset X$, provided
both limits exist.
\end{theo}
Many of Nigel's extension theorems have counterparts for Lipschitz
functions; for a detailed study see \kcite{Kal-NY07} and \kcite{Kal-Isr07}.
|
\subsection{Vacuum structure}
Now, let us discuss the vacuum structure of this theory \cite{SYfstr}.
The vacua of the theory (\ref{model}) are determined by the zeros of
the potential (\ref{pot}). In general, the theory has a number of the so called $r$-vacua, in which
$r$ quarks condense. The range of variation of $r$ is $r=0,...,N$. Say,
$r=0$ vacua (there are $N$ such vacua) are always at strong
coupling. We have already explained that they are called the
monopole vacua \cite{SW1,SW2}. In this paper we will focus
on a particular set of vacua with the maximal number of condensed quarks, $r=N$.
The reason for this choice is that all U(1) factors of the gauge group are spontaneously
broken in these vacua, and, as a result,
they support non-Abelian strings \cite{HT1,ABEKY,SYmon,HT2}. The occurrence of strings ensures
the monopole confinement in these vacua.
Let us first assume that our theory is at weak coupling, so that we can
analyze it quasiclassically. Below we will explicitly formulate conditions on the quark mass terms
and $\mu$ which will enforce such a regime.
With generic values of the quark masses we have
\begin{equation}
C_{N_f}^{N}= \frac{N_f!}{N!(N_f-N)!}
\label{numva}
\end{equation}
isolated $r$-vacua in which $r=N$ quarks (out of $N_f$) develop
vacuum expectation values (VEVs).
Consider, say, the vacuum in which the first $N$ flavors develop VEVs, to be denoted as (1, 2 ..., $N$).
In this vacuum the
adjoint fields develop
VEVs too, namely,
\begin{equation}
\left\langle \left(\frac12\, a + T^a\, a^a\right)\right\rangle = - \frac1{\sqrt{2}}
\left(
\begin{array}{ccc}
m_1 & \ldots & 0 \\
\ldots & \ldots & \ldots\\
0 & \ldots & m_N\\
\end{array}
\right),
\label{avev}
\end{equation}
For generic values of the quark masses, the SU$(N)$ subgroup of the gauge
group is
broken down to U(1)$^{N-1}$. However, in the {\em special limit}
\begin{equation}
m_1=m_2=...=m_{N_f},
\label{equalmasses}
\end{equation}
the adjoint field VEVs do not break the SU$(N)\times$U(1) gauge group.
In this limit the theory acquires a global flavor SU$(N_f)$ symmetry.
With all quark masses equal and to the leading order in $\mu$,
the mass term for the adjoint matter (\ref{msuperpotbr})
reduces to the Fayet--Iliopoulos $F$-term of the U(1) factor of the SU$(N)\times$U(1) gauge group,
which does {\em not} break ${\mathcal N}=2\;$ supersymmetry \cite{HSZ,VY}. In this limit the Fayet--Iliopoulos $F$-term
can be transformed into the Fayet--Iliopoulos $D$-term by an SU$(2)_R$ rotation; the theory reduces to
${\mathcal N}=2\;$ SQCD described in detail, say, in \cite{SYrev}.
Higher orders in the parameter $\mu$
break ${\mathcal N}=2\;$ supersymmetry by splitting all ${\mathcal N}=2\;$ multiplets.
If the quark masses are unequal the U($N$) gauge group is broken down to U(1)$^{N}$
by the adjoint field VEV's (\ref{avev}). To the leading order in $\mu$, the superpotential
(\ref{msuperpotbr}) reduces to $N$ distinct FI terms, one in each U(1) gauge factor.
${\mathcal N}=2\;$ supersymmetry in each individual low-energy U(1) theory remains unbroken \cite{SYfstr}. It is
broken, however, being considered in the full U($N$) gauge theory.
Using (\ref{msuperpotbr}) and (\ref{avev}) it is not difficult to obtain the quark field VEVs
from Eq.~(\ref{pot}). By virtue of a gauge rotation they can be written as
\begin{eqnarray}
\langle q^{kA}\rangle &=& \langle\bar{\tilde{q}}^{kA}\rangle=\frac1{\sqrt{2}}\,
\left(
\begin{array}{cccccc}
\sqrt{\xi_1} & \ldots & 0 & 0 & \ldots & 0\\
\ldots & \ldots & \ldots & \ldots & \ldots & \ldots\\
0 & \ldots & \sqrt{\xi_N} & 0 & \ldots & 0\\
\end{array}
\right),
\nonumber\\[4mm]
k&=&1,..., N\,,\qquad A=1,...,N_f\, ,
\label{qvev}
\end{eqnarray}
where we present the quark fields as matrices in the color ($k$) and flavor ($A$) indices.
The Fayet--Iliopoulos $F$-term parameters for each U(1) gauge factor are given (in the quasiclassical
approximation) by the following expressions:
\begin{equation}
\xi_P = 2\left\{\sqrt{\frac{2}{N}}\,\,\,\mu_1\,\hat{m}+\mu_2(m_P-\hat{m})\right\},
\qquad P=1,...,N
\label{xis}
\end{equation}
and $\hat{m}$ is the average value of the first $N$ quark masses,
\begin{equation}
\hat{m}=\frac1{N}\sum_{P=1}^{N} m_P\, .
\label{avm}
\end{equation}
While the adjoint VEVs do not break the SU$(N)\times$U(1) gauge group in the limit
(\ref{equalmasses}), the quark condensate (\ref{qvev}) does result in the spontaneous
breaking of both gauge and flavor symmetries.
A diagonal global SU$(N)$ combining the gauge SU$(N)$ and an
SU$(N)$ subgroup of the flavor SU$(N_f)$
group survives, however. This is color-flavor locking. Below we will refer to this diagonal
global symmetry as to $ {\rm SU}(N)_{C+F}$.
Thus, the pattern of the
color and flavor symmetry breaking
is as follows:
\begin{equation}
{\rm U}(N)_{\rm gauge}\times {\rm SU}(N_f)_{\rm flavor}\to
{\rm SU}(N)_{C+F}\times {\rm SU}(\tilde{N})_F\times {\rm U}(1)\,,
\label{c+f}
\end{equation}
where $\tilde{N}=N_f-N$.
Here SU$(N)_{C+F}$ is a global unbroken color-flavor rotation, which involves the
first $N$ flavors, while the SU$(\tilde N )_F$ factor stands for the flavor rotation of the
$\tilde N$ quarks.
The presence of the global SU$(N)_{C+F}$ group is instrumental for
formation of the non-Abelian strings \cite{HT1,ABEKY,SYmon,HT2,SYfstr}.
As we will see shortly, the global symmetry of the dual theory is, of course,
the same, albeit the physical origin is different.
With unequal quark masses, the global symmetry (\ref{c+f}) is broken down to
U(1)$^{N_f-1}$ both by the adjoint and squark VEVs. This should be contrasted with the
theory with the Fayet--Iliopoulos term introduced through the $D$-term, in which the
quark VEVs are all equal and do not break the color-flavor symmetry.
Since the global (flavor) SU$(N_f)$ group is broken by the quark VEVs anyway, it will be helpful for
our purposes
to consider
the following mass splitting:
\begin{equation}
m_P=m_{P'}, \qquad m_K=m_{K'}, \qquad m_P-m_K=\Delta m
\label{masssplit}
\end{equation}
where
\begin{equation}
P, P'=1, ... , N\,\,\,\, {\rm and} \,\,\,\, K, K'=N+1, ... , N_f\,.
\label{pppp}
\end{equation}
This mass splitting respects the global
group (\ref{c+f}) in the $(1,2,...,N)$ vacuum. Moreover, this vacuum becomes isolated.
No Higgs branch develops. We will often use this limit below.
\subsection{Perturbative excitations}
Now let us discuss the perturbative excitation spectrum.
To the leading order in $\mu$, in the limit (\ref{masssplit}), the superpotential
(\ref{msuperpotbr}) reduces to the Fayet--Iliopoulos $F$-term of the U(1) factor of the gauge group. Since
both U(1) and SU($N$) gauge groups are broken by the squark condensation, all
gauge bosons become massive. In fact, with nonvanishing $\xi_P$'s (see Eq.~(\ref{xis})), both the quarks and adjoint scalars
combine with the gauge bosons to form long ${\mathcal N}=2\;$ supermultiplets \cite{VY}, for a review see \cite{SYrev}.
In the limit (\ref{masssplit}) $$\xi_P\equiv\xi\,,$$ and all states come in
representations of the unbroken global
group (\ref{c+f}), namely, in the singlet and adjoint representations
of SU$(N)_{C+F}$,
\begin{equation}
(1,\, 1), \quad (N^2-1,\, 1),
\label{onep}
\end{equation}
and in the bifundamental representations
\begin{equation}
\quad (\bar{N},\, \tilde N), \quad
(N,\, \bar{\tilde N})\,.
\label{twop}
\end{equation}
We mark representations in (\ref{onep}) and (\ref{twop}) with respect to two
non-Abelian factors in (\ref{c+f}). The singlet and adjoint fields are (i) the gauge bosons, and
(ii) the first $N$ flavors of the squarks $q^{kP}$ ($P=1,...,N$), together with their fermion superpartners.
The bifundamental fields are the quarks $q^{kK}$ with $K=N+1,...,N_f$.
These quarks transform in the two-index representations of the global
group (\ref{c+f}) due to the color-flavor locking.
In the limit (\ref{masssplit}) the mass of the $(N^2-1,\, 1)$ adjoint fields is
\begin{equation}
m_{(N^2-1,\,1)}=g_2\sqrt{\xi}\,,
\label{Wmass}
\end{equation}
while the singlet field mass is
\begin{equation}
m_{(1,\,1)}=g_1\, \sqrt{\frac{N}{2}}\,\sqrt{\xi}\,.
\label{phmass}
\end{equation}
The bifundamental field masses are given by
\begin{equation}
m_{(\bar{N},\, \tilde N)}= \Delta m\,.
\label{bifund}
\end{equation}
The above quasiclassical analysis is valid if the theory is at weak coupling. This is the case if
the quark VEVs are sufficiently large so that the gauge coupling constant is frozen at a large scale.
From (\ref{qvev}) we see that
the quark condensates are of the order of
$\sqrt{\mu m}$ (see also \cite{SW1,SW2,APS,CKM}). As was mentioned, we assume that $\mu_1\sim\mu_2\sim\mu$. In this case the weak
coupling condition reduces to
\begin{equation}
\sqrt{\mu m}\gg\Lambda_{{\mathcal N}=2}\,,
\label{weakcoup}
\end{equation}
where $\Lambda_{{\mathcal N}=2}$ is the scale of the ${\mathcal N}=2\;$ theory, and we assume that all quark masses are of the same order $m_A\sim m$.
In particular, the condition (\ref{weakcoup}), combined with the condition (\ref{smallmu}) of smallness
of $\mu$, implies that the average quark mass $m$ is very large.
\section{Duality at small $\mu$ in the quark vacua }
\label{N2duality}
\setcounter{equation}{0}
\subsection{Dual theory}
Now we will relax the condition (\ref{weakcoup}) and pass
to the strong coupling domain at
\begin{equation}
|\sqrt{\xi_P}|\ll \Lambda_{{\mathcal N}=2}\,, \qquad | m_{A}|\ll \Lambda_{{\mathcal N}=2}\,.
\label{strcoup}
\end{equation}
${\mathcal N}=2\;$ SQCD with the Fayet--Iliopoulos term (introduced as the $D$-term) was shown \cite{SYdual,SYtorkink}
to undergo a
crossover transition upon reduction of the
FI parameter. The results obtained in \cite{SYdual} are based on
studying the Seiberg--Witten curve \cite{SW1,SW2,APS} in ${\mathcal N}=2\;$ SQCD on the Coulomb branch, and,
therefore, do not depend on the type of the FI deformation. We briefly review these results here
adjusting our consideration \cite{SYdual,SYtorkink} to fit the case of the Fayet--Iliopoulos $F$-term
induced by the adjoint mass $\mu$.
The domain (\ref{strcoup})
can be described in terms of weakly coupled (infrared free) dual theory
with with the gauge group
\begin{equation}
{\rm U}(\tilde N)\times {\rm U}(1)^{N-\tilde N}\,,
\label{dualgaugegroup}
\end{equation}
and $N_f$ flavors of light {\em dyons}.\footnote{ Previously the SU$(\tilde N)$
gauge group was identified as dual \cite{APS} on the Coulomb branch
at the root of the baryonic Higgs branch in the ${\mathcal N}=2\;$ supersymmetric SU($N$) Yang--Mills
theory with massless quarks.}
\vspace{2mm}
Light dyons $D^{lA}$
\begin{equation}
l=1, ... ,\tilde N\,, \quad A=1, ... , N_f
\label{ldnumb}
\end{equation}
are in
the fundamental representation of the gauge group
SU$(\tilde N)$ and are charged under the Abelian factors indicated in Eq.~(\ref{dualgaugegroup}).
In addition, there are $(N-\tilde N)$
light dyons $D^J$ ($J=\tilde N+1, ..., N$), neutral under
the SU$(\tilde N)$ group, but charged under the
U(1) factors.
In Appendix A we present the low-energy effective action for
the dual theory in a specific
example: $N=3$, $N_f=5$, and $\tilde N=2$. In particular, starting from this action, we find
the dyon condensates in the quasiclassical approximation.
Generalization of these results to arbitrary $N$ and $\tilde N$ has the following form
\begin{eqnarray}
\langle D^{lA}\rangle \! \! &=& \langle \bar{\tilde{D}}^{lA}\rangle =
\!\!
\frac1{\sqrt{2}}\,\left(
\begin{array}{cccccc}
0 & \ldots & 0 & \sqrt{\xi_{1}} & \ldots & 0\\
\ldots & \ldots & \ldots & \ldots & \ldots & \ldots\\
0 & \ldots & 0 & 0 & \ldots & \sqrt{\xi_{\tilde N}}\\
\end{array}
\right),
\nonumber\\[4mm]
\langle D^{J}\rangle &=& \langle\bar{\tilde{D}}^{J}\rangle=\sqrt{\frac{\xi_J}{2}},
\qquad J=\tilde N +1, ..., N\,.
\label{Dvev}
\end{eqnarray}
The most important feature apparent in (\ref{Dvev}), as compared to the squark VEVs of the
original theory (\ref{qvev}), is a ``vacuum jump'' \cite{SYdual},
\begin{equation}
(1, ... ,\, N)_{\sqrt{\xi}\gg \Lambda_{{\mathcal N}=2}} \to (N+1, ... , \,N_f,\,\,\tilde N+1, ... ,\, N)_{\sqrt{\xi}\ll \Lambda_{{\mathcal N}=2}}\,.
\label{jump}
\end{equation}
In other words, if we pick up the vacuum with nonvanishing VEVs of the first $N$ quark flavors
in the original theory at large $\xi$, Eq.~(\ref{model}), and then reduce $\xi$ below
$\Lambda_{{\mathcal N}=2}$,
the system goes through a crossover transition and ends up in the vacuum of the {\em dual} theory with
the nonvanishing VEVs of $\tilde N$ last dyons (plus VEVs of $(N-\tilde N)$ SU$ (\tilde N)$ singlets).
The Fayet--Iliopoulos parameters $\xi_P$ in (\ref{Dvev}) are determined by the
quantum version of the classical expressions
(\ref{xis}). They are expressible via the roots of the Seiberg--Witten curve
in the given $r=N$ vacuum \cite{SYfstr}.
Namely,
\begin{equation}
\xi_P=2\,\left\{\sqrt{\frac{2}{N}}\,\mu_1\,\hat{m}-\mu_2(\sqrt{2}e_P+\hat{m})\right\},
\label{qxis}
\end{equation}
where $e_P$ are the double roots of the Seiberg--Witten curve \cite{APS},
\begin{equation}
y^2= \prod_{P=1}^{N} (x-\phi_P)^2 -
4\left(\frac{\Lambda}{\sqrt{2}}\right)^{N-\tilde N}\, \,\,\prod_{A=1}^{N_f} \left(x+\frac{m_A}{\sqrt{2}}\right),
\label{curve}
\end{equation}
while $\phi_P$ are gauge invariant parameters on the Coulomb branch. We recall that $\hat{m}$
in Eq.~(\ref{qxis}) is still the average of the first $N$ quark masses (\ref{avm}).
The curve (\ref{curve}) describes the Coulomb branch of the theory for $\tilde N<N-1$. The case $ \tilde N=N-1$ is special. In this case we must make a shift in Eq.~(\ref{curve}) \cite{APS},
\begin{equation}
m_A\to \tilde{m}_A=m_A+\frac{\Lambda_{{\mathcal N}=2}}{N}, \qquad \tilde N=N-1\,.
\label{shift}
\end{equation}
We will not consider this special case at large $\mu$ since it is incompatible with the
condition $N_f<3/2\, N$ or $\tilde N< N/2$.
In the $r=N$ vacuum the curve (\ref{curve}) has $N$ double roots and reduces to
\begin{equation}
y^2= \prod_{P=1}^{N} (x-e_P)^2,
\label{rNcurve}
\end{equation}
where quasiclassically (at large masses) $e_P$'s are given by the
mass parameters, $\sqrt{2}e_P\approx -m_P$, $P=1,...,N$.
As long as we keep $\xi_P$ and masses small enough (i.e. in the domain (\ref{strcoup}))
the coupling constants of the
infrared free dual theory (frozen at the scale of the dyon VEVs) are small:
the dual theory is at weak coupling.
At small masses, in the region (\ref{strcoup}), the double roots of the Seiberg--Witten
curve are
\begin{equation}
\sqrt{2}e_I = -m_{I+N}, \qquad
\sqrt{2}e_J = \Lambda_{{\mathcal N}=2}\,\exp{\left(\frac{2\pi i}{N-\tilde N}J\right)}
\label{roots}
\end{equation}
for $\tilde N<N-1$, where
\begin{equation}
I=1, ... ,\tilde N\,\,\,\, {\rm and} \,\,\,\, J=\tilde N+1, ... , N\,.
\label{d1}
\end{equation}
In particular, the $\tilde N$ first roots are determined by the masses of the
last $\tilde N$ quarks --- a reflection of the fact that the
non-Abelian sector of the dual theory is not asymptotically free and is at weak coupling
in the domain (\ref{strcoup}).
From Eqs.~(\ref{Dvev}), (\ref{qxis}) and (\ref{roots}) we see that the VEVs of the non-Abelian
dyons $D^{lA}$ are determined by $\sqrt{\mu m}$ and are much smaller
than the VEVs of the Abelian dyons $D^{J}$ in the domain
(\ref{strcoup}). The latter are of the order of
$\sqrt{\mu \Lambda_{{\mathcal N}=2}}$.
This circumstance is most crucial for our analysis in this paper. It will allow us to eventually increase $\mu$
and decouple the adjoint fields without spoiling the weak coupling condition in the dual theory,
see Sec.~\ref{largemu}.
In the special case $\tilde N=N-1$ masses in (\ref{roots}) should be shifted according to
(\ref{shift}).
Now, let us consider either equal quark masses or the mass choice (\ref{masssplit}).
Both, the gauge group and the global flavor SU($N_f$) group, are
broken in the vacuum. However, the color-flavor locked form inherent to (\ref{Dvev})
under the given mass choice guarantees that the diagonal
global SU($\tilde N)_{C+F}$ symmetry survives. More exactly, the unbroken {\em global} group of the dual
theory is
\begin{equation}
{\rm SU}(N)_F\times {\rm SU}(\tilde N)_{C+F}\times {\rm U}(1)\,.
\label{c+fd}
\end{equation}
The SU$(\tilde N)_{C+F}$ factor in (\ref{c+fd}) is a global unbroken color-flavor rotation, which involves the
last $\tilde N$ flavors, while the SU$(N)_F$ factor stands for the flavor rotation of the
first $N$ dyons.
Thus, a color-flavor locking takes place in the dual theory too. Much in the same way as
in the original theory, the presence of the global SU$(\tilde N)_{C+F}$ group
is the reason behind formation of the non-Abelian strings.
For generic quark masses the global symmetry (\ref{c+f}) is broken down to
U(1)$^{N_f-1}$.
In the equal mass limit, or given the mass choice (\ref{masssplit}),
the global unbroken symmetry (\ref{c+fd}) of the dual theory at small
$\xi$ coincides with the global group (\ref{c+f}) which manifest in the
$r=N$ vacuum of the original theory at large
$\xi$. This has been already announced previously.
Note, however, that this global symmetry is realized in two very distinct ways in the dual pair at hand.
As was already mentioned, the quarks and U($N$) gauge bosons of the original theory at large $\xi$
come in the $(1,1)$, $(N^2-1,1)$, $(\bar{N},\tilde N)$, and $(N,\bar{\tilde N})$
representations of the global group (\ref{c+f}), while the dyons and U($\tilde N$) gauge
bosons form $(1,1)$, $(1,\tilde N^2-1)$, $(N,\bar{\tilde N})$, and
$(\bar{N},\tilde N)$ representations of (\ref{c+fd}). We see that the
adjoint representations of the $(C+F)$
subgroup are different in two theories. How can this happen?
The quarks and gauge bosons
which form the adjoint $(N^2-1)$ representation
of SU($N$) at large $\xi$ and the dyons and gauge bosons which form the adjoint $(\tilde N^2-1)$ representation of SU($\tilde N$) at small $\xi$ are, in fact, {\em distinct} states.
The $(N^2-1)$ adjoints of SU($N$) become heavy
and decouple as we pass from large to small $\xi$
along the line $\xi\sim \Lambda_{{\mathcal N}=2}$. Moreover, some
composite $(\tilde N^2-1)$ adjoints of SU($\tilde N$), which are
heavy and invisible in the low-energy description at large $\xi$ become light
at small $\xi$ and form the $D^{lK}$ dyons
($K=N+1,...,N_f$) and gauge bosons of U$(\tilde N)$. The phenomenon of
the level crossing
takes place (Fig.~\ref{figevol}). Although this crossover is smooth in the full theory,
from the standpoint of the low-energy description the passage from large to small $\xi$ means a dramatic change: the low-energy theories in these domains are
completely
different; in particular, the degrees of freedom in these theories are different.
\begin{figure}
\epsfxsize=7cm
\centerline{\epsfbox{Aevol.eps}}
\caption{\small Evolution of the SU$(N)$ and SU$(\tilde N)$ gauge bosons and light quarks (dyons) vs. $\xi$.
}
\label{figevol}
\end{figure}
This logic leads us to the following conclusion. In addition to light dyons and gauge bosons
included in the low-energy theory at small $\xi$ we must have
heavy fields which form the adjoint representation $(N^2-1,1)$ of the
global symmetry (\ref{c+fd}). These are screened quarks
and gauge bosons from the large-$\xi$ domain.
Let us denote them as $M_P^{P'}$ (here $P , P'=1, ... , N$).
As has been already noted in Sec.~\ref{intro},
at small $\xi$ they decay into the monopole-antimonopole
pairs on the curves of marginal stability (CMS).\footnote{An explanatory remark regarding
our terminology is in order. Strictly speaking,
such pairs can be formed by monopole-antidyons and
dyon-antidyons as well, the dyons carrying root-like electric charges.
In this paper we refer to all such states collectively as to
``monopoles." This is to avoid confusion with dyons which appear in Eq.~(\ref{Dvev}). The
latter dyons carry weight-like electric charges and, roughly speaking, behave as
quarks, see \cite{SYdual} for further details.}
This is in accordance with results obtained
for ${\mathcal N}=2\;$ SU(2) gauge theories \cite{SW1,SW2,BF} on the Coulomb branch at zero $\xi$,
while for the theory at hand it is proven in \cite{SYtorkink}.
The general rule is that the only states which exist at strong coupling inside CMS are those which can become massless on the Coulomb branch
\cite{SW1,SW2,BF}. For our theory these are light dyons shown in Eq.~(\ref{Dvev}),
gauge bosons of the dual gauge group and monopoles.
At small nonvanishing $\xi$ the
monopoles and antimonopoles produced in the decay process of the adjoint $(N^2-1,1)$ states
cannot escape from
each other and fly off to asymptotically large separations
because they are confined. Therefore, the (screened) quarks or gauge bosons
evolve into stringy mesons $M_P^{P'}$ ($P,P'=1, ..., N$) in the strong coupling domain of small $\xi$
-- the monopole-antimonopole
pairs connected by two strings \cite{SYdual,SYtorkink}, as shown in Fig.~\ref{figmeson}.
By the same token, at large $\xi$, in addition to the light quarks and gauge bosons,
we have heavy fields $M_K^{K'}$ (here $K, K'=N+1, ... , N_f$), which form the
adjoint $(\tilde N^2-1)$ representation of SU($\tilde N$).
This is schematically depicted in Fig.~\ref{figevol}.
The $M_K^{K'}$ states are (screened) light
dyons and gauge bosons of the dual theory. In the large-$\xi$ domain they decay into
the monopole-antimonopole
pairs and form stringy mesons \cite{SYdual} shown in Fig.~\ref{figmeson}.
\subsection{More on the non-Abelian strings and confined mono\-poles}
\label{stringsN2}
Since dyons develop VEVs in the $r=N$ vacuum which break
the gauge group, see (\ref{Dvev}), our dual theory supports strings.
In fact, the minimal stings in our theory are the $Z_N$ strings, progenitors of
the non-Abelian strings \cite{HT1,ABEKY,SYmon,HT2}.
At generic $m_A$ the dual gauge group (\ref{dualgaugegroup}) reduces to U(1)$^N$;
the low-energy theory is U(1)$^N$
gauge theory with the Fayet--Iliopoulos $F$-term for each U(1) factor. The $Z_N$ strings for this
theory are thoroughly studied in \cite{SYfstr}. In the low-energy approximation the $Z_N$ strings are BPS saturated.
Tensions of all $N$ elementary $Z_N$ strings are given by the FI parameters \cite{SYfstr},
\begin{equation}
T^{\rm BPS}_{P}=2\pi|\xi_P|, \qquad P=1,...,N\,.
\label{ten}
\end{equation}
In the limit (\ref{masssplit}) the color-flavor locking takes place and the
global group of the dual theory becomes
that of Eq.~(\ref{c+fd}). In this case $\tilde N$ of the set of $N$ $Z_N$ strings (associated with
windings of non-Abelian $D^{lA}$ dyons) acquire orientational
zero modes and become non-Abelian. They can be analyzed within the general framework developed in
\cite{HT1,ABEKY,SYmon,HT2}, see \cite{SYrev} for a review.
The internal dynamics of
the orientational zero modes are described by two-dimensional
${\mathcal N}=(2,2)\;$ supersymmetric CP model living on the string world sheet \cite{HT1,ABEKY,SYmon,HT2}.
For the original theory (\ref{model}) it is CP$(N-1)$ model for $\tilde N =0$. For nonzero $\tilde N$ the string
becomes semilocal. Semilocal strings do not have fixed transverse radius, they acquire
size moduli, see \cite{AchVas} for a review of
the Abelian semilocal strings. The non-Abelian semilocal strings in ${\mathcal N}=2\;$ SQCD with $N_f>N$ were studied in \cite{HT1,HT2,SYsem,Jsem}.
The internal dynamics of these strings is {\em qualitatively} described by a weighted
CP$(N_f-1)$ model with $N$ positive and $\tilde N$ negative charges associated with $N$
orientational moduli and $\tilde N$ size moduli. (Aspects of a quantitative description and
its interrelation with the weighted
CP$(N_f-1)$ model will be discussed in \cite{SYV}.)
In the dual theory $N$ and $\tilde N$ interchange; it is governed by
the weighted CP$(N_f-1)$ model with $\tilde N$ positive (orientations) and $N$ negative (size) charges
$n_K$, $K=(N+1),...,N_f$ and $\rho_P$, $P=1,...,N$, respectively \cite{SYtorkink}.
The above moduli are subject to the constraint
\begin{equation}
|n_K|^2 - |\rho_P|^2=2 \tilde{\beta},
\label{unitvec}
\end{equation}
where $\tilde{\beta}$ is a coupling constant of the dual world-sheet theory. It is determined by
the gauge coupling constant of the dual bulk theory at the scale $\sim \sqrt{\xi}$
\cite{ABEKY,SYmon,SYtorkink},
\begin{equation}
\tilde{\beta}= \frac{2\pi}{\tilde{g}_2^2}\,.
\label{betag}
\end{equation}
Distinct elementary non-Abelian strings
correspond to different vacua of the CP model under consideration.
Confined monopoles of the bulk theory are identified with the junctions of
two degenerate elementary non-Abelian strings \cite{T,SYmon,HT2}.
These are seen as kinks interpolating between different vacua of the CP model. Non-perturbative
generation of the dynamical scale $\Lambda_{CP}$ in
the CP model stabilizes these kinks in the non-Abelian regime,
making their inverse sizes and masses of the order of $\Lambda_{CP}$ \cite{SYmon,SYrev}.
Thus, the notion of the confined monopole becomes well-defined in the non-Abelian limit.
The identification of confined monopoles with the CP-model kinks reveals the global
quantum numbers of the monopoles. Say, it was known for a long time that
the kinks in the quantum limit form a fundamental representation
of the global SU$(N)$ group in the ${\mathcal N}=(2,2)\;$ supersymmetric CP$(N-1)\;$ models \cite{W79,HoVa}.
In \cite{SYtorkink} we generalized this result to the case of the
${\mathcal N}=(2,2)\;$ supersymmetric weighted CP models. We showed that
the kinks (confined monopoles) are in the fundamental representation of the global group (\ref{c+fd}).
More exactly, in the limit (\ref{masssplit}) they
form the $(N,1)+(1,\tilde N)$ representations of the global group (\ref{c+fd}).
This means that the total number of stringy mesons $M_A^B$ formed by
the monopole-antimonopole
pairs connected by two different elementary non-Abelian strings (Fig.~\ref{figmeson}) is $N_f^2$.
The mesons $M_P^{P'}$ form the
singlet and $(N^2-1,1)$ adjoint representations of the global group (\ref{c+fd}),
the mesons $M_P^{K}$ and $M_K^{P}$ form bifundamental representations
$(N,\bar{\tilde N})$ and $(\bar{N},\tilde N)$, while the mesons
$M_K^{K'}$ form the singlet and $(1,\tilde N^2-1)$ adjoint representations.
(Here, as usual, $P=1,...,N$ and $K=(N+1) , ... , N_f$.)
All these mesons with not too high spins have masses
\begin{equation}
m_{M_P^{P'}} \sim \sqrt{\xi}\, ,
\label{mstringyN2}
\end{equation}
as determined by the string tensions (\ref{ten}) .
They are heavier than the
elementary states, namely, dyons and dual gauge bosons which form
the (1,1), $(N,\bar{\tilde N})$, $(\bar{N},\tilde N)$,
and $(1,\tilde N^2-1)$ representations and have masses $\sim \tilde{g}_2\sqrt{\xi}$.
Therefore, the (1,1), $(N,\bar{\tilde N})$, $(\bar{N},\tilde N)$, and $(1,\tilde N^2-1)$ stringy mesons
decay into elementary states, and we are left with $M_P^{P'}$ stringy mesons
in the representation $(N^2-1,1)$.
Thus our confinement picture in the bulk theory outlined above is confirmed by the world-sheet analysis.
This concludes our extended introduction and adjustments necessary to pass to the study of
the ${\mathcal N}=1\;$ theories.
\section{Flowing to ${\mathcal N}=1\;$ QCD}
\label{largemu}
\setcounter{equation}{0}
With all preparatory work done, we begin our journey in the ${\mathcal N}=1\;$ theories.
In this section we increase the adjoint mass $\mu$ and decouple the adjoint matter. In the course of this process
the theory at hand flows to ${\mathcal N}=1\;$ SQCD. So, now we assume that
\begin{equation}
|\mu| \gg |m_A |, \qquad A=1, ... , N_f\,.
\label{mularge}
\end{equation}
Then, the ${\mathcal N}=2\;$ multiplets are split. We consider the
quark masses to be small enough to guarantee that the original theory (\ref{model})
is at strong coupling, while the dual theory is at weak coupling.
\subsection{Decoupling the U(1)$^{(N-\tilde N)}$ sector}
\label{41}
At first, we will impose the condition
\begin{equation}
|\mu |\ll \Lambda_{{\mathcal N}=2}\,,
\label{intermu}
\end{equation}
implying (in conjunction with (\ref{mularge})) that all
parameters $\sqrt{\xi_P}$ are much smaller than $\Lambda_{{\mathcal N}=2}$.
Then our dual theory is at weak coupling, see Eqs. (\ref{qxis}) and (\ref{roots}).
From (\ref{roots}) we see that
VEVs of non-Abelian dyons $D^{lA}$ are much smaller those of the Abelian dyons $D^J$. Consider
the low-energy limit of the dual theory, i.e. energies much lower than
$\sqrt{\mu\Lambda_{{\mathcal N}=2}}$. In this scale the Abelian dyons $D^J$
$$J=(\tilde N+1), ... , N$$
are heavy and decouple. These dyons interact with $(N-\tilde N +1)$ U(1) gauge fields, see
Eq.~(\ref{dualgaugegroup}). In this set of gauge bosons, $(N-\tilde N)$ U(1) fields also become heavy (with masses
$g\sqrt{\mu\Lambda_{{\mathcal N}=2}}$). Only one remains. As a result, in the low-energy limit we are
left with the dual theory with the gauge group
\begin{equation}
{\rm U}(\tilde N)
\label{dgg}
\end{equation}
and $N_f$ flavors of dyons $$D^{lA}\,,\quad l=1, ... ,\tilde N\,,\quad A=1, ... , N_f\,.$$
The superpotential in this theory is
\begin{equation}
{\mathcal W} = \sqrt{2}\,\sum_{A=1}^{N_f}
\left( \frac{1}{ 2}\,\tilde D_A b_{U(1)}
D^A + \tilde D_A b^p\,T^p D^A\right)\,
+{\mathcal W}_{[\mu ]}( b_{U(1)}, b^p),
\label{superpottN}
\end{equation}
Here $b_{U(1)}$ is a chiral superfield, the ${\mathcal N}=2\;$ superpartner of $B^{U(1)}_{\mu}$,
where $B^{U(1)}_{\mu}$ is a particular linear combination of the dual gauge fields not interacting
with the $D^J$ dyons. We renormalized $b_{U(1)}$ so that charges of the $D^{lA}$ dyons
with respect to this field
are $\frac{1}{2}$.
This amounts to redefining its coupling constant $\tilde{g}^2_{U(1)}$. Moreover,
$b^p$ is an SU($\tilde N$) adjoint chiral field, the ${\mathcal N}=2\;$ superpartner of the dual SU($\tilde N$) gauge field.
Finally, ${\mathcal W}_{[\mu ]}$ is a $\mu$ dependent part of the superpotential, cf. (\ref{msuperpotbr}).
The deformation superpotential ${\mathcal W}_{[\mu ]}$
given in Eq. (\ref{msuperpotbr}) can be expressed in terms
of invariants $u_k$, see Eq.~(\ref{u}). Namely,
\begin{equation}
{\mathcal W}_{[\mu ]}= \mu_2\,u_2 -\frac{\mu_2}{N}\,
\left(1-\sqrt{\frac2N}\,\frac{\mu_1}{\mu_2}\right)\,u_1^2\,,
\label{Wmuu}
\end{equation}
where $u_2$ and $u_1$ should be understood as functions of $b_{U(1)}$ and $b^p$.
These functions are determined by the exact Seiberg--Witten solution. We will treat them
in Sec. \ref{42}.
Note, that with the singlet dyons decoupled, the VEVs of the
non-Abelian dyons are
\begin{equation}
\langle D^{lA}\rangle \! \! = \langle \bar{\tilde{D}}^{lA}\rangle =
\!\!
\frac1{\sqrt{2}}\,\left(
\begin{array}{cccccc}
0 & \ldots & 0 & \sqrt{\xi_{1}} & \ldots & 0\\
\ldots & \ldots & \ldots & \ldots & \ldots & \ldots\\
0 & \ldots & 0 & 0 & \ldots & \sqrt{\xi_{\tilde N}}\\
\end{array}
\right),
\label{DvevN1}
\end{equation}
where the first $\tilde N$ $\xi$'s are of the order of $\mu m$, see (\ref{roots}).
\subsection{Decoupling the adjoint matter}
\label{42}
As will be shown in Sec.~\ref{43}, the masses of the gauge fields and dyons $D^{lA}$ in the U$(\tilde N$) gauge theory,
with the superpotential (\ref{superpottN}), do not exceed $\sqrt{\mu m}$, while
the adjoint matter mass of is of the order of $\mu$. Therefore, in the limit (\ref{mularge}) the adjoint matter
decouples. Below scale $\mu$ our theory becomes dual to ${\mathcal N}=1\;$ SQCD with the scale
\begin{equation}
\tilde{\Lambda}^{N-2\tilde N}= \frac{\Lambda_{{\mathcal N}=2}^{N-\tilde N}}{\mu^{\tilde N}}\,.
\label{tildeL}
\end{equation}
The only condition we impose to keep this infrared free theory in the weak coupling
regime is
\begin{equation}
\sqrt{\mu m} \ll \tilde{\Lambda}\,.
\label{wcdual}
\end{equation}
This means that at large $\mu$ we must keep the quark masses small enough. The
larger the value of $\mu$ the smaller the quark masses, so that the product $\mu m$
is constrained from above by $\tilde{\Lambda}^2$.
This is always doable.
We would like to stress that, although this procedure is perfectly justified in the $r=N$ vacuum we work in,
it does not work, say, in the
monopole vacua. In these vacua VEVs of the light matter
(the Abelian monopoles) are of the order of $\sqrt{\mu \Lambda_{{\mathcal N}=2}}$, which, in turn,
sets
the mass scale
in the dual Abelian U(1)$^{N}$ gauge theory \cite{SW1}. Therefore, we cannot decouple
the adjoint matter keeping the dual theory at weak coupling. As soon as we
increase $\mu$ well above the above scale, we break the weak coupling
condition in the dual U(1)$^{N}$ gauge theory.
In contrast, in the $r=N$ vacuum we can take $\mu$ much larger than the
quark masses and decouple the
adjoint matter. If the condition (\ref{wcdual}) is fulfilled, the dual theory stays at weak
coupling. The reason is that it is the quark masses rather
than $\Lambda_{{\mathcal N}=2}$ that determine the ``non-Abelian'' roots of
the Seiberg--Witten curve and VEVs of the non-Abelian dyons, see (\ref{roots}).
Given the superpotential (\ref{superpottN})
we can explicitly integrate out the adjoint matter. To this end we
expand ${\mathcal W}_{[\mu]}$ in powers of $b_{U(1)}$ and $b^p$,
\begin{eqnarray}
{\mathcal W}_{[\mu ]}( b_{U(1)}, b^p)
&= &
c_1 \, \mu_2\, b_{U(1)}^2 + c_2 \,\mu_2\, (b^p)^2
\nonumber\\[3mm]
&+& c_3 \, \mu_2 m \, b_{U(1)} + c_4 \, \mu_2 \,\Lambda_{{\mathcal N}=2}\, b_{U(1)}
\nonumber\\[3mm]
&+&
O\left(\frac{\mu_2\, (b^p)^4}{\Lambda^2_{{\mathcal N}=2}}\right) +
O\left(\frac{\mu_2\, b_{U(1)}^3}{\Lambda_{{\mathcal N}=2}}\right),
\label{supexpand}
\end{eqnarray}
where
\begin{equation}
m=\frac1{N_f}\,\, \sum_{A=1}^{N_f} m_A \,.
\label{m}
\end{equation}
We then note that
\begin{equation}
c_4=0\, .
\label{c40}
\end{equation}
Indeed, a nonvanishing $c_4$ would produce a VEV of $b_{U(1)}$ of the order of
$\Lambda_{{\mathcal N}=2}$ which, in turn, would imply VEVs of certain
dyons $D^{lA}$ to be of the order of $\sqrt{\mu\Lambda_{{\mathcal N}=2}}$, in direct contradiction with Eqs.
(\ref{DvevN1}) and (\ref{roots}).
Moreover, since VEVs of $b_{U(1)}$ and $b^p$ are of the order of the
quark masses (rather than $\Lambda_{{\mathcal N}=2}$) we can neglect higher-order
terms in the expansion (\ref{supexpand}) and keep only linear and quadratic
terms in the $b$ fields. Higher-order terms are suppressed by powers of $m/\Lambda_{{\mathcal N}=2}$.
Now, substituting (\ref{supexpand}) into (\ref{superpottN}) and integrating over
$b_{U(1)}$ and $b^p$ we get the superpotential which depends only on $D^{lA}$. Minimizing it and
requiring VEVs of $D^{lA}$ to be given by (\ref{DvevN1}) (see also (\ref{roots})) we fix the coefficients
$c_1$, $c_2$ and $c_3$,
\begin{equation}
c_1= \frac{\tilde N}{4}\left(1+\gamma\frac{\tilde N}{N}\right), \qquad c_2=\frac12,\qquad
c_3= \frac{\tilde N}{\sqrt{2}}\gamma\left(1+\frac{\tilde N}{N}\right),
\label{cs}
\end{equation}
where
\begin{equation}
\gamma= 1-\sqrt{\frac2N}\,\frac{\mu_1}{\mu_2}.
\label{gamma}
\end{equation}
After eliminating the adjoint matter the superpotential takes the form
\begin{eqnarray}
{\mathcal W} &=& -\frac1{2\mu_2}\,
\left[ (\tilde{D}_A D^B)(\tilde{D}_B D^A) - \frac{\alpha_{D}}{\tilde N} (\tilde{D}_A D^A)^2 \right]\,
\nonumber\\[3mm]
&+&
\left[m_A-\frac{\gamma\,(1+\frac{\tilde N}{N})}{1+\gamma\,\frac{\tilde N}{N}}\,m\right]\,(\tilde{D}_A D^A),
\label{superpotd}
\end{eqnarray}
where the color indices are contracted inside each parentheses, while
\begin{equation}
\alpha_{D}=\frac{\gamma\,\frac{\tilde N}{N}}{ 1+\gamma\,\frac{\tilde N}{N}}\,\,.
\label{alphaD}
\end{equation}
This equation presents
our final large-$\mu$ answer for the superpotential of the theory dual to ${\mathcal N}=1\;$ SQCD in the (1, ... , $N$) vacuum.
The second term is the dyon mass term while the first one describes the dyon interaction.
One can check that minimization of this superpotential leads to correct dyon VEVs, cf.
Eq.~(\ref{DvevN1}).
Of course,
the theory with the superpotential (\ref{superpotd})
possesses many other vacua in which different dyons (and different number of dyons) develop VEVs.
We consider only one particular vacuum here. As was explained in
Sec.~\ref{N2duality}, if we choose the (1, ... , $N$) vacuum in the original theory
above the crossover, then we end up in the $(0, ... ,0, N+1, ... ,N_f)$ vacuum in the dual
theory below the crossover, see (\ref{jump}). Vacua with different number of condensed $D$'s
seen in (\ref{superpotd}) are spurious. The reason is that if we start from an $r<N$ vacuum in the original theory
the dual gauge group (below the crossover) would be different from U$(\tilde N)$. Thus, the dual theory would not
be the U$(\tilde N)$ gauge theory of dyons $D^{lA}$ ($l=1,...,\tilde N$), with the superpotential (\ref{superpotd}).
Summarizing this section, we pass to the limit of large $\mu$ decoupling the adjoint matter
in the dual theory. This leaves us with the dual U$(\tilde N)$ gauge theory
with the superpotential (\ref{superpotd}). At this point one should ask:
Are we sure that $\mu$ is large enough
to decouple the adjoint matter in the original theory (\ref{model}), as well as in the dual theory,
so that the original theory becomes ${\mathcal N}=1\;$
SQCD?
Strictly speaking, it is not easy to directly answer this question since in the domain (\ref{wcdual})
the original theory is at strong coupling, and our control over its dynamics is limited.
Nevertheless, one can give the following argument. Let us denote
the low-energy scale of the original ${\mathcal N}=1\;$ SQCD as $\Lambda$.
In terms of the scale of the original theory (\ref{model}) at large $\mu$ we have
\begin{equation}
\Lambda^{2N-\tilde N}= \mu^{N}\,\Lambda_{{\mathcal N}=2}^{N-\tilde N}\,.
\label{Lambda}
\end{equation}
The (s)quark masses are small, and the scale of excitations in ${\mathcal N}=1\;$ SQCD
is determined by the parameter (\ref{Lambda}). The nonvanishing masses just
lift the Higgs branch making all vacua isolated. Therefore, if we require that
\begin{equation}
\mu\gg \Lambda
\label{largemuorig}
\end{equation}
we can be sure that the adjoint mater is decoupled in the original theory.
Now, the weak coupling condition for the dual theory (\ref{wcdual}) can be
rewritten in terms of $\Lambda$ as follows:
\begin{equation}
m\ll \Lambda\,\left(\frac{\Lambda}{\mu}\right)^{\frac{3N}{N-2\tilde N}}\, .
\label{wcdualL}
\end{equation}
Since the quark mass scale $m$ is at our disposal, we can always choose it to be sufficiently small.
Below we assume that both conditions (\ref{largemuorig}) and (\ref{wcdualL}) are met.
If we further increase $\mu$ (keeping the quark masses fixed) we hit the upper bound in (\ref{wcdualL})
and the dual theory (\ref{superpotd}) goes through a crossover into strong coupling.\footnote{To avoid this, one can
simultaneously decrease
$m$.} Still further increase of the parameter $\mu$, $$\sqrt{\mu m} \gg \Lambda\,,$$
brings us in the weak coupling regime in the original ${\mathcal N}=1\;$ SQCD.
In this regime the (s)quark fields condense thus completely Higgsing the U$(N)$ gauge group. Non-Abelian strings are formed which confine monopoles. This regime is quite similar to that studied in \cite{EdTo,SYhet,BSYhet} in
the massless version of the theory (\ref{model}), with a large Fayet--Iliopoulos $D$-term.
We stress
that in this domain (large $m$) the (s)quark fields condense, while in our present setup (small $m$)
the quarks and gauge bosons decay into the
monopole-antimonopole stringy mesons on CMS.
\subsection{Perturbative mass spectrum}
\label{43}
In this section we briefly discuss the perturbative mass spectrum of the dual U$(\tilde N)$
gauge theory, with the superpotential (\ref{superpotd}), at large $\mu$.
At first we assume the limit (\ref{masssplit}) for the quark masses.
The U$(\tilde N)$ gauge group is completely Higgsed, and the masses of the gauge bosons
are
\begin{equation}
m_{SU(\tilde N)}=\tilde{g}_2\sqrt{\xi}
\label{Wmassd}
\end{equation}
for the SU$(\tilde N)$ gauge bosons, and
\begin{equation}
m_{U(1)}=\tilde{g}_1\, \sqrt{\frac{N}{2}}\,\sqrt{\xi}\,.
\label{phmassd}
\end{equation}
for the U(1) gauge boson. Here $\tilde{g}_1$ and $\tilde{g}_2$ are dual gauge couplings
for U(1) and SU$(\tilde N)$ gauge bosons respectively, while $\xi$ is a common value of the first
$\tilde N$ $\xi_P$'s (see Eqs.~(\ref{qxis}) and (\ref{roots})),
\begin{eqnarray}
\xi &=& 2\,\left\{\sqrt{\frac{2}{N}}\,\mu_1\,\hat{m}+\mu_2(m_{\rm last}-\hat{m})\right\},
\qquad m_{\rm last}=m_K,
\nonumber\\[3mm]
K &=& (N+1),...,N_f\,.
\label{xi}
\end{eqnarray}
The dyon masses are determined by the $D$-term potential
\begin{equation}
V^{\rm dual}_D =
\frac{\tilde{g}^2_2}{2}
\left( \bar{D}_A T^p D_A -
\tilde{D}_A T^p \bar{\tilde{D}}^A \right)^2
+ \frac{\tilde{g}^2_1}{8}
\left(|D^A|^2 -|\tilde{D}_A|^2
\right)^2
\label{Dtermpot}
\end{equation}
and the $F$-term potential following from the superpotential (\ref{superpotd}). Diagonalizing the quadratic
form given by these two potentials we find that, out of $4\tilde N N_F$ real degrees of freedom
of the scalar dyons, $\tilde N^2$ are eaten by the Higgs mechanism, $(\tilde N^2-1)$ real scalar dyons have the same mass as the non-Abelian gauge fields, Eq.
(\ref{Wmassd}), while one scalar dyon has mass (\ref{phmassd}). These dyons are scalar superpartners
of the SU$(\tilde N)$ and U(1) gauge bosons in ${\mathcal N}=1\;$ massive vector supermultiplets, respectively.
Another $2(\tilde N^2-1)$
dyons form a $(1,\tilde N^2-1)$ representation of the global group (\ref{c+fd}). Their mass is as follows:
\begin{equation}
m_{(1,\tilde N^2-1)}=\frac{\xi}{\mu_2}=2 \left(m_{\rm last}-\gamma \hat{m}\right)\,,
\label{adjd}
\end{equation}
where $\xi$ is given in Eq. (\ref{xi}),
while two real singlet dyons have mass
\begin{equation}
m_{(1,\,1)}= \sqrt{\frac{N}{2}}\,\frac{\xi}{\mu_1}=2 \left(\hat{m} -
\sqrt{\frac{N}{2}}\,\frac{\mu_2}{\mu_1}\,\Delta m\right).
\label{singld}
\end{equation}
Masses of $4N\tilde N$ bifundamental fields are given by the mass split of $N$ first and $\tilde N$ last
quark masses, see (\ref{masssplit}),
\begin{equation}
m_{(\bar{N},\, \tilde N)}= \Delta m\,.
\label{bifundd}
\end{equation}
All these dyons are the scalar components of the ${\mathcal N}=1\;$ chiral multiplets.
We see that the masses of the gauge multiplets and those of chiral matter get a large split in the limit
of large $\mu$ and small $m_A$. Chiral matter become much lighter than
the gauge multiplets cf. \cite{SYnone,SYrev}.
Most important is the fact that in the theory (\ref{superpotd})
vacuum expectation values are developed by the light dyons, with masses given by (\ref{singld}) in the limit (\ref{masssplit}). Thus, we have an extreme type-I superconductivity in the vacuum of the dual theory.
For generic quark masses the perturbative excitation spectrum is rather complicated.
We summarize it here for a particular case
\begin{equation}
\sqrt{\frac{\tilde N}{2}}\,\tilde{g}_1=\tilde{g}_2\,, \qquad \gamma =0\,.
\label{singletr}
\end{equation}
The first condition means that (with our normalizations) the gauge couplings
in the SU$(\tilde N)$ and U(1) sectors are the same, while the last condition implies that we consider a single-trace deformation superpotential in (\ref{msuperpotbr}).
Under these conditions the masses of the gauge bosons $(A_{\mu})^k_l$ are
\begin{equation}
m_{\rm gauge}=\tilde{g}_2\sqrt{\frac{\xi_k +\xi_l}{2}}\,.
\label{gmassd}
\end{equation}
Moreover, $\tilde N^2$ real dyons have the same masses. They are the ${\mathcal N}=1\;$ superpartners of the
massive gauge bosons.
Another $2\tilde N^2$ real dyons form a $\tilde N\times\tilde N$ complex matrix. The masses of the elements of this matrix are
\begin{equation}
m_{KK'} = m_K+m_{K'},\qquad K,K'=(N+1),...,N_f\,.
\label{tN2dyonsmass}
\end{equation}
The remaining $4N\tilde N$ of dyons (which become bifundamentals in the limit (\ref{masssplit}))
have masses
\begin{equation}
m_{PK} = m_P-m_{K},\qquad P=1,...,N, \qquad K=(N+1),...,N_f\,.
\label{tNNdyonsmass}
\end{equation}
Again, we see that the dyons with masses (\ref{tN2dyonsmass}) and (\ref{tNNdyonsmass}) are much lighter than
the gauge bosons and their scalar superpartners. It is the diagonal elements of the dyon matrix with the masses (\ref{tN2dyonsmass}) that develop vacuum expectation values.
\section{Strings and confined monopoles at large $\mu$}
\label{largemustr}
\setcounter{equation}{0}
Since in the dual theory (\ref{superpotd}) the dyons develop vacuum expectation values, see Eq. (\ref{Dvev}),
this theory support strings. Consider the limit (\ref{masssplit}) in which the global color-flavor group
(\ref{c+fd}) is restored and these strings become non-Abelian.
As was discussed in Sec.~\ref{43}, the mass terms of those dyons that develop VEVs are much smaller
than the gauge boson masses in the dual gauge group U$(\tilde N)$. Therefore, we deal with the type-I superconductor.
A detailed discussion of the non-Abelian string solutions for this case will be presented elsewhere.
Here we briefly mention certain peculiar features of such strings.
These strings are not BPS-saturated; their profile functions satisfy second-order equations of motion.
These profile functions have logarithmic long-range tails formed by light dyonic scalars
with masses (\ref{adjd}) and (\ref{singld}), see \cite{Y99} where Abelian strings in the extreme
type-I superconducting vacuum were studied. The string tension in this regime is
\begin{equation}
T = \frac{4\pi |\xi |}{\log{(\tilde{g}\mu/m)}}\,,
\label{tentypeI}
\end{equation}
while their transverse sizes scale as
\begin{equation}
R \sim \frac{\log{(\tilde{g}\mu/m)}}{\tilde{g}\sqrt{\xi}} \,,
\label{R}
\end{equation}
with the logarithmic accuracy Here $\xi$ is given in Eq.~(\ref{xi}), and we assume that
$\tilde{g}_2\sim \tilde{g}_1\sim\tilde{g}$.
As was mentioned in Sec.~\ref{stringsN2},
the internal dynamics of the non-Abelian strings in
the ${\mathcal N}=2\;$ limit at small $\mu$ is qualitatively described by an ${\mathcal N}=(2,2)\;$ supersymmetric
weighted CP model \cite{HT1,ABEKY,SYmon,HT2}, see also \cite{SYV}.
In the dual bulk theory, the string world-sheet model is CP$(N_f-1)$
with $\tilde N$ positive charges associated with the orientational modes
and $N$ negative charges
associated with string's size moduli (the latter are specific for semilocal string)
\cite{HT1,HT2,SYsem,Jsem,SYdual,SYtorkink}.
At large $\mu$ the semilocal strings at hand are no longer BPS-saturated. Their
size moduli $\rho_P$ are lifted, and the string tends to shrink in
type-I superconductors and to expand in type-II superconductors \cite{Hind,AchVas}.
Remember, we deal with type I. Thus, the shrinkage of the semilocal strings
results in conventional local strings. They are stable. The size moduli of the semilocal strings acquire masses
of the order of
\begin{equation}
m_{\rho}\sim \frac{1}{\tilde{g}\sqrt{\xi}\, R^2}\sim \frac{\tilde{g}\sqrt{\xi}}{\log{(\tilde{g}\mu/m)}}\,,
\label{rhomass}
\end{equation}
cf. \cite{Hind}.
Then, the world-sheet theory effectively reduces to CP$(\tilde N-1)$ model which describes the orientational mode dynamics. In particular, as a matter of fact, the constraint (\ref{unitvec}) is replaced by
\begin{equation}
|n_K|^2 =2\tilde{\beta}\, .
\label{unitvecCP}
\end{equation}
Another feature of the non-Abelian strings in the extreme type-I superconductors
is that the coupling constant $\tilde{\beta}$ of the CP$(\tilde N-1)$ model becomes very large,
\begin{equation}
\tilde{\beta} \sim \frac{\tilde{g}^2\mu}{m}\,.
\label{betatypeI}
\end{equation}
This effect is due to the presence of a long-range tail in the string in the type-I superconductor.
Using the one-loop renormalization equation in the asymptotically free CP$(\tilde N-1)$ model
\begin{equation}
4\pi\tilde{\beta}(\xi) =
\tilde N \ln{\frac{\sqrt{\xi}}{\Lambda_{CP}}}
\label{d2coupling}
\end{equation}
we find that the CP$(\tilde N-1)$ model scale
becomes exponentially small,
\begin{equation}
\Lambda_{\rm CP}\sim {\sqrt{\xi}}\exp{\left(-{\rm const}\,\frac{\tilde{g}^2\mu}{m}\right)} \,.
\label{LambdaCP}
\end{equation}
\vspace{1mm}
Now, it is time discuss confined monopoles of the bulk theory corresponding to kinks
in the world-sheet CP model.
At large $\mu$ the
non-Abelian strings are no longer BPS-saturated, and, consequently,
the ${\mathcal N}=(2,2)\;$ supersym\-metry in the world-sheet CP
model is lost. Non-supersymmetric CP$(\tilde N-1)$ model no longer has $\tilde N$ degenerate vacua, the true vacuum is unique,
but the model has a family of quasi-vacua \cite{Wtheta,GSY05}. The splittings are of the order of $\Lambda_{CP}$.
Thus, $\tilde N$ different non-Abelian strings are split in their tensions. This implies
two-dimensional confinement of monopoles,
along the string \cite{GSY05}, in addition to their
permanent attachment to strings. The monopoles cannot move freely along the string. They are combined into
monopole-antimonopole pairs, the attraction is due to the fact that the string between the monopole and antimonopole
at hand has a slightly higher tension than the strings outside.
However, this effect is tiny (because of the small value of the
parameter $\Lambda_{\rm CP}$) and does
{\em not}
determine the distance between the monopole and antimonopole in the
stringy meson in Fig.~\ref{figmeson}. This distance is determined by
the classical string tension (\ref{tentypeI}) itself (and the kink masses),
rather than the tiny quantum differences between
the tensions of different non-Abelian strings. Therefore, we will ignore this effect, the tension splitting.
Another effect which affects the formation of monopole-antimonopole stringy mesons at large $\mu$
is the lifting of the size moduli of the semilocal string, see (\ref{rhomass}). Although the kinks that are in the
$(1,\tilde N)$ representation of the global group (\ref{c+fd}) are still light (their masses are of the order
of $\Lambda_{\rm CP}$), the kinks in the $(N,1)$ representation become heavier. We expect them to have masses
of the order of the masses of the $\rho$-excitations (see (\ref{rhomass})),
\begin{equation}
m^{{\rm kink}}_{(N,1)}\sim \frac{\tilde{g}\,\sqrt{\xi}}{\log{(\tilde{g}\mu/m)}}\,.
\label{mkink}
\end{equation}
These kinks (confined monopoles) form stringy mesons in the adjoint
representation of the SU$(N)$ subgroup of the global group.
We recall that the $(N^2-1,1)$ stringy mesons are
former (screened) quarks and gauge bosons of the original ${\mathcal N}=1\;$ SQCD.
As was already explained, below the crossover (at small $\sqrt{\mu m}$, see (\ref{wcdual})) the quarks and
gauge bosons decay into the monopole-antimonopole pairs and form stringy mesons $M_P^{P'}$,
$P=1, ... , N$, shown in Fig.~\ref{figmeson}.
From the kink mass formulas (\ref{rhomass}) and (\ref{tentypeI}) and
the string tension we expect the mass of the $M_P^{P'}$ mesons to be
\begin{equation}
m_{M_P^{P'}} \sim \sqrt{T}\,,
\label{mstringy}
\end{equation}
provided the meson spins are of the order of unity.
The masses of these stringy mesons are determined by the string tension, much in the same way as
in the ${\mathcal N}=2\;$ limit, see (\ref{mstringyN2}).
\section{Relation to Seiberg's duality}
\label{Seiberg}
\setcounter{equation}{0}
The last but not the least
topic to discuss
is the relation between our duality (and the
monopole confinement mechanism) at large $\mu$ and Seiberg's duality in ${\mathcal N}=1\;$ SQCD \cite{Sdual,IS}.
The
light dyons $D^{lA}$ of our U($\tilde N$) dual theory could be identified with Seiberg's ``dual quarks''.
This is natural since they carry the same quantum numbers: both are in fundamental representations of the dual gauge group U($\tilde N$) and the global flavor group SU$(N_f)$.
Moreover, the stringy mesons formed by the
monopole-antimonopole pairs correspond to Seiberg's neutral mesons $M_A^B$, $A,B=1,...,N_f$,
which are in the singlet or adjoint representations of global flavor group both in
our and Seiberg's dual descriptions of ${\mathcal N}=1\;$ SQCD. This conceptual similarity does not extend further, however.
There is a crucial distinction: in our dual theory the
stringy mesons are non-perturbative objects and are rather heavy, with masses
determined by the string tension, (\ref{mstringy}). The dual gauge bosons and in particular, dyons $D^{lA}$,
are much lighter,
see (\ref{Wmassd}), (\ref{phmassd}) and (\ref{adjd}), (\ref{singld}) respectively.
At the same time, in Seiberg's dual theory, the $M_A^B$ mesons
appear as fundamental fields at the Lagrangian level and are light.
As was already mentioned in Sec.~\ref{intro},
our understanding of these dramatic differences is that Seiberg's duality refers to
$N$ monopole vacua in which the meson fields $M_A^B$ condense, making the dyons (``dual quarks'')
heavy \cite{IS,IS2}. Let us briefly review how this happens. Consider the U$(\tilde N)$ version of Seiberg's dual theory with
the superpotential
\begin{equation}
{\mathcal W}_{S} =\sqrt{2}\,
(\tilde{D}_A D^B) M^A_B + \Lambda m_A\, M^A_A\,,
\label{seib}
\end{equation}
where we conjectured that Seiberg's ``dual quarks'' can be identified with our dyons $D^{lA}$.
Following \cite{IS,IS2}, we assume that the $M_A^B$ fields develop VEVs making dyons heavy and integrate dyons out.
The gluino condensation in the U$(\tilde N)$ gauge theory with no matter induces the superpotential
\begin{equation}
{\mathcal W}_{S}^{\rm eff} = \tilde N\,\Lambda^{\frac{2\tilde N -N}{\tilde N}} \left({\rm det}\,M\right)^{\frac{1}{\tilde N}} + \Lambda m_A\, M^A_A\,.
\label{seibsup}
\end{equation}
Strictly speaking the scale of Seiberg's dual theory should appear in the first term here. However,
this scale is estimated to be of the order of the scale of the original ${\mathcal N}=1\;$ QCD $\Lambda$ \cite{IS2},
and
in this estimate we do not distinguish between the two.
Minimizing this superpotential with respect to $M_A^B$ we find
\begin{equation}
\langle M\rangle \sim \Lambda^{\frac{N-\tilde N }{N}} m^{\frac{\tilde N}{N}}\,.
\label{MVEV}
\end{equation}
The presence of $N$ vacua in ${\mathcal N}=1\;$ SQCD is well-known and follows e.g. from Witten's index.
It is also known that these vacua
are continuously connected to $N$ monopole vacua of ${\mathcal N}=2\;$ SQCD through the
$\mu$
deformation \cite{KSS,EFGR,GVY,CKM}. Since the
``dual quarks'' do not condense in these vacua the non-asymptotically
free Seiberg's dual theory is in the Coulomb phase (``free dyonic phase''). This is true for energies
above the scale of the $M$-field VEVs (\ref{MVEV}). Below this scale, all dyons decouple and Seiberg's
dual theory becomes pure Yang-Milles theory with the U($\tilde N $) gauge group. It flows into the strong coupling, and
the SU($\tilde N $) sector becomes confining. The U(1) gauge factor remains unbroken.
In Fig.~\ref{figmuevol} we show schematically the evolution of different vacua versus $\mu$ at small $m$.
The vertical axis in this Figure corresponds to $\mu$, while the horizontal axis schematically represents
VEVs of various fields in the given vacuum.
At small $\mu$, near the Coulomb branch of ${\mathcal N}=2\;$ SQCD, we have the U(1)$^N$ Abelian gauge
theories in the $N$ monopole vacua. Condensation of the monopoles leads to formation of the
electric ANO strings and Abelian confinement of quarks \cite{SW1,SW2}. One U(1) factor remains unbroken.
At $\mu\sim \Lambda$ these vacua
go through a crossover into the non-Abelian phase. In the limit of infinite $\mu$ they are described
via Seiberg's dual theory. It is the U$(\tilde N)$ infrared-free non-Abelian gauge theory
with neutral mesonic fields described by the superpotential (\ref{seib}) \cite{Sdual,IS,IS2}. As was
reviewed above, the $M$ fields condense, and the theory is in the Coulomb phase for dyons $D^{lA}$.
\begin{figure}
\epsfxsize=10cm
\centerline{\epsfbox{Amuevol.eps}}
\caption{\small $\mu$-evolution of different vacua at small $m$.
$N$ monopole vacua are shown by thick solid lines
on the left, while $r=N$ vacuum is on the right. The ISS vacuum is shown by thick dashed line.
Gauge groups in different regimes are indicated as well as condensed or confined states.}
\label{figmuevol}
\end{figure}
Our dual theory applies to the $r=N$ quark vacuum of the original ${\mathcal N}=1\;$ SQCD, rather than
the monopole vacua. In the
strong coupling regime at small $m$ (described by a weakly coupled dual theory
in the domain (\ref{wcdual})) the
light dyons $D^{lA}$ condense in this vacuum, triggering formation of the
non-Abelian strings with confinement of monopoles ensuing automatically.
This vacuum has dyon condensate
proportional to $\sqrt{\mu m}$, see (\ref{DvevN1}), and represents a run-away vacuum not seen in Seiberg's
dual description, where $\mu$ is considered to be strictly infinite.
There exists a bunch of other ``hybrid" vacua in the theory, in which
at small $\mu$ we deal with $r<N$ quarks and some monopoles condensing.
All of them have unbroken U(1) gauge group \cite{Cachazo2}. We do not study them in this paper.
A few words about the relation of our $r=N$ vacuum to the Intriligator--Seiberg--Shih vacuum \cite{ISS}.
This vacuum looks rather similar to ours. The dyons $D^{lA}$ (Seiberg's ``dual quarks'') condense in both
of these vacua. However, clearly these vacua are different. In particular, for a generic choice of
the quark masses,
supersymmetry is broken in the ISS vacuum, while the
$r=N$ vacuum is supersymmetric. Also, the ISS vacuum has dyon VEVs
of the order of $\sqrt{m\Lambda}$ while in
the $r=N$ vacuum they are much larger, proportional to $\sqrt{\mu m}$. Still, the presence
of the $D^{lA}$ condensate indicates that the ISS vacuum could have physics similar to that in
our $r=N$ vacuum. In
particular, it could exhibit confinement of monopoles and a phenomenon similar to our
decay of quarks and gauge bosons of the original ${\mathcal N}=1\;$ SQCD
into the monopole-antimonopole stringy mesons.
Since the ISS vacuum
is not supersymmetric, it may not exist at all $\mu$. This would explain why we do not see this vacuum
in our dual theory (\ref{superpotd}) in the domain (\ref{wcdualL}).
We show this vacuum by the dashed line in Fig.~\ref{figmuevol}.
The fate of the ISS vacuum in the framework of our construction calls for further studies.
\section{Conclusions}
\label{conclu}
Let us summarize our findings. We started from our recent development of the
non-Abelian duality in the {\em quark vacua} of ${\mathcal N}=2\;$ super-Yang--Mills
theory with the U$(N)$ gauge group and $N_f$ flavors ($N_f>N$). The fact that $N_f>N$ is very crucial, as
will be emphasized below. The quark mass terms are introduced in a judiciously chosen way.
Instead of the Fayet--Iliopoulos term of the $D$ type, as previously, we introduce it through a superpotential (i.e. $F$ type).
We construct dual pairs.
Both theories from the dual
pair support non-Abelian strings which confine monopoles.
Next we undertake a next step, basically in the uncharted waters. we introduce an ${\mathcal N}=2\;$-breaking deformation,
a mass term $\mu{\mathcal A}^2$
for the adjoint fields. Our final goal is to make the adjoint fields heavy and thus pass to ${\mathcal N}=1\;$ SQCD.
Starting from a small deformation we eventually make it large which enforces
complete decoupling of the adjoint fields. We show that the ${\mathcal N}=2\;$ non-Abelian duality fully survives in
the limit of ${\mathcal N}=1\;$ SQCD, albeit some technicalities change. For instance, non-Abelian strings which used
to be BPS-saturated in the ${\mathcal N}=2\;$ limit, cease to be saturated in ${\mathcal N}=1\;$ SQCD.
They become strings typical of the extreme type-I superconducting regime.
Our
duality is a distant relative of Seiberg's duality in ${\mathcal N}=1\;$ SQCD. Both share common features
but have many drastic distinctions. This is due to the fact that Seiberg's duality apply to the monopole
rather than quark vacua.
More specifically, in our theory we deal with $N< N_f<\frac32 N $ massive quark flavors.
We consider the vacuum in which $N$ squarks condense.
Then we identify a crossover transition
from weak to strong coupling. At strong coupling we find a dual theory, U$(N_f-N)$ SQCD, with
$N_f$ light dyon flavors. The dual theory is at weak coupling provided $\mu m$ is small enough
(at large $\mu$ this requires taking $m$ to be rather small).
Condensation of light dyons
$D^{lA}$ in this theory triggers the formation of non-Abelian strings and confinement of monopoles.
Quarks and gauge bosons of the original ${\mathcal N}=1\;$ SQCD decay into the monopole-antimonopole pairs on CMS and
form stringy mesons shown in Fig.~\ref{figmeson}.
We would like to stress that the condition $\tilde N>1$ is crucial for our construction.
As was explained in Sec.~\ref{largemu}, the presence of the dual non-Abelian group allows us to increase $\mu$,
eventually
decoupling the adjoint field and, simultaneously, keeping the dual theory at weak coupling. The reason is that
we can take quark masses rather small to satisfy the weak coupling condition (\ref{wcdual}). If the dual
gauge group were Abelian, the light matter VEV's would be of the order of $\sqrt{\mu\Lambda_{{\cal N}=2}}$, hence
the theory would go into the strong coupling regime once we increase $\mu$ above
$\Lambda_{{\cal N}=2}$.
\section*{Acknowledgments}
The work of MS was supported in part by DOE
grant DE-FG02-94ER408.
The work of AY was supported
by FTPI, University of Minnesota,
by RFBR Grant No. 09-02-00457a
and by Russian State Grant for
Scientific Schools RSGSS-65751.2010.2.
\section*{Appendix: \\
U(3) theory with \boldmath{$N_f=5$} at small \boldmath{$\mu$}}
\renewcommand{\thesection.\arabic{equation}}{A.\arabic{equation}}
\setcounter{equation}{0}
\renewcommand{\thesubsection}{A.\arabic{subsection}}
\setcounter{subsection}{0}
In this Appendix
following \cite{SYdual} we consider specific example of U(3) gauge theory
with $N_f=5$ quark flavors (so that $N=3$, $\tilde N=2$) and present the low-energy dual
theory at small values of FI parameter, see (\ref{strcoup}). The gauge group (\ref{dualgaugegroup})
in this case has the form
\begin{equation}
{\rm U}(2)\times {\rm U}(1)_8\times {\rm U}(1)\,,
\label{U2dualgaugegroup}
\end{equation}
where U$(1)_8$ denotes a U(1) factor of the gauge group which is associated with
$T_8$ generator of the U(3) gauge group of the original theory.
The bosonic part of the effective low-energy
action of the theory in the domain (\ref{strcoup}) has the form
\begin{eqnarray}
S_{{\rm dual}} &=&\int d^4x \left[\frac{1}{4\tilde{g}^2_{2}}
\left(F^{p}_{\mu\nu}\right)^2
+\frac1{4g^2_1}\left(F_{\mu\nu}\right)^2 +\frac1{4\tilde{g}^2_8}\left(F^8_{\mu\nu}\right)^2
+\frac1{\tilde{g}^2_2}\left|\partial_{\mu}b^p\right|^2
\right.
\nonumber\\[4mm]
&+& \frac1{g^2_1}
\left|\partial_{\mu}a\right|^2 +\frac1{\tilde{g}^2_8}
\left|\partial_{\mu}b^8\right|^2
+\left|\nabla^1_{\mu}
D^A\right|^2 + \left|\nabla^1_{\mu} \tilde{D}_A\right|^2+
\left|\nabla^2_{\mu}
D^3\right|^2 + \left|\nabla^2_{\mu} \tilde{D}_{3}\right|^2
\nonumber\\[4mm]
&+&
\left. V(D,\tilde{D},b^p,b^8,a)\right]\,,
\label{SIII}
\end{eqnarray}
Here $B^{p}_{\mu}$ ($p=1,2,3$), $B^{8}_{\mu}$ and $A_{\mu}$ are gauge fields of (\ref{U2dualgaugegroup}),
while $F_{\mu\nu}^p$, $F_{\mu\nu}^8$ and $F_{\mu\nu}$ are their field strengths.
Their scalar ${\mathcal N}=2\;$ superpartners
$b^p$ and $b^8$ in terms of the fields of the original theory (\ref{model})
have the form
\begin{equation}
b^3= \frac{1}{\sqrt{2}}\,(a^{3}+a^{3}_D)\;\;\; {\rm for}\;\;\; p=3,
\qquad b^8= \frac{1}{\sqrt{10}}\,(a^{8}+3a^{8}_D),
\label{bbp}
\end{equation}
where subscript $D$ means dual scalar fields \cite{SW1,SW2},
while field $a$ is the same as in (\ref{model}).
Covariant derivatives are defined in accordance
with the charges of the $D^{lA}$ and $D^3$ dyons, see \cite{SYdual} for more details. Namely,
\begin{eqnarray}
\nabla^1_\mu & = &
=\partial_{\mu}-i\left(\frac12 A_{\mu}+\sqrt{2}\,B^p_{\mu}\frac{\tau^p}{2}
+\frac12\sqrt{\frac{10}{3}}\,B^8_{\mu}\right)\,,
\nonumber\\[3mm]
\nabla^2_\mu & = &
=\partial_{\mu}-i\left(\frac12 A_{\mu}
-\sqrt{\frac{10}{3}}\,B^8_{\mu}\right)\,.
\label{nablaD}
\end{eqnarray}
The coupling constants $g_1$, $\tilde{g}_8$ and $\tilde{g}_2$
correspond to two U(1) and the SU(2) gauge groups, respectively.
The scalar potential $V(D,\tilde{D},b^p,b^8,a)$ in the action (\ref{SIII})
is
\begin{eqnarray}
&& V(D,\tilde{D},b^p,b^8,a) =
\frac{\tilde{g}^2_2}{4}
\left( \bar{D}_A\tau^p D_A -
\tilde{D}_A \tau^p \bar{\tilde{D}}^A \right)^2
\nonumber\\[3mm]
&+& \frac{10}{3}\frac{\tilde{g}^2_8}{8}
\left(|D^A|^2 -|\tilde{D}_A|^2 -2|D^3|^2 +
2|\tilde{D}_3|^2 \right)^2
\nonumber\\[3mm]
&+& \frac{\tilde{g}^2_1}{8}
\left(|D^A|^2 -|\tilde{D}_A|^2 +|D^3|^2 -
|\tilde{D}_3|^2
\right)^2
\nonumber\\[3mm]
&+& \frac{\tilde{g}_2^2}{2}
\left| \sqrt{2}\tilde{D}_A \tau^p D_A +\sqrt{2}\,\,\frac{\partial{\mathcal W_{\mu}}}{\partial b^p}
\right|^2+
\frac{\tilde{g}^2_1}{2}\left| \tilde{D}_A D^A+
\tilde{D}_3 D_3 +\sqrt{2}\,\,\frac{\partial{\mathcal W}_{\mu}}{\partial a}\right|^2
\nonumber\\[3mm]
&+&
\frac{\tilde{g}_8^2}{2}\left| \sqrt{\frac{10}{3}}\tilde{D}_A D^A-
2\sqrt{\frac{10}{3}}\tilde{D}_3 D^3 + \sqrt{2}\,\,\frac{\partial{\mathcal W}_{\mu}}{\partial b^8}\right|^2
\nonumber\\[3mm]
&+&\frac12 \left\{ \left|(a+\tau^p\sqrt{2}\,b^p +\sqrt{\frac{10}{3}}\,b^8+\sqrt{2}m_A
)D^A\right|^2
\right.
\nonumber\\[3mm]
&+&
\left|(a+\tau^p\sqrt{2}\,b^p +\sqrt{\frac{10}{3}}\,b^8+\sqrt{2}m_A
)\bar{\tilde{D}}_A\right|^2
\nonumber\\[3mm]
&+&\left.
\left|\;a-2\sqrt{\frac{10}{3}}\,b^8+\sqrt{2}m_3 \;
\right|^2\left(|D^3|^2+|\tilde{D}_3|^2\right) \right\}\,.
\label{potIII}
\end{eqnarray}
\vspace{2mm}
The theory (\ref{SIII}) is at weak coupling in the domain (\ref{strcoup}) but
the derivatives of the superpotential (\ref{msuperpotbr}) entering in (\ref{potIII})
(which determine VEVs of dyons)
are rather complicated functions of fields $a$, $b^8$ and $b^p$. In \cite{SYfstr}
we used the exact
Seiberg-Witten solution of our theory to determine these derivatives in $r=N$
vacuum. Here we briefly review this calculation.
First we make a quantum generalization
\begin{equation}
\frac{\partial{\mathcal W}_{\mu}}{\partial b^p}\to\mu_2 \,\frac{\partial u_2}{\partial b^p}\,,
\qquad \frac{\partial{\mathcal W}_{\mu}}{\partial b^8}\to\mu_2 \,\frac{\partial u_2}{\partial b^8}\,,
\qquad \frac{\partial{\mathcal W}_{\mu}}{\partial a}\to\mu_1 \sqrt{\frac{2}{N}}\,\frac{\partial u_2}{\partial a}\,,
\label{dwda}
\end{equation}
where
\begin{equation}
u_k= \langle}\def\ket{\rangle {\rm Tr}\left(\frac12\, a + T^a\, a^a\right)^k\ket, \qquad k=1, ..., N\,,
\label{u}
\end{equation}
are gauge invariant parameters which describe the Coulomb branch.
To select the desired vacuum (1,2,3) ( which transforms into (4,5,3) vacuum below crossover) among all other vacua in the Seiberg-Witten
curve we require that the curve has the factorized form (\ref{rNcurve}), while
the double roots $e_P$ are semiclassically (at large masses) are given by mass
parameters, $\sqrt{2}e_P\approx -m_P$, $P=1,...,N$.
Using explicit expressions from \cite{ArFa,KLTY,ArPlSh,HaOz} which express derivatives
of $u_k$ with respect to scalar fields $a^a$ ($a=1,2,3$) of the original theory
(\ref{model}) and taking
into account monodromies which convert these derivatives into derivatives with respect to
$b^p$, $b^8$ and $a$ \cite{SYfstr,SYdual} we obtain
\begin{eqnarray}
\frac{\partial u_2}{\partial a}
&=&
e_1+e_2+e_3\,, \qquad \frac1{\sqrt{2}}\frac{\partial u_2}{\partial b^3}=e_1-e_2\,,
\nonumber\\[3mm]
\frac1{\sqrt{10}}\frac{\partial u_2}{\partial b^8}
&=&
\frac1{\sqrt{3}}(e_1+e_2-2e_3)\,,
\label{duda32}
\end{eqnarray}
where $e_P$ are double roots of the Seiberg-Witten curve (\ref{curve}) with shifted
masses (\ref{shift}) for the case $N=3$, $\tilde N=2$.
Vacua of the theory (\ref{SIII}) are determined by zeros of all $D$ and $F$-terms in
(\ref{potIII}). Using the derivatives of the superpotential (\ref{msuperpotbr})
obtained above we get the VEV's of dyons in the form
\begin{eqnarray}
\langle D^{lA}\rangle &=& \langle \bar{\tilde{D}}^{lA}\rangle =
\frac1{\sqrt{2}}\,
\left(
\begin{array}{ccccc}
0 & 0 & 0 & \sqrt{\xi_1} & 0\\
0 & 0 & 0 & 0 & \sqrt{\xi_2}\\
\end{array}
\right),
\nonumber\\[4mm]
\langle D^{3}\rangle &=& \langle\bar{\tilde{D}}^{3}\rangle = \sqrt{\frac{\xi_3}{2}},
\label{Dvev32}
\end{eqnarray}
where FI parameters $\xi_P$ are determined by (\ref{qxis}). The obvious
generalization of this formula to an arbitrary $N$ and $\tilde N$ gives Eq.~(\ref{Dvev}) quoted in the main text.
\newpage
\small
|
\section{Introduction}
Let $\mathbb{S}^{d}$ denote the unit sphere in the Euclidean space
$\mathbb{R}^{d+1}$. Suppose that it is insulated and has a total
positive charge of $+1$. In the absence of an external field the
charge will distribute uniformly with respect to the normalized
surface area measure (unit Lebesgue measure) $\sigma_d$. Now we
introduce a positive unit point charge exterior to the sphere that
repels the charge on the sphere in accordance with the Newton
potential $1/r^{d-1}$, where $r$ represents the distance between
point charges. If this point charge is very close to the sphere,
then one would expect it to cause a negatively charged spherical cap
to appear, while if the point charge is very far from the sphere its
influence is negligible and the charge on the sphere will be
everywhere positive and nearly uniformly distributed over the entire
sphere. We consider the following question: what is the smallest
distance from the unit point charge to $\mathbb{S}^d$ such that the
distribution of the positive charge on the sphere covers all of the
sphere? We will denote this critical distance by $\rho(d)$. As we
shall show, $1+\rho(d)$ equals the largest positive root of the
following polynomial equation of degree $2d-1$:
\begin{equation} \label{polynomial.G}
G(d;z) {:=} \left[ \left( z - 1 \right)^d - z - 1 \right] z^{d-1} +
\left( z - 1 \right)^d=0, \qquad d = 1, 2, 3, \dots .
\end{equation}
As the question above was communicated to the authors by A. A.
Gonchar, we shall refer to $G(d;z)$ as {\it Gonchar
polynomials}.
Rather surprisingly, $\rho(2)$ turns out to be the {\it Golden
ratio}, which is the limit of the ratio of successive terms in the
{\it Fibonacci sequence} $F_{n}=F_{n-1} +F_{n-2}$, $F_0=F_1=1$ and
$\rho(4)$ is the so-called {\it Plastic number} \cite{St1996}, which
is the limit of ratios of successive terms for the less known {\it
Padovan sequence} $P_n=P_{n-2}+P_{n-3}$, $P_0=P_1=P_2=1$ (sequence
A000931 in Sloane's OEIS~\cite{Sl2003}). That is,
\begin{align*}
\rho(2) &= \lim_{n\to \infty}\frac{F_{n+1}}{F_n}=\frac{1+\sqrt{5}}{2}=1.618033988\dots,\\
\rho(4) &= \lim_{n\to \infty}\frac{P_{n+1}}{P_n}=\frac{( 9 -
\sqrt{69} )^{1/3} + ( 9 + \sqrt{69} )^{1/3}}{2^{1/3} 3^{2/3}} =
1.3247179572\dots.
\end{align*}
In addition to this curious coincidence, the polynomials $ G(d;z)$
exhibit rather fascinating properties with regard to their
irreducibility over the ring of polynomials with integer
coefficients, as well as the asymptotic behavior of their zeros. Our
goal in the next section is to show how the Gonchar polynomials are
derived and then, in Section 3, to explore some of their properties
and draw the reader's attention to some related conjectures.
\section{Signed Equilibrium}
A general charge distribution on $\mathbb{S}^d=\{\PT{x}\in R^{d+1}\,
:\, |\PT{x}|=1\}
$ will be modeled by a signed measure $\eta$
supported on $\mathbb{S}^d$ with $\eta(\mathbb{S}^d)=1$. The
corresponding {\it Newtonian potential and energy} are given by
\begin{equation}\label{SignedPot}
V^\eta (\PT{x})=\int \frac{1}{|\PT{x}-\PT{y}|^{d-1}}\, d
\eta(\PT{y}), \qquad \mathcal{I}(\eta)=\int V^\eta (\PT{x})\,
d\eta(\PT{x}).\nonumber
\end{equation}
For example, $V^\eta (\PT{x})$ is the familiar Coulomb potential
when $d=2$. For $\eta=\sigma_d$ we expect from the rotational
invariance that the potential $V^{\sigma_d}$ is constant on
$\mathbb{S}^d$ and on concentric spheres. This is a well-known fact in potential theory, but
for completeness we provide a proof below.
\begin{lem} \label{L1}
The potential $V^{\sigma_d}(\PT{x})$ satisfies the following relations
\begin{equation} \label{EquilPotSphere}
V^{\sigma_d}(\PT{x}) =
\begin{cases}
1 & \text{for $|\PT{x}|\leq 1$,} \\
1 / |\PT{x}|^{d-1} & \text{for $|\PT{x}| > 1$.}
\end{cases}
\end{equation}
\end{lem}
\begin{proof}
Indeed, for any fixed $\PT{x}$ the Newtonian kernel
$f(\PT{y})=|\PT{x}-\PT{y}|^{1-d}$ is a harmonic function
in $\mathbb{R}^{d+1}\setminus \{ \PT{x} \}$,
that is $\Delta f (\PT{y})=\sum_{i=1}^{d+1} f_{y_i y_i}
(\PT{y}) \equiv 0$. Therefore,
by the mean value property for harmonic functions we have
\begin{equation}
V^{\sigma_d}(\PT{x})=\int_{\mathbb{S}^d} f(\PT{y})\, d
\sigma_d(\PT{y})=f(\PT{0})= \frac{1}{|\PT{x}|^{d-1}} \qquad
\text{{for $|\PT{x}|>1$.}} \label{LebesgueSurfacePotential}
\end{equation}
Considering the sequence of functions $f_n (y)=|(n+1)\PT{x}/n-\PT{y}|^{1-d}$ and applying the monotone convergence theorem allows us to extend this to the unit sphere
giving $V^{\sigma_d}(\PT{x}) \equiv 1$ for $|\PT{x}|=1$. By the
maximum principle for harmonic functions, the potential
$V^{\sigma_d}$ is constant everywhere in the closed unit ball
(``Faraday cage effect''), since it assumes its extreme values
on the boundary $\mathbb{S}^d$, where it is constant. (One can also verify
this directly by applying the identity $| \PT{x} - \PT{y} |^2 = |
\PT{x} |^2 | \frac{\PT{x}}{| \PT{x} |^2} - \PT{y} |^2$ and
\eqref{LebesgueSurfacePotential}.)
\end{proof}
In the classical Coulomb case ($d=2$), a standard electrostatics
problem (see \cite[Ch.~2]{Ja1998}) is to find the charge
distribution on a charged, insulated, conducting sphere in the
presence of an external field (such as generated by a
positive point charge $q$ off the sphere). This motivates the
following definition.
\begin{defn} Given a continuous function $Q(\PT{x})$ on $\mathbb{S}^d$,
we call a signed measure $\eta_{Q}$ supported on $\mathbb{S}^d$
and of total charge $\eta_{Q}(\mathbb{S}^d)=1$ {\em a signed
equilibrium associated with $Q$} if its weighted Newtonian potential
is constant on $\mathbb{S}^d$; that is, for some constant $F_Q$,
\begin{equation}
V^{\eta_{Q}}(\PT{x}) + Q(\PT{x}) = F_Q \qquad \text{everywhere on
$\mathbb{S}^d$.} \label{signedeq}\nonumber
\end{equation}
\end{defn}
In the above definition, the function $Q$ is referred to as an {\em
external field}. Thus Lemma \ref{L1} establishes that $\sigma_d$ is
the (signed) equilibrium measure on $\mathbb{S}^d$ in the absence of
an external field. The uniqueness of the signed equilibrium is
settled by the following known proposition \cite[Lemma~23]{BrDrSa2009}, whose proof is presented in
the Appendix.
\begin{prop} \label{prop:uniqueness}
If a signed equilibrium $\eta_Q$ associated with an external field
$Q$ on $\mathbb{S}^d$ exists, then it is unique.
\end{prop}
Here, we are concerned with the external field generated by a
positive point charge of amount $q$ located at $\PT{a}=R\PT{p}$, $R>1$, where $\PT{p}$ is the North Pole of $\mathbb{S}^d$; that is,
$\PT{p}=(0,0,\dots,0,1)$. Such a field is given by
\begin{equation} \label{externalfield}
Q_{\PT{a},q}(\PT{x}) {:=} \frac{q}{|\PT{x}-\PT{a}|^{d-1}}, \qquad \PT{x}\in
\mathbb{R}^{d+1},\end{equation}
and the associated signed
equilibrium is described in the next result. {(For the general result, see \cite{BrDrSa2009}.)}
\begin{lem}
The signed equilibrium $\eta_{Q}$ associated with the external
field $Q = Q_{\PT{a},q}$ of \eqref{externalfield} is absolutely
continuous with respect to the (normalized) surface area measure on
$\mathbb{S}^d$; that is, $d \eta_Q(\PT{x}) =
\eta_{R,q}^\prime(\PT{x}) d \sigma_d(\PT{x})$, and its density is
given by
\begin{equation} \label{eta.Q}
\eta_{R,q}^\prime(\PT{x}) = 1 + \frac{q}{R^{d-1}} -
\frac{q\left(R^2-1\right)}{\left|\PT{x}-\PT{a}\right|^{d+1}}{,}
\qquad \PT{x}\in \mathbb{S}^d.
\end{equation}
\end{lem}
For the classical Coulomb case this relation is well-known from
elementary physics (cf. \cite[p.~61]{Ja1998}).
\begin{proof} We note first that if $|\PT{x}|>1$, then the function
$u(\PT{z}) {:=} |\PT{z}-\PT{x}|^{1-d}$ is harmonic in $|\PT{z}|\leq 1$,
and since the Poisson integral formula \cite{He2009} preserves harmonic functions, we have
\begin{equation} \label{PoissonIdentity} \frac{1}{|\PT{z}-\PT{x}|^{d-1}}=\int_{\mathbb{S}^d} \frac{1}{|\PT{y}-
\PT{x}|^{d-1}}\frac{1-|\PT{z}|^2}{|\PT{z}-\PT{y}|^{d+1}}\,d\sigma_d(\PT{y}), \qquad | \PT{z} | < 1.
\end{equation}
Using a monotone convergence theorem argument as in Lemma \ref{L1}
we extend \eqref{PoissonIdentity} to $|\PT{x}|=1$. We shall make use of the identity
$|R\PT{y}-\PT{x}|=|\PT{y}-R\PT{x}|$ for any $|\PT{x}|=|\PT{y}|=1$. Then from \eqref{PoissonIdentity} we obtain with $\PT{z}=\PT{p}/R$ and $| \PT{x} | = 1$,
\begin{equation}\label{PoissonComp}
\begin{aligned}
Q( \PT{x} ) = \frac{q}{|R\PT{p}-\PT{x}|^{d-1}}=&\frac{q}{|\PT{p}-R\PT{x}|^{d-1}}=\frac{q}{R^{d-1}|\PT{z}-\PT{x}|^{d-1}}=\int_{\mathbb{S}^d} \frac{q}{R^{d-1}|\PT{y}-
\PT{x}|^{d-1}}\frac{1-|\PT{z}|^2}{|\PT{z}-\PT{y}|^{d+1}}\,d\sigma_d(\PT{y})\\
=&\int_{\mathbb{S}^d} \frac{q}{|\PT{y}-
\PT{x}|^{d-1}}\frac{R^2-1}{|\PT{p}-R\PT{y}|^{d+1}}\,d\sigma_d(\PT{y})=
\int_{\mathbb{S}^d} \frac{1}{|\PT{y}-
\PT{x}|^{d-1}}\frac{q(R^2-1)}{|\PT{a}-\PT{y}|^{d+1}}\,d\sigma_d(\PT{y}).
\end{aligned}
\end{equation}
The Poisson integral formula applied to the constant function $u(\PT{z}) \equiv q/R^{d-1}$ and $\PT{z}=\PT{p}/R$ yields
\begin{equation*}
\frac{q}{R^{d-1}}=\int_{\mathbb{S}^d} \frac{q(R^2-1)}{|\PT{a}-\PT{y}|^{d+1}}\,d\sigma_d(\PT{y}).
\end{equation*}
This together with \eqref{PoissonComp} implies that for $d \nu {:=} \eta_{R,q}^\prime d \sigma_d$ we have
\begin{equation*}
\int_{\mathbb{S}^d}\, d \nu = 1 \qquad \text{and} \qquad V^{\nu}(\PT{x})+Q(\PT{x})=1+\frac{q}{R^{d-1}}, \quad \PT{x} \in \mathbb{S}^d,
\end{equation*}
which proves the lemma.
\end{proof}
\begin{rmk}
A surprising aspect of the equilibrium support, which varies with $R$, is illustrated in Figure~\ref{fig1} for $\mathbb{S}^2$ and $q=1$.
Writing $\eta_{Q_{\PT{a},1}} = \eta_{Q} = \eta_{Q}^+ - \eta_{Q}^-$ in terms of its Jordan decomposition, it is clear on geometrical and physical grounds that the support of $\eta_Q^+$ is a spherical cap centered at the South Pole, while its complement (a spherical cap centered at the North Pole) is the support of $\eta_Q^-$, see Figure~\ref{fig1a}. Although somewhat counter-intuitive, we see from Figure~\ref{fig1b} that, as the unit charge approaches the sphere ($R \to 1$), from a certain
distance on, the support of $\eta_Q^+$ occupies increasingly more of
the sphere. This phenomenon can be explained by the geometry of the sphere, where an increasingly 'needle-like'
negative sink forming at the North Pole is balanced by an increasingly
uniform positive part so that the total charge is always one.
\end{rmk}
\begin{figure}[h!t]
\begin{center}
\subfloat[]{\label{fig1a}\includegraphics[scale=.5]{etaQdpi720.pdf}}
\hfill
\subfloat[]{\label{fig1b}\includegraphics[scale=1]{positivepartA.pdf}}
\hfill
\caption{\label{fig1} Signed equilibrium $\eta_Q$ on $\mathbb{S}^2$
(left) and projection onto the polar axis of the positive part $\eta_Q^+$.}
\end{center}
\end{figure}
From formula \eqref{eta.Q} we observe that the minimum value of the density
$\eta_{R,q}^\prime(\PT{x})$ is attained at the North Pole $\PT{p}$,
and its value there is
\begin{equation}\label{eta:eq}
\eta_{R,q}^\prime(\PT{p}) = 1 + \frac{q}{R^{d-1}} -
\frac{q\left(R^2-1\right)}{\left|\PT{a}-\PT{p}\right|^{d+1}} = 1 +
\frac{q}{R^{d-1}} -
\frac{q\left(R^2-1\right)}{\left(R-1\right)^{d+1}}.
\end{equation}
This leads to the following theorem.
\begin{thm} \label{thm:Gonchar}
For the external field $Q_{\PT{a},q} (\PT{x})$ of
\eqref{externalfield} with $\PT{a}=R\PT{p}$, the signed equilibrium
is a positive measure on all of $\mathbb{S}^d$ if and only if $R\geq
R_q$, where $R_q$ is the unique (real) zero in $(1,+\infty)$ of the
polynomial
\begin{equation} \label{equation}
G(d,q;z) {:=} \left[ \left( z - 1 \right)^d / q - z - 1 \right] z^{d-1} + \left( z - 1 \right)^d.
\end{equation}
In particular, the solution to Gonchar's problem is given by
$\rho(d)=R_1 -1$.
\end{thm}
\begin{proof}
We use the fact that the support of $\eta_{Q_{\PT{a},q}}$ is all of $\mathbb{S}^d$ if and only if $\eta_{R,q}^\prime(\PT{p}) \geq 0$ in \eqref{eta:eq} or equivalently that $G(d,q;R) \geq 0$. Hence we seek the number $R > 1$ such that $G(d,q;R) = 0$. Observing that $G(d,q;1) < 0$ and $G(d,q;x) > 0$ for $x>1$ sufficiently large, there exists at least one value $R_q$ such that $G(d,q; R_q) = 0$. Moreover, this root in $(1, \infty)$ is unique as can be seen by applying Descartes' Rule of Signs to
\begin{equation*}
G(d,q; 1 + w) = \frac{1}{q}\sum_{m=0}^{d-1} \binom{d-1}{m} w^{m+d} - \sum_{m=0}^{d-1} \left[ \binom{d}{m} + \binom{d-1}{m} \right] w^m.
\end{equation*}
\end{proof}
Curiously, for $d=2$ and $d=4$ we obtain the Golden
ratio and the Plastic number as answers to Gonchar's problem as mentioned in the Introduction.
Furthermore, for large values of $d$ the asymptotic analysis (provided in the Appendix) shows that
\begin{equation} \label{AsymptoticEquation}
R_q = 2 + \left[ \log (3 q) \right] / d + \mathcal{O}(1/d^2) \qquad
\text{as $d\to\infty$.}
\end{equation}
The appearance of $2$ as the limit as $d\to\infty$ of the
critical distances $R_q$ ($q>0$ fixed) can be explained by studying the asymptotic behavior of
$G(d,q; R)$ as $d \to \infty$. Indeed, for $R \geq 2 + \varepsilon$ ($\varepsilon > 0$
fixed) it goes to $+\infty$ as $d\to\infty$, while for
$1 < R \leq 2 - \varepsilon$ ($\varepsilon > 0$ fixed) it goes
to $-\infty$, leaving the number $2$ as the only
candidate.
\section{The polynomials $G(d;z)$}
\label{sec2}
In the following we investigate the family of Gonchar polynomials $G(d,z) = G(d,1;z)$ given in \eqref{polynomial.G}. Aside from the solution to Gonchar's problem, these polynomials are interesting in themselves and their distinctive properties merit further study.
A polynomial $P$ with real coefficients is called {\em
(self-)reciprocal} if its {\em reciprocal polynomial} $P^*(z) {:=}
z^{\deg P} P(1/z)$ coincides with $P(z)$. In other words, the
coefficients of $z^k$ and of $z^{\deg P - k}$ in $P(z)$ are the
same. Notice that {\em $G(d; z)$ is self-reciprocal} for even $d$ since
\begin{equation*}
z^{2 d - 1} G(d; 1/z) = \left[ \left( 1 - z \right)^d - z - 1
\right] z^{d-1} + \left( 1 - z \right)^d.
\end{equation*}
Consequently, if $\zeta$ is a zero of $G(d; z)$, then so is $1 /
\zeta$ for even $d$. For odd $d$ we infer from
\begin{equation*}
G(d; z) + G^*(d; z) = \left[ 1 + ( -1 )^d \right] \left( z - 1
\right)^d \left( z^{d-1} + 1 \right) - 2 \left( z^d + z^{d-1}
\right) = - 2 \left( z^d + z^{d-1} \right)
\end{equation*}
that the coefficients of $z^k$ and $z^{2 d - 1 - k}$ in $G(d; z)$
sum to zero except for the 'innermost' pair.
\subsection{Factorizations and Irreducibility}
With the aid of symbolic computation programs one can find factorizations and check irreducibility of explicitly given polynomials. For $d = 1, 2, \dots, 7$, we thereby obtain the following factorizations over the integers of $G(d; z)$.
\begin{align*}
G(1; z) &= z-3, \\
G(2; z) &= (z+1) \left(z^2-3 z+1\right), \\
G(3; z) &= z^5-3 z^4+3 z^3-5 z^2+3 z-1, \\
G(4; z) &= (z+1) \left(z^3-3 z^2+2 z-1\right) \left(z^3-2 z^2+3 z-1\right), \\
G(5; z) &= z^9-5 z^8+10 z^7-10 z^6+5 z^5-7 z^4+10 z^3-10 z^2+5 z-1, \\
G(6; z) &= (z+1) \left(z^2-z+1\right) \left(z^8-6 z^7+15 z^6-21 z^5+21 z^4-21 z^3+15 z^2-6 z+1\right), \\
G(7; z) &= z^{13}-7 z^{12}+21 z^{11}-35 z^{10}+35 z^9-21 z^8+7 z^7-9 z^6+21 z^5-35 z^4+35 z^3-21 z^2+7 z-1. \\
\end{align*}
One can easily verify that $(z + 1)$ divides $G(d; z)$ if and only if $d$ is even. Furthermore, the factor $z^2 - z + 1$ arises in the following cases.
\begin{prop} \label{prop:cyclot.divisor}
The cyclotomic polynomial $z^2-z+1$ divides $G(d;z)$ if and only if $6$ divides $d$.
\end{prop}
\begin{proof}
Note that if $\zeta$ is a zero of $z^2-z+1$, then $\zeta^2=\zeta-1$ and $\zeta^3 = -1$. Using formula \eqref{polynomial.G}, it readily follows that $G(d;\zeta)=0$ whenever $d\equiv 0 \pmod{6}$ and $G(d;\zeta) \neq 0$ otherwise.
\end{proof}
A general irreducibility result has so far eluded the authors. The Eisenstein
criterion and, in general, reduction to finite fields, seem not to be effective tools for studying the polynomials $G(d;z)$. However, using Mathematica, we verified the following conjecture for $d$ up to $500$.
\begin{conj}
Set $\ell(d; z) \equiv 1$ for $d$ odd, $\ell(d; z) {:=} z + 1$ for $d$
even but not divisible by $6$, and $\ell(d; z) {:=} ( z + 1 ) (
z^2 - z + 1 )$ if $6$ divides $d$. Then $G(d; z) / \ell(d; z)$ is
irreducible over the rationals except for $d = 4$, $8$ and $12$.
\end{conj}
Regarding the exceptional cases we record that, in addition to $G(4; z)$ as given above,
\begin{align*}
G(8; z) &= (z+1) \left(z^4-3 z^3+3 z^2-3 z+1\right) \\
&\phantom{=}\times \left(z^{10}-6 z^9+16 z^8-24 z^7+24 z^6-21 z^5+24 z^4-24 z^3+16 z^2-6 z+1\right), \\
G(12; z) &= (z+1) \left(z^2-z+1\right) \left(z^6-4 z^5+5 z^4-3 z^3+5 z^2-4 z+1\right) \\
&\phantom{=}\times \big(z^{14}-8 z^{13}+29 z^{12}-62 z^{11}+85 z^{10}-77 z^9+48z^8 \\
&\phantom{=\times}-33 z^7+48 z^6-77 z^5+85 z^4-62 z^3+29 z^2-8
z+1\big).
\end{align*}
\subsection{Zeros of Gonchar polynomials}
Figure~\ref{fig5} illustrates some features of $G(d;z)$ on the
real line and suggests some of its general properties. Note the qualitatively
different behavior for even and odd $d$.
\begin{figure}[h!t]
\begin{center}
\includegraphics[scale=.85]{polygraphoddB.pdf}
\includegraphics[scale=.85]{polygraphevenA.pdf}
\caption{\label{fig5} The polynomial $G(d;z)$ on the real line for
$d=1,2,\dots,7$.}
\end{center}
\end{figure}
Depending on the parity of $d$, the polynomial equation $G(d; z) =
0$ has either one ($d$ odd) or three ($d$ even) real simple roots.
More precisely, the following holds.
\begin{prop} \label{prop:real.zeros}
If $d$ is odd, then $G(d; z)$ has precisely one real zero, which is simple and lies in the interval $(2,3]$. If $d$ is even, then $G(d; z)$ has exactly three real zeros: one at $-1$, one in the interval $(1/3, 1/2)$ and one in the interval $(2, 3)$; all these zeros are simple.
\end{prop}
\begin{proof}
Note that for $d \geq 2$ there holds:
\begin{equation*}
G(d; -1) = ( -2 )^d \left[ 1 - ( -1 )^d \right], \quad G(d; 0) = (
-1 )^d, \quad G(d; 1) = -2, \quad G(d; 2) = 1 - 2^d.
\end{equation*}
The assertion of the proposition is trivial for $G(1; z) = z - 3$. So let $d \geq 2$. Since $G(d;2) < 0$ and $G(d;3) > 0$, the polynomial $G(d;z)$ has at least one real zero in the interval $(2,3)$ and, as we observed in the proof of Theorem~\ref{thm:Gonchar}, this is its only zero on $[1, \infty)$ and must be simple. For odd $d$, each of the terms $[(x-1)^d-x-1]x^{d-1}$ and $(x-1)^d$ is negative for $x < 1$ and hence so is $G(d; x)$; thus $G(d;x)$ has no zeros outside $(2,3)$. For even $d$, the self-reciprocity of $G(d;x)$ implies that to each zero $\xi(d)$ of $G(d;x)$ in $[1,\infty)$ there is a zero $\xi^*(d) = 1 / \xi(d)$ in $(0,1]$. Consequently, by the first part of the proof, $G(d; x)$ has one and only one zero in $(0,1)$ and this zero is simple and lies in the interval $(1/3, 1/2)$. It remains only to consider the interval $(-\infty, 0]$. Clearly $G(d;-1) = 0$, and by computing $G^\prime(d;x)$ and analyzing its sign (in particular, $G^\prime(d;-1) > 0$), one can show that $G(d;x)$ is strictly increasing on $(-\infty,-1)$. Thus $G(d;x) < 0$ on $(-\infty, -1)$ and, by self-reprocity, $G(d;x) > 0$ on $(-1,0)$.
\end{proof}
Regarding the behavior as $d$ increases of the zeros in $(2,3)$, M. Lamprecht \cite{La2011} proved the following.
\begin{prop}
The zeros $\xi(d)$ of $G(d; z)$ in the interval $(2,3]$ form a
strictly monotonically decreasing sequence with limit point $2$
(compare with \eqref{AsymptoticEquation}).
\end{prop}
We now turn to the study of the complex zeros of $G(d;z)$. Observe that by Proposition~\ref{prop:real.zeros}, $G(d; z)$ has either $d - 1$ (if $d$ is odd) or $d - 2$ (if $d$ is even) pairs of complex conjugated zeros (counting multiplicity). In Figure~\ref{fig6} we have plotted the zeros of $G(d;z)$ for $d = 9, 10, 11,$ and $12$ along with the two unit circles $\mathcal{C}_0$, $\mathcal{C}_1$ centered respectively at $0$ and $1$. Notice that these circles intersect at the points $(1 \pm i \sqrt{3} ) / 2$ and, by Proposition~\ref{prop:cyclot.divisor}, these points are zeros of $G(d; z )$ if and only if $d$ is a multiple of $6$. The zeros of $G(d;z)$ seem to occur roughly into three categories: zeros close to $\mathcal{C}_0$ (indicated by $\times$), zeros close to $\mathcal{C}_1$ (indicated by $+$) and zeros close to the vertical line $x = 1 / 2$ (indicated by $\blacklozenge$). The numbers $N_1$, $N_2$, $N_3$ of zeros in each of the categories are listed in Table~1 for $d = 1, \dots, 12, 42$, from which it appears that these numbers are nearly the same. We will discuss this further in Section~\ref{sec:asymptotics} (see Conjecture~\ref{conj:unit.circle}).
Figure~\ref{fig6} suggests that for even $d$ the polynomial $G(d;z)$ may have zeros lying precisely on $\mathcal{C}_0$. This, in fact, is the case as we now prove.
\begin{figure}[h!t]
\begin{center}
\hfill
\begin{minipage}{0.7\linewidth}
\includegraphics[scale=.625]{RootsGd09q1.pdf} \hfill
\includegraphics[scale=.625]{RootsGd10q1.pdf}
\includegraphics[scale=.625]{RootsGd11q1.pdf} \hfill
\includegraphics[scale=.625]{RootsGd12q1.pdf}
\end{minipage}
\hfill
\begin{minipage}{0.25\linewidth}
\small
\begin{tabular}{rr|rrrr}
$d$ & $n$ & $N_1$ & $N_2$ & $N_3$ \\
& & {\tiny($\times$)} & {\tiny($\blacklozenge$)} & {\tiny($+$)} \\
\hline
1 & 1 & 0 & 0 & 1 \\
2 & 3 & 1 & 1 & 1 \\
3 & 5 & 2 & 2 & 1 \\
4 & 7 & 1 & 3 & 3 \\
5 & 9 & 2 & 4 & 3 \\
\rowcolor[gray]{.8}
6 & 11 & 3 & 3 & 3 \\
7 & 13 & 4 & 4 & 5 \\
8 & 15 & 5 & 5 & 5 \\
9 & 17 & 6 & 6 & 5 \\
10 & 19 & 5 & 7 & 7 \\
11 & 21 & 6 & 8 & 7 \\
\rowcolor[gray]{.8}
12 & 23 & 7 & 7 & 7 \\
$\vdots$ & $\vdots$ & $\vdots$ & $\vdots$ & $\vdots$ \\
\rowcolor[gray]{.8}
42 & 83 & 27 & 27 & 27
\end{tabular}
\end{minipage}
\end{center}
\captionlistentry[table]{. . . }
\captionsetup{labelformat=andtable}
\caption{\label{fig6} Counting the zeros of $G(d; z)$, $n = 2d - 1$. Two additional zeros appear when $6$ divides $d$, see shaded entries in table.}
\end{figure}
For this purpose it is convenient to define $\delta_{6 \mid d} {:=} 1$ if $6$ divides $d$ and $\delta_{6 \mid d} {:=} 0$ otherwise.
\begin{prop} \label{prop:Zeros}
If $d$ is even, $G(d;z)$ has exactly $4(\lfloor (d-1)/6\rfloor+\delta_{6 \mid d})+1$ zeros on the unit circle $\mathcal{C}_0$; all are simple and satisfy $\mathop{\mathrm{Re}} z \leq 1/2$. Their positions are determined by the solutions of the equation $g(\theta) = f(\theta)$ given in \eqref{ThetaEquation} below.
If $d$ is odd, there are no zeros of $G(d; z)$ on the unit circle $\mathcal{C}_0$.
\end{prop}
\begin{proof}
Since $z=1$ is not a zero of $G(d;z)$, the equation $G(d;z)=0$ is equivalent to
\begin{equation*}
z^{d-1}+1=\frac{z+1}{z-1}\frac{z^{d-1}}{(z-1)^{d-1}}.
\end{equation*}
Substituting $z=e^{i\theta}$ and changing to trigonometric functions we arrive at
\begin{equation} \label{ThetaEquation}
g(\theta) {:=} (-1)^{d/2} \cos\left( \frac{d-1}{2} \theta \right) = (-1)^{d/2} \frac{\cos(\theta/2)}{\left[2 i \sin(\theta/2) \right]^d} {=:} f(\theta).
\end{equation}
Suppose $d$ is even. Then $f(\theta) = \cos( \theta / 2 ) / [ 2 \sin( \theta / 2 ) ]^d$ is real-valued. Since the complex zeros of $G(d;z)$ occur in complex conjugate pairs, we may assume $0\leq \theta \leq \pi$. Using elementary calculus one shows that the function $f(\theta)$ is monotone decreasing and convex on $(0,\pi)$. Hence, $f(\theta)\geq f(\pi/3)+f'(\pi/3)(\theta-\pi/3)$ on $(0,\pi/3)$. This implies that $f(\theta)>1$ for $\theta\in (0, \pi/3-\alpha)$, where $\alpha=(4-2\sqrt{3})/(3d+1)$. On the other hand, on $[ \pi/3-\alpha,\pi/3)$ one can show that $|g(\theta)|<f(\theta)$. Hence, in \eqref{ThetaEquation} we have only to consider the range $[\pi/3,\pi]$ whereupon $f(\theta) \leq f(\pi/3) = \sqrt{3}/2 < 1$.
The function $g$ is a cosine function with period $4\pi/(d-1)$, where the sign factor ensures that $g$ has a positive derivative at $\theta = \pi$ (which is also an intersection point of $f$ and $g$). The other intersection points of $f$ and $g$ occur in the half-periods where $g \geq 0$. On such a half-period $I$ the convex function $f$ and the concave function $g$ (when restricted to $I$) can intersect in at most two points counting multiplicity, as can be seen by applying Rolle's theorem to the strictly convex function $f-g$.
That there are at least two such points on $I$ can be seen from the intermediate value theorem applied to the same function ($f-g>0$ at the endpoints of $I$ and $f-g<0$ at the midpoint of $I$) as illustrated in Figure~\ref{fig:typical.cases} for the three canonical cases.
\begin{figure}[h!t]
\begin{center}
\includegraphics[scale=.55]{solutiond12.pdf}
\includegraphics[scale=.55]{solutiond14.pdf}
\includegraphics[scale=.55]{solutiond16.pdf}
\end{center}
\caption{\label{fig:typical.cases} Graphs of $f(\theta)$ and $g(\theta)$; typical cases when solving equation~\eqref{ThetaEquation}.}
\end{figure}
Since there are $\lfloor (d-1)/6\rfloor$ full periods of $g$ in $[\pi/3,\pi]$ plus one more period (partially contained in $[\pi/3,\pi]$) whenever $6$ divides $d$, we have $2(\lfloor (d-1)/6\rfloor+\delta_{6 \mid d})$ zeros in the upper half-plane and that many conjugate zeros of $G(d;z)$ in the lower half-plane.
For odd $d$, equation~\eqref{ThetaEquation} has no real solution in $(0,\pi]$. Since $1$ is not a zero of $G(d; z)$, there are no zeros of $G(d; z)$ on $\mathcal{C}_0$.
\end{proof}
\begin{thm}
All zeros of $G(d; z)$ are simple for each $d \geq 1$.
\end{thm}
\begin{proof}
The cases $d=1$ and $d=2$ are obvious, so assume $d\geq 3$. Let
$\zeta$ be a zero of $G(d; z)$. Then $\zeta$ is simple if
$G^\prime(d; \zeta) \neq 0$. By means of some helpful substitutions for the expressions in braces\footnote{The replacements are $(\zeta-1)^d(\zeta^{d-1}+1) \mapsto \zeta^d + \zeta^{d-1}$ , $[(\zeta-1)^d - \zeta - 1] \zeta^{d-1} \mapsto -(\zeta-1)^{d}$ and $\zeta^d - ( \zeta - 1 )^d \mapsto \zeta^{d-1} [ ( \zeta - 1 )^d - 1]$.}, we find that
\begin{align*}
\zeta \left( \zeta - 1 \right) G^\prime(d; \zeta)
&= d \zeta \left\{ \left( \zeta - 1 \right)^d \left( \zeta^{d-1}
+ 1 \right) \right\} - \zeta^d \left( \zeta - 1 \right) +
\left( d - 1 \right) \left\{ \left[ \left( \zeta - 1 \right)^d -
\zeta - 1 \right] \zeta^{d-1} \right\} \left( \zeta - 1 \right) \notag \\
&= \left( d - 1 \right) \left[ \zeta^{d+1} - \left( \zeta -
1 \right)^{d+1} \right] + \left( d + 1 \right) \zeta^d \\
&= \left( d - 1 \right) \zeta \left\{ \zeta^d - \left( \zeta - 1 \right)^d \right\} + \left( d - 1 \right) \left( \zeta - 1 \right)^d + \left( d + 1 \right) \zeta^d \\
&= \left( d - 1 \right) \left( \zeta - 1 \right)^d \left( \zeta^d + 1 \right) + 2 \zeta^d {=:} P(d; \zeta).
\end{align*}
Suppose to the contrary that $P(d;\zeta) = 0$. Then on replacing $(\zeta-1)^d$ by $-2 \zeta^d /[(d-1)(\zeta^d+1)]$ in the formula \eqref{polynomial.G} for $G(d;\zeta)$ we get
\begin{equation} \label{eq:G.EQ.Q}
- \left( \zeta^d + 1 \right) G(d; \zeta) = \zeta^{d-1} \left( \zeta^{d+1} + \frac{d+1}{d-1} \zeta^d + \frac{d+1}{d-1} \zeta +1 \right)
\end{equation}
The polynomial $Q(d; z)$ of degree $d+1$ obtained by replacing $\zeta$ by $z$ in the second parenthetical expression in \eqref{eq:G.EQ.Q} has three real zeros (at $-1$) and $d-2$ complex zeros for even $d$ and two negative zeros ('near' $-1$) and $d-1$ complex zeros for odd $d$.\footnote{Interestingly, if $\frac{d+1}{d-1}$ is changed to $\frac{d-3}{d-1}$ in $Q(d; z)$, then the new polynomial has all its zeros on the unit circle.} This can be seen from the facts that
\begin{equation*}
Q(d; -1) = - \frac{2\left[ 1 - (-1)^d \right]}{d-1}, \qquad Q^\prime(d; -1) = \left[ 1 - (-1)^d \right] \frac{d+1}{d-1}, \qquad Q^{\prime\prime}(d; -1) = 0
\end{equation*}
and Descartes' Rule of Signs. Substituting $w = e^{i \phi}$ and using trigonometric functions, we arrive at
\begin{equation*}
Q(d; e^{i \phi}) = \frac{4 e^{i [(d+1)/2] \phi}}{d-1} \cos \frac{\phi}{2} \cos \frac{d \phi}{2} \left( d + \tan \frac{\phi}{2} \tan \frac{d \phi}{2} \right).
\end{equation*}
By symmetry, the number of solutions in the open set $(0,2\pi) \setminus \{ \pi \}$ of the equation
\begin{equation} \label{TheotherEquation}
\tan \frac{\phi}{2} \tan \frac{d \phi}{2} = - d
\end{equation}
equals the number of zeros of $\tan(d \phi/2)$ in $(0, 2 \pi)$ (cf. Figure~\ref{fig:simplicity}), which is $d-2$ for even $d$ and $d-1$ for odd $d$.
\begin{figure}[h!t]
\begin{center}
\includegraphics[scale=.75]{simplicityd6.pdf} \hspace{10mm}
\includegraphics[scale=.75]{simplicityd7.pdf}
\end{center}
\caption{\label{fig:simplicity} Typical cases when solving equation~\eqref{TheotherEquation}.}
\end{figure}
Thus, all zeros of $Q(d; z)$ are accounted for. They are either negative or complex conjugate pairs of zeros located on the unit circle.
For odd $d$, no zero $\zeta$ of $G(d; z)$ is on the unit circle (Proposition~\ref{prop:Zeros}) or negative (Proposition~\ref{prop:real.zeros}), so $Q(d; \zeta) \neq 0$, which contradicts \eqref{eq:G.EQ.Q}. Suppose $d$ is even. If $\zeta$ is on the unit circle $\mathcal{C}_0$, then it is simple (Proposition~\ref{prop:Zeros}). If $\zeta$ is not on $\mathcal{C}_0$, then $Q(d; \zeta) \neq 0$, which again contradicts \eqref{eq:G.EQ.Q}.
\end{proof}
\subsection{Asymptotics of Gonchar polynomials}
\label{sec:asymptotics}
Numerically computing the zeros for $G(d;z)$ for small values of $d$, we observe (cf. Figures~\ref{fig6} and \ref{fig:limitdistr} and Table~1) that they essentially form three groups separated by the sets
\begin{subequations}
\begin{align}
A_1 &{:=} \left\{ z \in \mathbb{C} : \mathop{\mathrm{Re}} z < 1/2, | z - 1 | > 1 \right\}, \\
A_2 &{:=} \left\{ z \in \mathbb{C} : | z | < 1, | z - 1 | < 1 \right\}, \\
A_3 &{:=} \left\{ z \in \mathbb{C} : \mathop{\mathrm{Re}} z > 1/2, | z | > 1
\right\}.
\end{align}
\end{subequations}
For the purpose of asymptotic analysis (large $d$) we rewrite the
equation $G(d; z) = 0$ in three different ways to emphasize an
exponentially decaying right-hand side when considering zeros
of $G(d;z)$ from the indicated part of the complex plane:
\begin{subequations} \label{eq:asympt.equations}
\begin{align}
z^{d-1} + 1 &= \frac{z+1}{z-1} \left( \frac{z}{z-1} \right)^{d-1},
& ( \mathop{\mathrm{Re}} z &< 1 / 2 ) \label{eq:approx1} \\
\left( z - 1 \right)^d - \left( z + 1 \right) z^{d-1} &= -
\left( z - 1 \right) \left[ \left( z - 1 \right) z \right]^{d-1},
& ( \left| z - 1 \right| \left| z \right| &< 1 ) \label{eq:approx2} \\
\left( z - 1 \right)^d - z - 1 &= - \left( z - 1 \right) \left( \frac{z - 1}{z} \right)^{d-1}. & ( \mathop{\mathrm{Re}} z &>
1 / 2 ) \label{eq:approx3}
\end{align}
\end{subequations}
The following theorem concerning the limit behavior of the zeros of $G(d;z)$ as $d \to \infty$ is illustrated in Figure~\ref{fig:limitdistr}.
\begin{thm}
Let $\Gamma$ be the set consisting of the boundary of the union of the two unit disks centered at $0$ and $1$ and the line-segment connecting the intersection points as indicated in Figure~\ref{fig:limitdistr}. Then, as $d \to \infty$, all the zeros of $G(d;z)$ tend to $\Gamma$, and every point on $\Gamma$ attracts zeros of these polynomials.
\end{thm}
\begin{figure}[h!t]
\centerline{\includegraphics[scale=1]{limitdistr.pdf}}
\caption{\label{fig:limitdistr} Zeros of $G(d;z)$ for $d = 1, 2, \dots, 40$.}
\end{figure}
\begin{proof}
First, we observe that a closed set $K$ in $\mathbb{C} \setminus \Gamma$ is free of zeros of $G(d;z)$ for sufficiently large $d$ as can be seen from the following relations obtained from \eqref{eq:approx1}:
\begin{align*}
\lim_{d \to \infty} \left| 1 + \frac{1}{z^{d-1}} - \frac{z+1}{z-1} \left( \frac{1}{z-1} \right)^{d-1} \right|^{1/d} &=
\begin{cases}
1 & \text{if $| z | > 1$, $| z - 1 | > 1$,} \\
1 / \left| z - 1 \right| & \text{if $| z | > 1$, $| z - 1 | < 1$,}
\end{cases} \\
\lim_{d \to \infty} \left| \frac{z+1}{z-1} - \left( \frac{z - 1}{z} \right)^{d-1} \left( z^{d-1} + 1 \right) \right|^{1/d} &=
\begin{cases}
1 & \text{if $| z | < 1$ , $| \frac{z-1}{z} | < 1$ ($\mathop{\mathrm{Re}} z > 1/2$),} \\
\left| \frac{z-1}{z} \right| & \text{if $| z | < 1$ , $| \frac{z-1}{z} | > 1$ ($\mathop{\mathrm{Re}} z < 1/2$),}
\end{cases} \\
\lim_{d \to \infty} \left| z^{d-1} + 1 - \frac{z+1}{z-1} \left( \frac{z}{z-1} \right)^{d-1} \right|^{1/d} &=
\begin{cases}
\left| z \right| & \text{if $| z | > 1$, $| \frac{z}{z-1} | < 1$ ($\mathop{\mathrm{Re}} z < 1/2$),} \\
1 & \text{if $| z | < 1$, $| \frac{z}{z-1} | < 1$ ($\mathop{\mathrm{Re}} z < 1/2$).}
\end{cases}
\end{align*}
The second part of the assertion, that every point of $\Gamma$ attracts zeros, is proved by contradiction.
Given a supposedly non-attracting point $w$ on $\Gamma$, there is a sufficiently small\footnote{The disk is small enough that its intersection with $\Gamma$ is contained either in $\mathcal{C}_0$, $\mathcal{C}_1$ or the open line-segment connecting the intersection points of $\mathcal{C}_0$ and $\mathcal{C}_1$.} open disk $D_w$ centered at $w$ containing no zeros of $G(d;z)$ for all sufficiently large $d$. It is possible to then define a single-valued analytic branch of the $d$-th root of any of the rational functions whose moduli appear on the left-hand sides above. Thereby, we obtain sequences of functions which are analytic and uniformly bounded in $D_w$. Such sequences form normal families in $D_w$. According to the right-hand sides above, at least one limit function of these families (which is necessarily analytic in $D_w$) will have the property that its modulus is $1$ in one part and is non-constant in the other part of $D_w$ which is separated by $\Gamma$. This gives the desired contradiction, since an analytic function in a domain that has constant modulus on a subdomain must be constant throughout the whole domain. Consequently, each point of $\Gamma$ attracts zeros of $G(d;z)$ as $d \to \infty$.
\end{proof}
It is inviting to compare the zeros of $G(d;z)$ with the ones of the polynomials given at the left-hand sides of \eqref{eq:approx1}, \eqref{eq:approx2} and \eqref{eq:approx3}. Such comparisons will likely lead to a finer analysis of the properties of the zeros of $G(d;z)$.
We conclude this note with some challenging conjectures.
\begin{conj} \label{conj:unit.circle}
For every positive integer $d$, the zeros of $G(d; z)$ form three
groups separated by the sets $A_1$, $A_2$ and $A_3$ except when $6$
divides $d$ in which case one also has the zeros $( 1 \pm \mathrm{i}
\sqrt{3} ) / 2$.
\end{conj}
When counting the zeros in the sets $A_1$, $A_2$ and $A_3$ a very regular pattern emerges, which can be seen from Table~1. In fact, inspection of this table shows that the values of column $N_1$ (increased by $2$ when $6$ divides $d$) agree with the number of zeros on $\mathcal{C}_0$ obtained in Proposition~\ref{prop:Zeros}. Assuming that this is true for all $d \geq 1$, by self-reprocity of $G(d;z)$, it would follow that $N_1 = N_2 = N_3 = 4k - 1$ if $d = 6 k$. {\em We expect that the zero counting scheme indicated in the Table~1 generalizes to all $d \geq 1$.}
Numerically, the zeros in $A_1$ and $A_3$ can be found near the respective unit circle $\mathcal{C}_0$ and $\mathcal{C}_1$.
\begin{conj}
If $d$ is even, all the zeros of $G(d;z)$ in $A_1$ are the zeros on the unit circle $\mathcal{C}_0$ given in Proposition~\ref{prop:Zeros}. If $d$ is odd, the zeros of $G(d;z)$ in $A_1$ alternately lie inside and outside $\mathcal{C}_0$. Furthermore, the zeros of $G(d;z)$ in $A_3$ are always outside of $\mathcal{C}_1$.
\end{conj}
\begin{conj} \label{conj:convex.left}
The zeros of $G(d; z)$ in $A_2$ are located on a curve which is convex from the left.
\end{conj}
{\bf Acknowledgment.} The first author is grateful to Don Zagier and
Wadim Zudilin for inspiring this work when attending the workshop
``Geometry and Arithmetic around Hypergeometric Functions'' from
September 28th --- October 4th, 2008, which was made possible by the
Mathematisches Forschungsinstitut Oberwolfach (MFO) and the
Oberwolfach-Leibniz-Fellow Programme (OWFL).
|
\section{#1}\setcounter{equation}{0}}
\def\thesection{\arabic{section}}
\def\theequation{\arabic{equation}}
\def\simgt{\mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}}}
\def\simlt{\mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}}}
\newcommand{\s}{\mbox{$\sigma$}}
\newcommand{\bi}[1]{\bibitem{#1}}
\newcommand{\fr}[2]{\frac{#1}{#2}}
\newcommand{\gm}{\mbox{$\gamma_{\mu}$}}
\newcommand{\gn}{\mbox{$\gamma_{\nu}$}}
\newcommand{\Le}{\mbox{$\fr{1+\gamma_5}{2}$}}
\newcommand{\R}{\mbox{$\fr{1-\gamma_5}{2}$}}
\newcommand{\GD}{\mbox{$\tilde{G}$}}
\newcommand{\gf}{\mbox{$\gamma_{5}$}}
\newcommand{\nub}{\mbox{$\nu_b$}}
\newcommand{\tb}{\tan\beta}
\newcommand{\Ima}{\mbox{Im}}
\newcommand{\Rea}{\mbox{Re}}
\newcommand{\Tr}{\mbox{Tr}}
\newcommand{\psl}{\slash{\!\!\!p}}
\newcommand{\cp}{\;\;\slash{\!\!\!\!\!\!\rm CP}}
\newcommand{\qq}{\langle \ov{q}q\rangle}
\newcommand{\uGu}{\bar{u}g_s(G\si) u}
\newcommand{\dGd}{\bar{d}g_s(G\si) d}
\newcommand{\nc}{\newcommand}
\newcommand{\uu}{\bar{u}u}
\newcommand{\dd}{\bar{d}d}
\nc{\gone}{\bar g_{\pi NN}^{(1)}}
\nc{\gzero}{\bar g_{\pi NN}^{(0)}}
\nc{\al}{\alpha}
\nc{\ga}{\gamma}
\nc{\de}{\delta}
\nc{\ep}{\epsilon}
\nc{\ze}{\zeta}
\nc{\et}{\eta}
\nc{\ka}{\kappa}
\nc{\rh}{\rho}
\nc{\si}{\sigma}
\nc{\ta}{\tau}
\nc{\up}{\upsilon}
\nc{\ph}{\phi}
\nc{\ch}{\chi}
\nc{\ps}{\psi}
\nc{\om}{\omega}
\nc{\Ga}{\Gamma}
\nc{\De}{\Delta}
\nc{\La}{\Lambda}
\nc{\Si}{\Sigma}
\nc{\Up}{\Upsilon}
\nc{\Ph}{\Phi}
\nc{\Ps}{\Psi}
\nc{\Om}{\Omega}
\nc{\ptl}{\partial}
\nc{\del}{\nabla}
\nc{\ov}{\overline}
\nc{\newcaption}[1]{\centerline{\parbox{15cm}{\caption{#1}}}}
\nc{\us}{U(1)$_S$}
\nc{\ub}{U(1)$_B$}
\nc{\co}{CoGeNT}
\nc{\ctw}{$^{12}$C}
\nc{\cth}{$^{13}$C}
\nc{\ctwm}{{\rm ^{12}C}}
\nc{\cthm}{{\rm ^{13}C}}
\nc{\neff}{${\cal N}_{\rm eff}$}
\nc{\neffm}{{\cal N}_{\rm eff}}
\nc{\bore}{$^{8}$B}
\def\beq{\begin{equation}}
\def\eeq{\end{equation}}
\def\bmat{\begin{displaymath}}
\def\emat{\end{displaymath}}
\def\bear{\begin{eqnarray}}
\def\eear{\end{eqnarray}}
\def\ba{\begin{eqnarray}}
\def\ea{\end{eqnarray}}
\def\bery{\begin{array}}
\def\ery{\end{array}}
\def\bit{\begin{itemize}}
\def\eit{\end{itemize}}
\def\ben{\begin{enumerate}}
\def\een{\end{enumerate}}
\def\btab{\begin{tabular}}
\def\etab{\end{tabular}}
\def\btbl{\begin{table}}
\def\etbl{\end{table}}
\def\bfig{\begin{figure}[htb]}
\def\efig{\end{figure}}
\def\bpic{\begin{picture}}
\def\epic{\end{picture}}
\def\st{\scriptstyle}
\def\ss{\scriptscriptstyle}
\def\hsx{\hspace{0.06in}}
\def\hse{\hspace{0.08in}}
\def\hst{\hspace{0.12in}}
\def\nnl{\nonumber \\}
\def\nl{\nonumber \\ &&}
\def\hocm{\hspace{1cm}}
\def\htcm{\hspace{2cm}}
\def\ga{\mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}}}
\def\la{\mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}}}
\def\gappeq{\mathrel{\rlap {\raise.5ex\hbox{$>$}}
{\lower.5ex\hbox{$\sim$}}}}
\def\lappeq{\mathrel{\rlap{\raise.5ex\hbox{$<$}}
{\lower.5ex\hbox{$\sim$}}}}
\def\ohsq{\Omega_{\widetilde\chi}\, h^2}
\def\gyr{{\rm \, G\kern-0.125em yr}}
\def\mev{{\rm \, Me\kern-0.125em V}}
\def\gev{{\rm \, Ge\kern-0.125em V}}
\def\tev{{\rm \, Te\kern-0.125em V}}
\def\cp{C\!P}
\def\tsq{|{\cal T}|^2}
\def\halft{{\textstyle{1\over2}}}
\def\slash#1{\rlap{\hbox{$\mskip 1 mu /$}}#1
\def\tbt{\tan \beta}
\def\ttbt{\tan^2 \beta}
\def\hc{{\rm h.c.}}
\def\emunu{\eta^{\hspace{0.01in} \mu \hspace{0.01in} \nu}}
\def\bfp{{\bf p}}
\def\nhat{{\bf \hat{n}}}
\def\half{\frac{1}{2}}
\def\athird{\frac{1}{3}}
\def\aforth{\frac{1}{4}}
\def\Tr{\rm Tr}
\def\Ker{\rm Ker}
\def\index{\rm index}
\def\bmtheta{\mbox{\boldmath $\theta$}}
\def\bmphi{\mbox{\boldmath $\phi$}}
\def\bmalpha{\mbox{\boldmath $\alpha$}}
\def\bmsigma{\mbox{\boldmath $\sigma$}}
\def\bmgamma{\mbox{\boldmath $\gamma$}}
\def\bmomega{\mbox{\boldmath $\omega$}}
\newcommand{\FRAME}[1]{\fbox{\mbox{$#1$}}}
\begin{document}
\begin{titlepage}
\setcounter{page}{1}
\vspace*{0.2in}
\begin{center}
\hspace*{-0.6cm}\parbox{17.5cm}{\Large \bf \begin{center}
Neutrino Physics with Dark Matter Experiments and \\
the Signature of New Baryonic Neutral Currents
\end{center}}
\vspace*{0.5cm}
\normalsize
\vspace*{0.5cm}
\normalsize
{\bf Maxim Pospelov$^{\,(a,b)}$}
\smallskip
\medskip
$^{\,(a)}${\it Perimeter Institute for Theoretical Physics, Waterloo,
ON, N2J 2W9, Canada}
$^{\,(b)}${\it Department of Physics and Astronomy, University of Victoria, \\
Victoria, BC, V8P 1A1 Canada}
\smallskip
\end{center}
\vskip0.2in
\centerline{\large\bf Abstract}
New neutrino states \nub, sterile under the Standard Model interactions,
can be coupled to baryons via the isoscalar vector currents that are much stronger than
the Standard Model weak interactions.
If some fraction of solar neutrinos oscillate into \nub\ on their way to Earth,
the coherently enhanced elastic \nub-nucleus
scattering can generate a strong signal in the dark matter detectors.
For the interaction strength a few hundred times stronger than the weak force,
the elastic \nub-nucleus scattering via new
baryonic currents may account for the existing anomalies in the
direct detection dark matter experiments at low recoil. We point out that for
solar neutrino energies the baryon-current-induced inelastic scattering is suppressed,
so that the possible enhancement of new force is not in
conflict with signals at dedicated neutrino detectors.
We check this explicitly by calculating the \nub-induced deuteron breakup,
and the excitation of 4.4 MeV $\gamma$-line in $^{12}$C. Stronger-than-weak
force coupled to baryonic current implies the existence of new abelian
gauge group \ub\ with a relatively light gauge boson.
\vfil
\leftline{March 2011}
\end{titlepage}
\subsection*{1. Introduction}
Standard Model (SM) of particles and fields must be augmented
to include neutrino mass physics and perhaps extended
even further to account for the "missing mass" of the Universe, or cold dark matter (DM).
During the last decade the underground experiments
\cite{cdms,xe,DAMA,cogent,cresst,zep,edel,picasso} aimed at direct detection
of DM in the form of weakly interacting massive particles (WIMPs) \cite{WIMPS}
have made significant inroads into
the WIMP-nucleon cross section vs WIMP mass parameter space. Since no DM-induced nuclear recoil signal was found
with the exception of two hints to be discussed below,
they constrained many models of dark matter and ruled out some portion of the
parameter space in the best motivated cases such as {\em e.g.} supersymmetric neutralino
DM \cite{witten}, Higgs-portal singlet DM \cite{HiggsDM} etc.
It has been argued by some authors that although primarily designed to
search for WIMPs, these experiments are in fact multi-purpose devices that
can be also used for alternative signatures of other effects beyond Standard
Model. In particular, using same instruments one can look for the
absorption of keV-scale bosonic super-WIMPs \cite{sWIMPs},
search for the axion emission from the Sun \cite{aSun},
and also investigate some additional signatures of WIMP-atom scattering that exist in
"non-minimal" WIMP models \cite{nonMin}. This paper extends this point
further and opens a new direction: we show that neutrino physics beyond SM can also
be probed with the dark matter experiments.
Elastic scattering of neutrinos on nuclei \cite{Stodolsky} is enhanced by the coherence factor $N^2$,
where $N$ is the number of neutrons.
Straightforward analysis \cite{FM,nuDM} of the SM solar neutrino elastic scattering rates on nuclei
used in DM experiments reveal
several basic points:\\
\begin{itemize}
\item Despite the coherent enhancement, the scattering rates are way too small,
leading to counting rates not exceeding
$10^{-3}~{\rm kg}^{-1}{\rm day}^{-1}{\rm keV}^{-1}$. Such low rates do not introduce
any $\nu$-background to WIMP searches at the current levels of sensitivity.
\item The nuclear recoil spectrum is usually very soft, $E_r\sim (E_\nu)^2/M_{\rm Nucl} ~
\sim {\rm few~KeV}$ or less.
\item Solar boron ($^8$B) neutrinos are the best candidates for producing an observable signal,
because of the compromise between the relatively large flux and the energy spectrum extending to 15 MeV.
\end{itemize}
Of course, at this point the DM experiments typically target much harder recoil
and are far away from low counting rates induced by solar neutrinos. On the other hand,
the last generation of dedicated solar neutrino experiments \cite{SuperK,SNO,Borexino} have been
extremely successful in
detecting solar neutrinos via charged current reactions (CC), elastic scattering on electrons (ES),
and $Z$-exchange mediated (NC) deuteron break-up \cite{SNO}.
It is the combination of all these three signals that led to a very credible resolution of
the long-standing solar-neutrino deficit problem via the neutrino oscillation and the
MSW mechanism \cite{osc,MSW}.
However, it is easy to imagine that three active SM species $\nu_e,~\nu_\mu,~\nu_\tau$ with
their (almost completely) established mass/mixing parameters may not be the last word in the
neutrino story. In this paper we consider a model of "quasi-sterile" neutrino $\nu_b$ that
has no charged currents with normal matter, no ES on electrons or other leptons, but has much
enhanced NC with baryons (NCB).
We shall consider the strength of new NCB interaction to be much larger than Fermi constant,
$G_B \sim (10^2-10^3)\times G_F$. Such interactions can be mediated by new vector
bosons of the gauged baryon number \ub, and for that reason we call this new hypothetical neutrino
state as the "baryonic" neutrino \nub. If any of the solar SM neutrino flavors
oscillate into $\nu_b$ within one astronomical unit, then the current DM experiments will
in principle be able to pick it up via the coherently enhanced NCB signal. Whether such strong NCB would lead to a
measurable energy deposition in the standard neutrino experiments requires special investigation and
is addressed in this paper. We find that although the inelastic NCB scattering is enhanced by
a huge factor $G_B^2/G_F^2$, it is also suppressed by a tiny factor $E_\nu^4 R_N^4$,
where $R_N$ is a nuclear radius-related parameter. The resulting rate for NCB processes in
neutrino detectors is then can be made comparable or smaller than the regular neutrino counting rates.
It is somewhat tempting to relate the proposed $\nu_b$ model with the recently reported
anomalies/signals in the direct DM detection. For a long time of course the DAMA and its successor
DAMA/LIBRA experiments have been claiming \cite{DAMA} the annual modulation of the energy deposition
in NaI crystals with the maximum in early June and minimum in December, which is consistent with the
expected seasonal modulation of the WIMP-nucleus scattering rate. Then, last year the CoGeNT collaboration \cite{cogent}
has reported
unexpected (in the null hypothesis) rise of their signal at recoil energies below $E_r=$1 keVee. Given the mass of
Ge nuclei, and typcial quenching factors in germanium,
it is plausible that the rise of \co\ signal at low KeVee can be produced
by \nub\ resulting from oscillations of boron neutrinos, $\nu_{SM} \to \nu_b$,
and hypothesized enhancement of NCB can compensate for small neutrino scattering rates.
It is also clear that mimicking DAMA signal with $\nu_b$ is also possible in a more restricted sense.
Of course the usual seasonal modulation of the neutrino flux due to the eccentricity
of the Earth orbit will have a minimum in the early July and a maximum in early January.
However, the neutrino oscillation phenomenon is not monotonic in distance \cite{Kamland},
and if the oscillation length for $\nu_{SM} \to \nu_b$ is comparable to the Sun-Earth distance, the annual modulation phase of the
\nub\ scattering signal can be reversed by $\pi$. We investigate this
opportunity, and conclude that both \co\ and DAMA signal can be described with \nub-type models
(provided that DAMA data can tolerate a $\sim 1$ month phase shift).
We further argue that if indeed this is the case, the model is very predictive, and
there will be further ample opportunites for probing $\nu_b$ both at DM and neutrino detectors,
as well as at more conventional particle physics experiments.
This paper is organized as follows: in the next section we introduce the class of \nub\ models
and specify a parameter range that is the most perspective for the DM experiments. Section 3
addresses NCB elastic and inelastic scattering, including calculations of
\nub-induced activation of carbon and the deuteron breakup reactions.
Section 4 studies the possibility of phase reversal in the annual modulation
signal. Our conclusions are reached in Section 5.
\subsection*{2. Baryonic neutrino and baryonic neutral currents}
The basic features of the model are as follows: we introduce a new gauge group
\ub, and give all quark fields of the SM, $Q,~U, D$, the same charge under
\ub, which we call $g_b/3$. We also introduce a new left-handed neutrino species $\nu_b$
that has a charge $g_l$ under this new group and no charge under any of the SM gauge groups.
In the interest of anomaly cancellations it is also
desirable to introduce a right-handed partner of $\nu_b$ with the same charge. Then the new group couples
to the "vector-like" matter multiplets, and although SM+\ub\ will in general be anomalous, the anomalies
can be cancelled at some heavy scales. Variants of this model may include some partial
gauging of the SM lepton species under \ub. Neither the right-handed $\nu_b$ nor the details
of anomaly cancellation will be important for this paper.
Furthermore, we assume that \ub\ is spontaneously broken by the \ub-Higgs vacuum expectation value $\langle \phi_b \rangle$,
and exactly how this happens will not be of direct consequence for us either.
The relevant gauge part of the Lagrangian is then given by
\be
\label{start}
{\cal L} = - \fr14V_{\mu\nu}^2 + \fr12m_V^2V_\mu^2 + \bar \nu_{b} \gamma_\mu(i\partial_\mu + g_lV_\mu) ~\nu_{b}
+\sum_{q} \bar q (iD_{SM}
\!\!\!\!\!\!\!\!\!\!\!/ \;\;\;\;\;
+\fr{1}{3}g_b \gamma_\mu V_\mu) q + {\cal L}_m.
\ee
First two terms in (\ref{start}) are the standard Maxwell-Proca terms for $V_\mu$, the sum extends over all
quark types and flavors, and $D_{SM}$ is the SM covariant derivative that includes
gauge interactions appropriate for each quark species $q$. The mass part of the
Lagrangian ${\cal L}_m$ besides the usual SM mass terms should also account for
neutrino masses, and generate mixing to a new state \nub. In this paper we are not going to consider vector
exchange with virtualities beyond $O(10~{\rm MeV})$, and therefore it is convenient to
switch from quarks to nucleons,
\be
\label{isosc}
\fr{1}{3}V_\mu g_b~\sum_{q} \bar q \gamma_\mu q ~\to~ g_b V_\mu(\bar p \gamma_\mu p + \bar n \gamma_\mu n) + ...
\ee
Ellipses stands for $O(m_N^{-1})$ terms associated with the $V_{\mu\nu} \bar N \sigma_{\mu\nu}N$
part of the form factor, which will be small for any process we consider in this paper.
The coupling of $V_\mu$ to the isoscalar vector current of nucleons
$J_\mu^{(0)}=\bar p \gamma_\mu p + \bar n \gamma_\mu n$ will have important implications
for both the elastic and inelastic scattering of \nub\ on nuclei. The exchange by the \ub\ gauge boson
creates the NCB Lagrangian,
\be
\label{exchange}
{\cal L}_{NCB} = \bar \nu_b \gamma_\mu \nu_b ~\fr{g_lg_b}{m_V^2 + \Box} ~J_\mu^{(0)},
\ee
that in the limit of $m_V^2\gg Q^2$ is just a new contact dimension 6 operator with the
effective coupling constant $G_B$:
\be
{\cal L}_{NCB} = G_B\times \bar \nu_b \gamma_\mu \nu_bJ_\mu^{(0)};~~
G_B = \fr{g_l g_b}{m_V^2} \equiv {\cal N} \times \fr{10^{-5}}{\rm GeV^2}.
\ee
Here we have introduced an "enhancement" parameter ${\cal N}$ that quantifies how much stronger
$G_B$ is compared to the weak-scale value of $10^{-5}{\rm GeV}^{-2}$. We note that stronger-than-weak
interactions among four neutrino species were considered earlier in {\em e.g.} Ref. \cite{Khlopov0}.
The use of baryonic force as a mediator between SM and dark matter was considered recently in \cite{Ko}.
One may wonder if ${\cal N}$ as large as a 100 or a 1000 can be consistent with low energy data
on meson decays, such as $K\to \pi \bar \nu_b \nu_b$. It turns out that due to the conservation
of the baryon current, the loop amplitude for the underlying $s\to d \bar \nu_b \nu_b$ decay is
additionally suppressed by $G_F Q^2$, which compensates for all possible enhancements due to ${\cal N}$.
(In contrast, the quark axial vector analogue of (\ref{start}) will be strongly constrained to
have ${\cal N} \la 1$.) Thus, from quark flavor perspective, the baryonic portal (\ref{start}) is
one of the two "safe" portals (the other being the kinetic mixing with hypercharge \cite{Holdom})
that allow attaching stonger-than-weak interactions to the quark currents. We shall not pursue
the meson decay constraints on the model any further in this paper, and turn our attention
to the neutrino mass sector.
The most natural way of having a UV-complete theory of neutrino masses is via the introduction
of right-handed neutrinos states $N_R$. We can use the same singlet right-handed neutrinos
coupled to the Higgs-lepton bilinears $LH$ and Higgs$_b$--neutrino \nub\ bilinears $\nu_b \phi_b$
in a gauge-invariant way,
\be
{\cal L}_m = LH {\bf Y} N + \nu_{bL}\phi {\bf b} N +(h.c.) + \fr{1}{2} N^T{\bf M}_RN.
\ee
Here ${\bf M}_R$ and ${\bf Y}$ are the familiar 3$\times$3 right-handed neutrino mass matrix and Yukawa matrix,
while ${\bf b} $ is the new Yukawa vector parametrizing the couplings of the left-handed part of $\nu_b$ to $N$.
Integrating out $N$ states results in the low-energy 4$\times$4 neutrino mass and
mixing matrices, $M_{ij}$, where $i,j$ run over $e,\mu,\tau,b$ flavors.
While of course a full four-state analysis can be done, we shall simplify
our discussion by the following assumptions:
\begin{enumerate}
\item The entries of $3\times 3$ submatrix $M_{\rm active,~active}$ will in general be somewhat larger than
$M_{{\rm active},b}$ and $M_{b,b}$ components so that the mixing pattern can be addressed
sequentially: first the mixing of the SM neutrinos and then the admixture of the $\nu_b$.
\item A tri-bimaximal ansatz will be taken for the 3$\times$3 mixing of the SM neutrino species
for simplification, although having $\theta_{13}=0$ is not crucial.
\item We shall assume a preferential mixing of
$\nu_b$ to $\nu_2$, with the relevant parameters that we call $\Delta m_b^2$ and $\theta_b$,
so that true mass eignestates are $\nu_I = \cos\theta_b \nu_2 + \sin\theta_b\nu_b$;
$\nu_{II} = -\sin\theta_b \nu_2 + \cos\theta_b\nu_b$.
\item The sign of $G_B$ will be chosen to ensure that the matter effects for $\nu_b$
will not lead to the matter-induced $\nu_{\rm active} \to \nu_b$ transitions.
\end{enumerate}
The combination of these assumptions forms the following (simplified) picture of neutrino oscillations:
inside the Sun the neutrino oscillations occur largely between $\nu_e$ and
$\nu_+ \equiv (\nu_\mu+\nu_\tau)/\sqrt{2}$,
\be
\left(\begin{array}{c}
\nu_e\\\nu_+
\end{array}
\right) \simeq
\left(\begin{array}{cc}
\sqrt{\fr{2}{3}}& \sqrt{\fr{1}{3}}\\-\sqrt{\fr{1}{3}}&\sqrt{\fr{2}{3}}
\end{array}
\right)
\left(\begin{array}{c}
\nu_1\\\nu_2
\end{array}
\right)
\ee
while the "$-$" combination and $\nu_b$ stay unexcited. We would need only the higher-end of the Boron neutrino
spectrum where MSW effect dominates.
Upon the neutrino exit from the dense region of the Sun, it represents
an almost pure $\nu_2$ state, $ \nu_2 = \sqrt{\fr{1}{3}}\nu_e + \sqrt{\fr{2}{3}}\nu_+$, with individual
flavor probabilities
\be
P_e({\rm Sun}) \simeq \fr{1}{3};~~ P_+({\rm Sun}) \simeq \fr{2}{3};~~ P_b({\rm Sun})=0.
\ee
Then vacuum oscillations start building a non-zero probability for \nub\ due to
the $\nu_I$ and $\nu_{II}$ being the true mass eigenstates in vacuum. Upon the arrival to the Earth, the following
energy-dependent probabilities will approximate the neutrino flavor composition:
\begin{eqnarray}
P_b({\rm Earth}) \simeq \sin^2(2\theta_b)\sin^2\left[\fr{ \Delta m^2_b L(t)}{4E }\right] \nonumber\\
\label{P's}
P_e({\rm Earth}) \simeq \fr{1}{3}\left( 1 - \sin^2(2\theta_b) \sin^2\left[\fr{ \Delta m^2_b L(t)}{4E }\right] \right)\\
P_+({\rm Earth}) \simeq \fr{2}{3}\left( 1 - \sin^2(2\theta_b)\sin^2\left[\fr{ \Delta m^2_b L(t)}{4E }\right] \right), \nonumber
\end{eqnarray}
where $L(t)$ is the Earth-Sun distance with a slight eccentricity modulation,
\be
L(t) \simeq L_0\left(1-\epsilon \cos\left[
\fr{2\pi(t-t_0)}{T} \right]\right); ~~ L_0 = 1.5\times 10^8{\rm km}; ~~ \epsilon \simeq 0.167;~~ t_0 \simeq 3~{\rm Jan} .
\ee
The most interesting range for $\Delta m^2_b$ to consider is
\be
\label{hier}
10^{-10} ~{\rm eV}^2 \la \Delta m^2_b ~\ll ~ \Delta m^2_{\rm Solar,~atm}.
\ee
A scale of $O(10^{-10}) ~{\rm eV}^2 $ is the so-called "just so" mass splitting that may introduce significant
changes to the otherwise very predictable $\propto L^{-2}$ seasonal
variations of the \nub\ flux at Earth's location. With $\Delta m^2_b$ being much smaller than
$10^{-5}~{\rm eV}^2$ there is no danger of distorting KamLAND results \cite{Kamland} even for a
relatively large angle $\theta_b$, although the matter effects for $\bar\nu_b$ could be significant.
We also find it convenient to introduce the energy parameter $E_0$ directly related to the mass splitting,
\be
E_0 = \fr{\Delta m_b^2 L_0}{4\pi} = 6.05~{\rm MeV} \times \fr{\Delta m_b^2}{10^{-10}~{\rm eV}^2},
\ee
that defines last zero of $P_b$ as the function of energy, $P_b(L=L_0,E=E_0) = 0$.
Since in all NCB rates $P_b$ will enter in the combination with $G_B^2$, it is
also convenient to define
\be
{\cal N}_{\rm eff}^2 = {\cal N}^2 \times \fr{1}{2}\times \sin^2(2\theta_b),
\ee
so that in the limit of large $E_0$ the oscillations average out and
$P_bG_B^2 \to {\cal N}_{\rm eff}^2 \times 10^{-10}~{\rm GeV}^{-4}$.
The pattern of masses and mixing considered here is not the most natural:
we assume a pair of very degenerate $\nu_{I}$ and $\nu_{II}$ mass eigenstates
replacing $\nu_2$ and \nub. Given that the mass of $\nu_2$, regardless of the hierarchy pattern,
is always in between $\nu_1$ and $\nu_3$, this will require some specific adjustments of
the full 4$\times$4 mass matrix. The search for more natural realizations of
$\nu_{SM} \to \nu_b$ oscillations with long oscillation length, including matter effects for
a different sign of $G_B$, goes beyond the
scope of this paper. The goal of the next two sections will be to find the sensitivity to \neff\
in various processes involving the elastic and inelastic scattering of \nub.
We would like to close this section with some model-building comments.
A very intriguing question to ask is whether SM neutrinos would tolerate new large
NCB. A conventional answer is "no", as the so-called non-standard neutrino
interactions (NSI) with quarks and charged leptons were addressed in a number of papers \cite{NSI}
and almost no room at $O(1)G_F$ level was found, let alone the much enhanced NCBs hypothesized in this paper.
However, NSI studies \cite{NSI} with rare exceptions \cite{Nelson} assume that the scale of the
mediation is comparable to the weak scale, and ignore the possibility of light
vector bosons communicating between neutrinos and baryons. As a matter of counterexample one could consider
a model with two new gauge groups, \ub\ and the other being a quantized lepton flavor,
{\em e.g.} $L_\mu$ or $L_\tau$. The connection between two new vector sectors is given by the
kinetic mixing term $\eta V^{(1)}_{\mu\nu} V^{(2)}_{\mu\nu}$. Then additional effective interaction of
a SM neutrino with the baryonic current is given by
\be
{\cal L}_{\rm eff} \propto
\bar \nu_{SM} \gamma_\alpha \nu_{SM} ~\fr{\eta g_lg_b \Box}{(m_{V1}^2 + \Box)(m_{V2}^2 + \Box)} ~J_\alpha^{(0)}
\label{howdoyoulikeit}
\ee
Such interaction gives no contribution to the forward scattering amplitude and thus is not affecting neutrino
oscillation, and it is $1/Q^2$ suppressed in the large $Q^2$ regime, avoiding strong constraints from
deep-inelastic neutrino scattering. It is then clear that the
choice of $m_{V1}, ~ m_{V2}$ in the MeV range may allow having (\ref{howdoyoulikeit}) at
$Q^2 \sim (1-10)~ {\rm MeV}^2$ to be considerably stronger than the SM weak force\footnote{The author would like to acknowledge
very stimulating discussions with B. Batell and I. Yavin on the possibility of the NSI enhancement.}.
The interactions of type (\ref{howdoyoulikeit}) can lead to the
detectable recoil signal from elastic scattering
of solar SM neutrinos, along the same lines as the \nub-scattering idea advocated in this paper.
A possibility of modified SM neutrino interactions such as (\ref{howdoyoulikeit}) can be
very effectively tested
using the proposed neutrino-nucleus elastic scattering detectors placed near the intense
source of stopped pions \cite{CLEAR}.
\subsection*{3. Elastic and inelastic scattering of \nub}
Elastic scattering of $\nu_b$ on nuclei will create a recoil signal regulated by the
strength of NCB, and the probability of oscillation (\ref{P's}). It can be picked up by the
direct dark matter detection experiments with low recoil thresholds.
Also, \nub\ neutrinos can deposit a significant amount of energy
on the order of a few MeV by activating excited nuclear states or via extra neutrons
created by nuclear breakup.
The main finding of this section can be summarized as follows: the ratio of the
elastic to inelastic cross sections in the interesting neutrino energy range $E_\nu \la 15$ MeV,
is governed by the following relation:
\be
\fr{\sigma_{\nu_b-{\rm Nucl}}({\rm elastic})}{\sigma_{\nu_b-{\rm Nucl}}({\rm inelastic})}\sim \fr{A^2}{E_\nu^4R_N^4} \sim 10^8,
\label{ratio}
\ee
where we took $A\sim 100$, $R_N^{-1} \sim 100 $ MeV, and $E_\nu \sim 10$ MeV.
It is this huge ratio that makes small-scale experiments such as \cite{cogent} competitive in
sensitivity to \nub\ with the large-scale neutrino detectors.
\subsubsection*{3.1. Elastic scattering}
The differential cross section for the NCB elastic scattering of left-handed \nub\ on a nucleus of mass
$M_N$ with $A$ nucleons is given by
\be
\fr{d\sigma_{\rm el}}{d(\cos\theta)}= \fr{E^2A^2g_b^2g_l^2(1+\cos\theta)}{4\pi(M_V^2 +{\bf q}^2)^2}
\simeq \fr{1}{4\pi}\times G_B^2E^2A^2(1+\cos\theta),
\label{elastics}
\ee
where the elastic scattering momentum transfer is $q = ({\bf q}^2)^{-1/2} = 2 E\sin(\theta/2)$ and cannot
exceed twice the neutrino energy $E$. In the second relation we took $M_V \gg E$, which allows to shrink vector propagator.
Using relations between the neutrino scattering angle $\theta $, nuclear recoil energy $E_r$,
and the minimum neutrino energy required to produce $E_r$-recoiling nucleus,
\be
E_r = \fr{E^2}{m_N}\times(1-\cos \theta);~ E^{\rm min} = \sqrt{\fr{E_r M_N}{2}},
\ee
we rewrite the elastic cross section (\ref{elastics}) in the following form
\be
\fr{d\sigma_{\rm el}}{dE_r}= \fr{1}{2\pi}\times G_B^2A^2m_N\left(1-\fr{(E^{\rm min})^2}{E^2}\right).
\label{NCBel}
\ee
One can readily see that the NCB cross section (\ref{NCBel}) is related to the SM elastic neutrino-nucleus
cross section \cite{Stodolsky} by
$ G_B^2 A^2 \to G_F^2 (N/2)^2$ substitution, where $N$ is the number of neutrons
(with small corrections in $1-4\sin^2\theta_W$ parameter). For the momentum transfers and nuclei considered
in this paper the form factor corrections are $<5\%$ percent and are ignored.
Using cross section (\ref{NCBel}), the flux and the energy distribution
of \bore\ neutrinos \cite{Bahcall} ({\em hep} solar neutrinos provide a small correction),
we derive the counting rates as the function of interaction strength and the oscillation probability.
For a moment, we neglect small seasonal modulations and take the limit of $\epsilon\to 0$ .
For the medium composed only of atoms
with atomic number $A$, we approximate these rates by:
\begin{eqnarray}
\fr{dR}{d E_r} \simeq \fr{A^2m_N}{2\pi}\times \fr{1}{2}\sin^2(2\theta_b)G_B^2\Phi_{\rm ^8B}\times I(E_r,E_0)\nonumber\\
\simeq 85 ~\fr{\rm recoils}{\rm day\times kg\times KeV} \times
\left( \fr{A}{70} \right)^3\times \fr{\neffm^2}{10^4}\times I(E_r,E_0).
\label{dR}
\end{eqnarray}
The input (total flavor) flux of \bore\ neutrinos is taken to be $\Phi_{^8{\rm B} } = 5.7\times 10^6{\rm cm}^{-2}{\rm s}^{-1}$
and $m_N \propto Am_p$.
The recoil integral $I(E_r,E_0)$ in eq. (\ref{dR}) is given by the convolution of the \bore\ energy
distribution, the energy-dependent part of oscillation probability, and the kinematic factor in the
cross section reflecting neutrino helicity conservation:
\be
I(E_r,E_0) = \int_{E^{\rm min}(E_r)}^\infty dE ~
\left(1-\fr{(E^{\rm min})^2}{E^2}\right) \times f_{\rm ^8B}(E) \times 2\sin^2\left[\fr{ \pi E_0}{E }\right].
\label{IEr}
\ee
Here the distribution function is normalized as $\int_{{\rm all} ~ E} f_{^8{\rm B}}(E)dE = 1$. For the limit of large $E_0$ (fast oscillations), the last multiplier in (\ref{IEr}) becomes 1. If a detector threshold
corresponds to recoil energies that require $E^{\rm min}$ to be above the end-point of \bore\ neutrino spectrum,
$I\equiv 0$ (apart from small corrections from {\em hep} and diffuse supenova neutrinos). It is the case for most of
the existing WIMP detectors, but not for all of them.
In real detectors registering ionization such as \cite{DAMA,cogent}
it is the electron equivalent of the
energy release rather than the recoil energy that is detected. We take the relation between the
two by following recent DM-related analyses \cite{Itay,Collar}
\begin{eqnarray}
{\rm Ge:}~~~E_r({\rm keVee}) \simeq 0.2\times (E_r({\rm keV}))^{1.12}\nonumber\\
{\rm Na ~in ~NaI:} ~~~ E_r({\rm keVee}) \approx 0.33\times E_r({\rm keV}).
\label{quenching}
\end{eqnarray}
The second relation is far less precise than the first one \cite{Collar}.
The counting rates in germanium resulting from scattering of \nub\ created by the oscillations of
\bore\ and {\em hep} solar neutrinos are presented in Fig. 1.
We have taken three cases of mass splitting:
$E_0 \gg E_\nu^{\rm max}$, and $E_0 = 12, ~14$ MeV.
The NCB rates are plotted for the value of $\neffm^2 = 10^4$.
For this enhancement factor, the resulting counting rates are clearly within
reach of current generation of low-threshold germanium detectors ({\em i.e.} \co).
\begin{figure}
\includegraphics[width=0.9\textwidth]{Ge1}
\caption{Expected recoil event rate in Germanium in units of recoils/day/kg/keVee as the function of
$E_r$ in keVee. The NCB enhancement factor, $\neffm=100$. A, B and C lines correspond to
$E_0=\infty$, $E_0=12$ MeV and $E_0 = 14$ MeV. }
\label{Ge}
\end{figure}
Inspection of Figure 1 shows that choice of different mass splitting that makes the oscillation
length comparable to 1 A.U. influences
the shape of the spectrum. This is because the most important part of the
spectrum for the recoil in excess of 0.5 keVee is above neutrino energies of 10 MeV, where the \bore\ neutrino
spectrum is already sharply falling. The neutrino oscillations with $E_0$ close to 12 MeV will lead to the
suppression of higher $E_r$ and to the steep rise of the signal at lower $E_r$. The sharp
end of the neutrino spectrum prevents other Ge experiments with higher threshold like CDMS \cite{cdms} to
probe the NCB scattering in the regime of large recoil where \cite{cdms} has strong sensitivity.
The signal from the recoil due to \nub\ neutrinos is very similar in morphology to that of sub-10 GeV scale WIMPs.
This is because a typical momentum transfer in a heavy nucleus - light WIMP collision is
$q \sim m_{\rm wimp} v \sim 10$ MeV, which is about the same for \bore\ neutrino scattering.
There is one kinematic difference though: the back-scattering of WIMPs that produces hardest
recoil is kinematically allowed, while for neutrinos it is forbidden by helicity
conservation. This additionally limits the capabilities of high-threshold experiments to detect
\nub\ neutrinos in comparison with light WIMPs.
Is it possible to use \nub\ as another speculative explanation of \co\ results \cite{cogent}?
The overall event rate can indeed be reproduced well with $\neffm \sim 10^2 - {\rm few}\times 10^2$,
depending on $E_0$. For the large $E_0$ parameter, the enhancement factor of $\neffm = 10^2$
seems sufficient: it gives 7 recoils/day/keVee at $E_r = 0.7$ keVee, which is about the same as
the experimental data suggest after accounting for the efficiency \cite{cogent,Collar}.
The shape of the predicted signal is also similar to the counting rate profiles observed by
CDMS at the Stanford Underground facility \cite{SUF}.
Fitting the exact spectral shape of excess events
at \co\ falls outside the scope of our current investigation. We should also note that the
expected total counting rate for the material used in CRESST detectors \cite{cresst}
due to the neutrino-oxygen scattering is given by
\be
R_{\rm O ~in~ CaWO_4}(E_r> 10~{\rm keV}) \simeq 0.2\times \fr{\rm recoils}{\rm day\times kg} \times \fr{\neffm^2}{10^4},
\ee
which is well within their detection capabilities for $\neffm \sim 100$.
Other methods in development that use liqud helium as a detecting medium with
a potentially very low energy threshold \cite{he} also look promising for detecting \nub-induced recoil.
It is also important that the choice of very low-mass target such as $^4$He will allow discriminating
between $\ga 5$ GeV WIMPs and \nub's: the effective recoil energy goes down
at $M_N<M_{\rm WIMP}$, while it becomes larger for \nub\ scattering.
\subsubsection*{3.2. Inelastic scattering}
Unlike light WIMPs that can cary significant momentum but very little energy, \nub\ can easily lead to
an MeV-scale energy deposition.
Here we turn our attention to the NCB inelastic processes and will adress the following issues:
the NCB deuteron breakup, and the NCB excitation of the first $2^+$ resonance in $^{12}$C resulting in 4.4 MeV $\gamma$ line:
\begin{eqnarray}
\label{breakup}
d+\nu_b &\to& \nu_b+n+p\\
^{12}{\rm C} +\nu_b &\to& \nu_b +^{12}{\rm C}^*(4.44~{\rm MeV}) \to \nu_b + ^{12}{\rm C}+ \gamma
\label{excitation}
\end{eqnarray}
The main scientific question to answer is whether the enhanced values of $G_B^2P_b$ can be consistent with the
constraints provided by SNO on "extra neutrons" from (\ref{breakup}) and by Borexino and other liquid scintillator detectors
on "extra gammas" from (\ref{excitation}). There are of course other processes that one has to consider in a more
comprehensive study, including the excitation of oxygen, the breakup of $^{13}$C to $^{12}$C$+n$ etc, but they will all
follow the scaling in Eq. (\ref{ratio}). The earlier studies of the nuclear excitations due to the
differenty type of neutrino couplings can be found in \cite{Khlopov1}. There are also elastic channels of energy deposition
via $\nu_b+p\to \nu_b +p$ \cite{JohnPetr}, but the proton recoil from the scattering of \bore\ neutrinos
would fall below the detector thresholds.
To understand the origin of ratio (\ref{ratio}) one does not have to perform any sophisticated calculations.
We consider the scattering of $\sim 10$ MeV energy neutrinos, so that their wavelengths are much larger
than the characteristic nuclear size of a few fm. Therefore, one can safely expand the nuclear matrix elements
in series in $q$, or in neutrino energy $E$, as $q$ is bounded by $E$. Here is how the inelastic matrix element of the
$\mu=0$ component of the isoscalar vector current $J_\mu^{(0)}$
between the deuteron bound state
and $np$ continuum will look like in this expansion:
\begin{eqnarray}
\label{expansion}
\langle d| \exp(i{\bf q r}^{(n)}) + \exp(i{\bf q r}^{(p)}) |np\rangle
\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\\
= 2\langle d| np\rangle + i {\bf q}\cdot \langle d| {\bf r}^{(n)} + {\bf r}^{(p)} |np\rangle
-\fr{q_kq_l}{2}\langle d| r_k^{(n)}r_l^{(n)} + r_k^{(p)}r_l^{(p)}|np\rangle
= -\fr{q_kq_l}{4}\langle d| r_kr_l|np\rangle,
\nonumber
\end{eqnarray}
where ${\bf r}^{(n)} ,~ {\bf r}^{(p)} $ are the position operators for the neutron and the proton.
The zeroth and first order terms in $q = |{\bf q}|$ are trivially zero due to the orthogonality of
the wave functions (${\bf r}^{(n)} + {\bf r}^{(p)} $ is the center-of-mass operator and cannot
mediate inelastic transitions). In the last line we have introduced
the relative position vector ${\bf r} = {\bf r}^{(n)} - {\bf r}^{(p)} $,
and the quadratic in $r$ operator can be further separated into the isotropic "charge-radius"
and quadrupole components. For the $0^+ \to 2^+$ transition in \ctw\ only the quadrupole part will matter.
It is of course instructive to revisit the SM deuteron breakup \cite{Butler}, and observe that
isoscalar vector component of the standard weak current gives a very minor contribution to the total
cross section at low $E_\nu$ due to this $q^2$ suppression of the amplitude. The SM rate is of course dominated by the
isovector axial-vector current that corresponds to the difference of nucleon spins ${\bf s}^{(n)} - {\bf s}^{(p)}$,
an operator that has non-zero inelastic matrix elements even in the $O(q^0)$ order.
We perform the calculation of the NCB-induced deuteron breakup using the "zero-radius" approximation
of the initial and final state wave functions,
\begin{eqnarray}
\psi_{\rm in}({\bf r}) = \fr{\sqrt{2\kappa}}{\sqrt{4\pi }r}\exp(-\kappa r); ~~
\psi_{\rm f}({\bf r}) = \exp(i{\bf pr})\\
\kappa = \sqrt{2E_b \mu} = 45 ~{\rm MeV} ; ~~ p^2={\bf p}^2 = 2\mu(E - E_f - E_b).
\end{eqnarray}
Here $E$ and $E_f$ are the initial and final energy of \nub, $E_b =2.2$ MeV is the absolute value of the deuteron binding energy,
and $\mu \simeq (m_n+m_p)/4$ is the reduced mass of the proton-neutron two-body system. Relative momentum $p$
of the final state is fully determined from the neutrino energies, as the recoil of the deuteron center-of-mass is
negligible. Parameter $\kappa$ is the familiar "bound state momentum", related to the inverse size of the
deuteron, $\kappa \sim R_d^{-1}$, and its relative smallness reflects large spatial extent of the deuteron.
In a slightly excessive for the problem at hand language our calculations correspond to the leading order
of the pionless effective field theory \cite{Butler,Pionless}. They can be systematically improved if needed,
or treated with the more elaborate nuclear physics tools (see {\em e.g.} \cite{Kubodera}). None of this
will of course change the order of $q$ in which the effect first occurs.
Straightforward calculations give the differential over the final neutrino energy cross section:
\begin{eqnarray}
\frac{d\sigma_{d\to np}}{dE_f}= \frac{G_B^2 E_f^2 m_p}{8\pi^2}
\fr{\kappa^5 p}{( p^2 + \kappa^2)^6}\left[ E^2E_f^2 +
\fr{12 p^4}{5\kappa^4}(E^4-\fr{2}{3}E^3E_f +\fr{10}{9}E^2E_f^2 -\fr{2}{3}EE_f^3 +E_f^4) \right]
\label{dtonp}
\end{eqnarray}
The result for $d\sigma_{d\to np}$
shows a $O(E^4\kappa^{-4})$ suppression in agreement with $O(q^2)$ of the deuteron matrix element
and in agreement with (\ref{ratio}). Judging by the size of the subleading corrections in the SM calculations \cite{Butler}
we expect this answer to hold within $\sim 20\%$ accuracy.
\begin{figure}
\centering
\includegraphics[width=0.85\textwidth]{dnpsigma1}
\vspace{0.5 cm}
\hspace{-0.5cm}\includegraphics[width=0.85\textwidth]{fsigma1}
\caption{Top pannel: Deuteron breakup cross section for the SM NC processes (top curve) and for the
NCB \nub-neutrinos (bottom curve). The NCB cross section is plotted for ${\cal N}^2$ enhancement factor of $10^4$.
Bottom pannel: same cross sections convoluted with \bore\ energy distribution and the energy dependent part of
the oscillation probability. The top curve is the SM NC distribution of effective cross section,
and A, B, C are the same for the NCB with $E_0=\infty, ~12, ~14$
MeV and $\neffm^2=10^4$. The areas under the curves give the proportion of neutrons produced via SM NC and NCB processes.}
\label{dnp}
\end{figure}
The final integral over $E_f$ in the interval from $0$ to $E-E_b$
gives the total NCB deuteron breakup cross section. In Figure 2, upper pannel, we plot $\sigma_{d\to np}$
for the Standard Model neutrino and for \nub\ with the choice of enhancement factor ${\cal N}^2=10^4$.
As expected, the NCB cross section has a faster rise with neutrino energy due to quadrupolar nature of the NCB interaction.
In Figure 2, lower pannel, we also show the convolution of the cross section with the energy
distribution of boron neutrinos times the energy-dependent part of the oscillation probability,
$2\sin^2(\pi E_0/E )$. As in the previous subsection, the NCB rates are considered
for large $E_0$ and for $E_0 = 12,~14$ MeV, while the enhancement factor is kept at
$\neffm =100$. The areas under curves give the total effective cross sections,
and when multiplied by $\Phi_{^8{\rm B}}$, correspond to the breakup rate per deuteron.
For the three NCB cases considered here have the following comaprison to the SM rate:
\be
\fr{\sigma_{\rm NCB}}{\sigma_{\rm SM~NC}}\simeq \fr{\neffm^2}{10^4}\times (0.14,~0.06,~0.13)~~ ~{\rm at}~ ~E_0 = \infty;~ 12~{\rm MeV};~14~{\rm MeV}.
\ee
A 15\% increase in the neutron production rate at SNO can be tolerated, and if one chooses a
sizeable $\theta_b$ so that the active neutrino flux is slightly less than SM+SSM predict, the
total neutral current rate may not even change.
We conclude that SNO NC events leave enough room for the possible $\neffm^2 \sim O(10^4)$ (and slightly
higher) enhancement of the NCB rate.
We now address the $\ctwm\to\ctwm^*$ activation due to \nub. To avoid the complications
arising from nuclear physics we shall assume that both the ground state and the first excited state
of \ctw\ are given by $3\alpha$ configurations. This is a very well justified assumption, which leads to
a relation between the matrix elements of the baryonic current and electric current,
\be
\langle 0^+ | J_i^{(0)} | 2^+ \rangle = 2 \langle 0^+ | J_i^{\rm em} | 2^+ \rangle =
2 \fr{(E_{2^+} - E_{0^+}) q_j}{6} \langle 0^+ | Q_{ij}^{\rm em} | 2^+ \rangle
\label{relation}
\ee
Only the lowest order in $q$ terms are retained here, and the
$\mu=0$ component can be restored from gauge invariance. The factor of 2 in (\ref{relation})
comes from the fact that the baryonic charge of $\alpha$-particles is twice larger than its electric charge.
The information on the value of the transitional quadrupole moment $\langle Q_{ij}\rangle$
can be extracted from the $\ctwm^*$ decay width $\Gamma = 1.08\times 10^{-2}$ eV:
\be
\label{quadr}
\overline{|\langle Q_{ij}^{\rm em}\rangle|^2}= \fr{90\Gamma}{\alpha (\Delta E)^5} = ( 3.3 ~{\rm fm} )^4,
\ee
where the value of quadrupole moment squared is averaged over the
arbitrary projection of the $2^+$ angular momentum, and $\Delta E =
E_{2^+} - E_{0^+}=E-E_f = 4.439$ MeV. Note that we define
electric quadrupole and electric current {\em without} the "$e$",
and account for $\alpha$ explicitly in (\ref{quadr}). The total
inelastic cross section for the \nub-induced $\ctwm\to\ctwm ^*$ transition is given by
\be
\sigma_{^{12}{\rm C}\to ^{12}{\rm C}^*}=
\fr{8G_B^2E^5(E-\Delta E)\overline{|\langle Q_{ij}^{\rm em}\rangle|^2}}{81\pi}
\left[ 1-3x+\fr{39}{8}x^2-\fr{19}{4}x^3+\fr{39}{16}x^4-\fr{9}{16}x^5\right],
\ee
where $x = \Delta E/E$. The benchmark value for this cross section at $E=8$ MeV and ${\cal N}=1$
is $2.5 \times 10^{-48}~{\rm cm}^2$.
With this cross section the effective rate of injection of 4.4 MeV gamma quanta in pseudocumene
(scintillating material used by the Borexino experiment) is estimated to be
\be
R(4.4 ~{\rm MeV}) \sim (0.05-0.15)\times \fr{\gamma~{\rm injections}}{\rm 100~tons\times day} \times \fr{\neffm^2}{10^4}.
\label{borex}
\ee
This is not a large rate by any measure, but it is nevertherless comparable to the counting rates
in 3-5 MeV window from \bore\ ES processes and from the $^{208}$Tl background events \cite{BorexB}.
The actual counting rate should be obtained by applying to (\ref{borex}) the efficiency factor that the collaboration
can extract from their calibration data and simulations. At this point we can only conclude that
there must be some sensitivity to NCB at $\neffm \sim 10^2$ level at the large scale neutrino detectors that
use carbon-based scintillators. More definitive statements and perhaps stronger sensitivity to \nub\ can be
derived from dedicated analyses. Moreover, the search for the
extra $\gamma$-lines in a different energy range was already performed by the Borexino collaboration
in connection with hypothetical Pauli-forbidden \ctw\ decays \cite{PauliB}. The search for NCB would represent a far
less exotic physics cause in our opinion. One could also conduct similar searches of \nub-induced
excitation of $^{16}$O nuclei using SNO and SuperK data.
\subsection*{4. Annual modulation of \nub\ rates}
In this section we would like to address the question of seasonal modulation of the NCB rate.
The seasonal modulation of solar neutrino rate was observed by SNO and SuperK collaborations
\cite{SuperK,Snoseason}. It exhibits full agreement with expected $\propto L^{-2}(t)$, 3.3\% modulation of the
neutrino flux, with an appropriate minimum in the summer (Northern hemisphere).
The hypothetical NCB elastic scattering rate will have the same modulation pattern as long as
$\Delta m_b^2$ is large, or in other words at $E_0\gg E_{\nu~{\rm solar}}$.
In the opposite limit of $E_0\ll E_{\nu~{\rm solar}}$ the modulation effects are suppressed because
$\Phi_\nu \propto L_0^2L^{-2}(t)\sin^2[\pi E_0 L(t)(L_0 E)^{-1} ] \to (\pi E_0/E)^2$, which is time-independent.
However, it is easy to imagine that the flux of \nub\ neutrinos can have a more intricate seasonal modulation
pattern. For example, if $E_0$ is between the maximum of the \bore\ neutrino spectrum and its end-point,
the high-energy fraction of the distribution will have a higher flux in the summer. This is best illustrated
in Figure 3, where the expected flux of \nub\ resulting from oscillations of boron neutrinos is
convoluted with the time-dependent part of $P_b$,
$L_0^2L^{-2}(t)\sin^2[\pi E_0 L(t)(L_0 E)^{-1} ]$ at $E_0=12$ MeV. The two curves correspond to
$t=t_{\rm perihelion}$ ($\sim$ 3 Jan) and $t_{\rm aphelion}$ ($\sim$ 4 Jul). Although on average there is more
\nub\ neutrinos
arriving to the Earth in January, in the most relevant range of energies, $E > 10 $ MeV, the flux in July is
larger. Therefore, for this fraction of neutrinos there is a phase reversal, and the elastic scattering rates
will reflect that.
\begin{figure}
\centering
\includegraphics[width=1\textwidth]{Seasons1}
\caption{Boron \nub\ neutrino flux modified by the time-dependent part of the oscillation probability
$ 2L_0^2L^{-2}(t)\sin^2[\pi E_0 L(t)(L_0 E)^{-1} ]$ with $E_0=12$ MeV. The black curve is for July, and
the gray curve is for January. Although the total integral under the gray curve is bigger than under the black one, it is the
high-end of the spectrum that would determine rates at the existing DM detectors, where July
rates are larger. }
\label{Seasons}
\end{figure}
Next we calculate the expected seasonal modulation in the counting rate,
\be
\label{sinus}
\fr{dR_{\rm mod}}{dE_r} = \fr12\left( \left.\fr{dR}{dE_r}\right|_{\rm Jul} -\left.\fr{dR}{dE_r}\right|_{\rm Jan} \right),
\ee
for NaI detectors using the quenching factor from Eq. (\ref{quenching}).
We would like to remark in passing that for some ranges of neutrino energies
there can be a significant departure from a simple time-sinusoidal function,
but to observe such effects one would probably require very high statistics and very good energy
resolution. Modulation rates, $dR_{\rm mod}/dE_r$, as defined in Eq. (\ref{sinus}) are plotted in Figure 4
for the same three choices of $E_0$ and $\neffm=100$ as before. One can see that indeed modulation of both signs is
possible, and that the rate of modulated NCB signal at this \neff\ is
indeed probed by DAMA/LIBRA experiment \cite{DAMA}, which is sensitive to modulation amplitudes $O(10^{-2})$cpd/kg/keVee.
\begin{figure}
\centering
\includegraphics[width=0.90\textwidth]{Modulation1}
\caption{Modulation of the counting rate in recoils/kg(NaI)/keVee for \nub-scattering on
Na. As before, A curve is for large $E_0$, B is for $E_0 = 12$ MeV and
C is for $E_0=14$ MeV while $\neffm^2=10^4$. Both signs of modulation are possible. }
\label{Modulation}
\end{figure}
Is it possible that \nub-Na elastic scattering is behind the DAMA/LIBRA seasonal modulation anomaly?
The magnitude of the predicted modulation can be in a very good agreement with DAMA results \cite{DAMA}.
Moerover, as we saw in the previous section, $\neffm^2\sim 10^4$ is thus far consistent with other observations
and constraints (and with simultaneous explanation of
\co\ low $E_r$ anomaly). Of course the phase of the DAMA results will require $E_0$ to be in the right range.
But even then, would the early July maximum be consistent with DAMA/LIBRA claims of the oscillation phase?
The best fit point for the maximum is about 4-5 weeks different from the $t_{\rm aphelion}$ \cite{DAMA}.
It would be interesting to know if the early July maximum is actually excluded by DAMA data,
and the criticism expressed in Ref. \cite{phase} about the errors on the phase being too tight be
properly addressed by the collaboration.
Going away from \nub\ idea, one can also notice that many other exotic physics explanations of DAMA
signal can be invoked (if it allows to tolerate $\sim$ 1 month phase shift).
For example, the emission of solar axions with their subsequent absorption in DM detection experiments
can be a cause of low-energy ionization signal \cite{aSun}. On this picture, one can super-impose the oscillations of axions
into some "sterile axions" with the oscillation length similar to $L_0$ in order to break the monotonic
$L(t)$-dependence, and flip the phase of the modulation, achieving results similar to those of Figure 3 and 4.
\subsection*{5. Discussion and conclusions}
The oscillation of SM neutrinos to some new neutrino state on the way from the Sun to the Earth
is a realistic possibility.
We have shown that there exists the whole class of models where neutrino physics
beyond SM can be probed by the low-threshold DM detectors, which become equally sensitive
or even more sensitive to this type of neutrinos than the large scale neutrino detectors.
Such models require that new neutrino states \nub\ (or modified SM model neutrino interactions
in the spirit of Eq. (\ref{howdoyoulikeit}))
couple almost exclusively to the baryon current. The isoscalar vector properties of this current
lead to a very strong enhancement of the elastic over inelastic scattering, $\sigma_{\rm elastic}/\sigma_{\rm inelastic}
\sim 10^8$, providing an unexpected competitiveness factor to small scale experiments such as \co.
We have shown the effective strength of the NCB can be much larger
than the weak-scale value without being in conflict with any of the observational data.
We have also shown that the recent anomalies in direct DM detection, such as DAMA and \co,
can be explained by the $\nu_{\rm SM}\to \nu_b$ oscillation of \bore\ neutrinos
with subsequent scattering of \nub\ on Ge and Na nuclei. (This statement
relies on the assumption of $t_{\rm aphelion}$ being consistent with DAMA/LIBRA modulation phase.)
This may look counter-intuitive at first, but we have shown that the phase flip of seasonal modulations
is possible for the high-energy end of the \bore\ spectrum. This is a very speculative
explanation (and perhaps equally speculative as the WIMP recoil explanation), but it
is interesting enough to motivate further studies. In particular, we believe that the
Borexino collaboration can perform the search of the \nub-activated 4.4 MeV line in \ctw,
and probably surpass the sensitivity to the NCB enhancement factor $\neffm$ of $100$.
At the same time it seems apparent that further technological developments of
low-threshold WIMP/\nub\ detectors are required. Should the current low-energy anomalies in DM detectors
firm up to constitute a definitive signal of new-physics-induced recoil, some significant efforts and different mass targets
would be required to observationally distinguish between the low-mass WIMP and \nub\ signals.
Below, we would like to discuss further implications of the models involving new neutrino states with
enhanced baryonic currents.
\begin{itemize}
\item {\em Collider implications.} If the $G_B \ga 100 G_F$, and if the new
interaction is truly contact, the proton-antiproton collisons
will lead to strong new sources of missing energy signals in $\nu_b \bar\nu_b$ channel
and will most likely be excluded by the Tevatron experiments. This will not happen, however, in models
of relatively weakly coupled mediators with sub-GeV mass. Therefore the collider searches should be able
to place an {\em upper} bound on $m_V$.
\item {\em Fixed target implications.} A GeV-scale \ub\ baryonic vector, the carrier of
NCB interaction, can be produced in the
collisions of energetic proton beams with a target. Immediate decays of these vectors
will generate a flux of \nub\ state that can be searched for at near detectors via their
NCB interactions. This is very similar to the ideas of the "MeV-scale DM beams"
discussed previously in \cite{target}. There can be also implications for the
terrestrial anti-neutrino physics, as matter effects induced by $V$-exchange can be large
for \nub\ antineutrinos \cite{Nelson}. Enhanced neutral currents of \nub\ neutrinos may
help explaining the long-standing puzzle of the LSND anomaly \cite{LSND}, perhaps borrowing some elements of the
recent suggestion \cite{Gnin}. It also has to be said that over the last two years a lot of efforts have been
invested in systematically searching for the "kinetic mixing" (or hypercharge) portal (see {\em e.g.} \cite{ourguys}
and references therein).
Barionic portal is another example of a perfectly safe from the model-building perspective way of
introducing stronger-than-weak forces at low energy, and therefore it should be systematically
searched for using proton-on-target facilities. But perhaps the most NCB-search effective type of experiments
to perform with proton beams is the proposal \cite{CLEAR} of a neutrino-nucleus elastic scattering detector.
\item {\em Cosmological implications.} A new light neutrino state (and two neutrino states if right-handed
copy of \nub\ is also light) can be at borderline of what is allowed by early cosmology and observations
of light elemental abundances (most recent analysis can be found in \cite{bbn}). Is the model
with much enhanced baryonic currents has a chance to be consistent with these constraints? Actually, despite the interaction
strength of 100 or 1000 of $G_F$, \nub\ will decouple from thermal plasma {\em earlier} than the SM neutrinos.
That is because its thermalization rate will be proprtional to the baryon-to-photon ratio, which is a
small number $O(10^{-10})$. Therefore actual decoupling of \nub\ may happen with the decays and annihilations
of abundant hadronic species at temperatures of $\sim 100$ MeV, and therefore BBN bounds
from over-population of radiative degrees of freedom can be easily evaded.
\item {\em Astrophysical implications.} Another interesting aspect of \nub\ models is their NCB production
in stars. In the SN \nub\ will not provide new effective energy sinks because they would not escape
freely the explosion zone. However one should expect that a comparable to the SM number of \nub\ neutrinos is
created, so that one could detect them using the same DM/\nub\ detectors. Should a nearby SN explosion
happen, the existing neutrino scintillator detectors can pick up the \nub-NCB signal that would appear as a much enhanced
$\nu_\mu,\nu_\tau$-NC signal considered in \cite{Beacomtoday} (modulo the uncertainty in effective temperature for
\nub).
\item {\em Rare decay implications.} Relatively large NCB currents should open new channels
for the missing energy decays of $B$ and $K$ mesons. As argued in this paper, the conservation of
the baryonic current makes it a relatively safe portal compared to {\em e.g.} scalar or axial-vector portal.
Nevertheless, if the $K\to \pi V$ decay is kinematically allowed, it may lead to the underlying
two-body signature of $K\to \pi$ plus missing energy decays, making it an appealing target for the
next generation of precision kaon physics experiments.
\end{itemize}
\subsection*{Acknowledgements}
I would like to thank B. Batell, D. McKeen, J. Pradler and I. Yavin for helpful discussions.
I am also grateful to A. de Gouvea, R. Harnik and J. Kopp for pointing out inconsistencies
of the neutrino oscillation picture in version 1 of this work.
The work is supported in part by the Government of Canada through NSERC and by the Province of Ontario through MEDT.
|
\section*{Introduction}
\label{sec:intro}
Because of Darboux's theorem and because any open subset
of~$\mathbb{C}^n$ contains a Lagrangian torus, it is possible to construct
a Lagrangian torus in any symplectic manifold.
The construction of Lagrangian submanifolds can become far more
difficult as soon as we require some conditions - we will be here
interested in monotonicity - on the Lagrangian submanifold.
Recall that a Lagrangian submanifold $L$ of a symplectic
manifold~$(W,\omega)$ is said to be monotone (see Oh~\cite{Oh_I}) if
there exists a non-negative constant~$K_L$ such that for every disc
$u \in \pi_2(W,L)$,
$$\int u^\star \omega = K_L \; \mu_L (u),$$
where $\mu_L (u)$ is the Maslov class of the disc $u$.
Recall also that the existence of a monotone Lagrangian submanifold
implies that the ambient symplectic manifold itself must be monotone,
which means that there exists a non-negative constant $K_W$ such
that for every sphere $v~\in~\pi_2(W)$,
$$\int v^\star \omega = K_W \; c_1(v),$$
where $c_1$ is the first Chern class of $(W,\omega)$.
Note that because of the relation between the first Chern class and
the Maslov class of a sphere in~$\pi_2(W)$, in the case the first
Chern class of $W$ does not vanish identically on $\pi_2(W)$, we
have the following relation between the monotonicity constants:
$$K_W = 2 K_L.$$
In particular in this case, the monotonicity constant of any Lagrangian
submanifold is prescribed by the monotonicity constant of the ambient
symplectic manifold, and this gives already many restrictions on the
Lagrangian.
Even in the case of $\mathbb{C}^n$ endowed with its canonical symplectic
structure, not many constructions of monotone Lagrangian tori are
available.
The first and simplest example is the Clifford (or split) torus,
that is the
product of $n$~circles enclosing disks of the same area.
As any construction of a Lagrangian torus gives rise to infinitely many
tori with the same symplectic invariants by applying a Hamiltonian
diffeomorphism, we are rather interested in equivalence classes of tori
under the action of the group of Hamiltonian diffeomorphisms.
It was only in 1995 that Chekanov~(\cite{Chekanov96}) gave the first
examples of monotone Lagrangian tori in $\mathbb{C}^n$ which are not
Hamiltonian isotopic to the Clifford torus.
Such tori are said exotic.
The Clifford torus can be embedded in the complex projective space
and the product of spheres via the embedding of a ball or a polydisc of
suitable size in~$\mathbb{C} \mathbb{P}^n$ or in~$\times_n \mathbb{S}^2$. It is a monotone
Lagrangian torus in $\mathbb{C} \mathbb{P}^n$ or in~$\times_n \mathbb{S}^2$ called again
Clifford torus. The Clifford torus was also the only known example
of monotone Lagrangian torus in~$\mathbb{C} \mathbb{P}^n$ and in~$\times_n \mathbb{S}^2$
till Chekanov and Schlenk~(\cite{Chek-SchI},
\cite{Chek-Sch-MSRI}) described in~2006 their families of
exotic monotone Lagrangian tori.
Biran et Cornea (\cite{Bir-Cor-uniruling}) have also given recently a
construction of monotone Lagrangian tori thanks to the circle bundle
construction of Biran~(\cite{Biran-NonIntersections}).
It is believed that one should recover the families of monotone
Lagrangian tori of Chekanov and Schlenk with the constructions
through circle bundles. The present note proves
this fact for $\mathbb{C} \mathbb{P}^2$ and $\mathbb{S}^2 \times \mathbb{S}^2$ and we will deal with the higher
dimensions in some future work.
The tori of Chekanov and Schlenk in $\mathbb{C} \mathbb{P}^2$ and $\mathbb{S}^2 \times \mathbb{S}^2$ are defined by
modifying slightly the construction of the exotic monotone Lagrangian
torus in $\mathbb{C}^2$ of Eliashberg and Polterovich in~\cite{El-Pol-1993}:
given a ``suitable'' (see Section~\ref{sec:CS-torus}
and~\ref{sec:CS-torus-insphere}) curve
$\gamma : [0, 2 \pi] \longrightarrow \mathbb{C}$,
consider
$$\left\{{ \left. \left({\gamma(s) e^{i\theta}, \gamma(s) e^{-i\theta}
}\right) \right| \theta \in [0, 2 \pi], s \in [0, 2 \pi] } \right\}$$
and then embed it into $\mathbb{C} \mathbb{P}^2$ or $\mathbb{S}^2 \times \mathbb{S}^2$.
The monotone torus in $\mathbb{C} \mathbb{P}^2$ (or analogously in~$\mathbb{S}^2 \times \mathbb{S}^2$) defined by
Biran and Cornea in~\cite{Bir-Cor-uniruling} is not coming from a
torus in $\mathbb{C}^2$ and an embedding of a ball or a polydisc: one
starts with a symplectic sphere $\Sigma$ (coming from a polarisation
of $\mathbb{C} \mathbb{P}^2$ or $\mathbb{S}^2 \times \mathbb{S}^2$, see~\cite{Biran-Barriers}), and constructs
the torus as the restriction to an equator of the sphere $\Sigma$
of a circle subbundle of a disc bundle~$E_{\Sigma}$ over $\Sigma$.
In the case of $\mathbb{S}^2 \times \mathbb{S}^2$, this construction is known to be equal to the
first example in the litterature of exotic monotone Lagrangian torus
given by Entov and Polterovich in~\cite{En-Pol-Rigid}.
In this note we prove the following:
\begin{thmintro}
The construction of monotone exotic torus of Chekanov and Schlenk and
the one given by the circle bundle construction of Biran are
Hamiltonian isotopic in $\mathbb{C} \mathbb{P}^2$ and $\mathbb{S}^2 \times \mathbb{S}^2$.
\end{thmintro}
The idea of the proof for the projective plane is to relate
Biran and Cornea's torus to the
first construction of exotic monotone torus in $\mathbb{C}^2$ given by
Chekanov in~\cite{Chekanov96}.
This construction in $\mathbb{C}^2$ is known to be Hamiltonian isotopic to
the monotone exotic torus of Eliashberg and Polterovich
in~\cite{El-Pol-1993}.
A similar isotopy between this tori can be used in the case of~$\mathbb{C} \mathbb{P}^2$
to define an isotopy between Chekanov and Schlenk's torus and a
modified Chekanov's torus.
After embedding into~$\mathbb{C} \mathbb{P}^2$, one can prove that this modified
Chekanov's torus is a circle subbundle of the disc bundle~$E_{\Sigma}$
over an equator of $\Sigma$ and then isotope this torus inside the
disc bundle to the torus of Biran and Cornea.
The strategy for~$\mathbb{S}^2 \times \mathbb{S}^2$ is analogous.
The article is organised the following way: in the first section we
recall the two constructions of exotic monotone Lagrangian torus in
$\mathbb{C}^2$, namely the first one by Chekanov~\cite{Chekanov96} and the one
of Eliashberg and Polterovich~\cite{El-Pol-1993}, and we exhibit a
Hamiltonian isotopy between the two.
In the second section, we focus on the case of~$\mathbb{C} \mathbb{P}^2$, recall the
constructions of Chekanov and Schlenk~(\cite{Chek-SchI},
\cite{Chek-Sch-MSRI}), and Biran and Cornea~(\cite{Bir-Cor-uniruling}),
and prove the existence of the Hamiltonian isotopy in this case. The
third section deals with the case of~$\mathbb{S}^2 \times \mathbb{S}^2$
\begin{acknowledgement}
\emph{This work follows from a question raised by Felix Schlenk. I wish
to thank him, Paul Biran, Octav Cornea and Leonid Polterovich for their
explanations and remarks on these constructions. I also want to thank
Kai Cieliebak, Ivan Smith and Maksim Maydanskyi for helpful
discussions.}
\end{acknowledgement}
\noindent
This work was carried out while the author held a postdoctoral
fellowship attached to the SNF project "Complexity and recurrence
in Hamiltonian systems" at the University of Neuch\^atel, a
Mathematical Sciences Research Institute postdoctoral fellowship,
and a postdoctoral fellowship at the University of Cambridge
funded by the European Research Council
grant ERC-2007-StG-205349.
The author wishes to thank these three institutions for their
hospitality and stimulating research atmosphere.
\section{The first constructions in $\mathbb{C}^2$}
\label{sec:C2}
In this section, we recall the first constructions of monotone
exotic tori in~$\mathbb{C}^2$ as they will be useful to understand the
constructions in the compact symplectic manifolds $\mathbb{C} \mathbb{P}^2$ and
$\mathbb{S}^2 \times \mathbb{S}^2$.
In order to simplify the computations and to avoid too many constants,
we will use normalisations of the standard symplectic forms
on~$T^\ast M$ for $M = \mathbb{R}^{n}$ or $M = \mathbb{S}^1$ (the circle $\mathbb{S}^1$
being identified with $\mathbb{R} / 2 \pi \mathbb{Z}$) such that the Liouville
$1$-form on $T^\ast M$ is:
\begin{equation}
\label{normalisation}
\lambda = \dfrac{1}{\pi} \Sigma p_i d q_i,
\end{equation}
where $(p,q)$ are the usual local coordinates on cotangent bundles,
$q=(q_1, \ldots, q_n)$ coordinates on the basis and
$p=(p_1, \ldots, p_n)$ coordinates in the fibers.
For example, with such a normalisation, the integral of the
Liouville form along the circle centered at the origin and of
radius~$r$ in $T^\ast \mathbb{R} \simeq \mathbb{R}^2$ is $r^2$.
\subsection{The first description by Chekanov}
Chekanov has given a first description of exotic monotone tori in
$\mathbb{R}^{2n}$ in~\cite{Chekanov96}. We recall here his construction.
For any Lagrangian submanifold $L$ in $\mathbb{R}^{2n}$, he has defined a
Lagrangian submanifold $\Theta (L)$ in $\mathbb{R}^{2n+2}$
the following way.
Consider the embedding
$$
\begin{array}{cccc}
i_n: & \mathbb{S}^1 \times \mathbb{R}^n & \longrightarrow & \mathbb{R}^{n+1} \\
& (\theta,x_1,\ldots,x_n) & \longmapsto & (e^{x_1}
\cos(\theta),e^{x_1} \sin(\theta),x_2,\ldots,x_n).
\end{array}
$$
Then $I_n = (i_n^\ast)^{-1}$ is a symplectic embedding of
$T^\ast (\mathbb{S}^1 \times \mathbb{R}^n)$ into $T^\ast \mathbb{R}^{n+1}$.
Denote by $\Theta (L)$ the image by the symplectic embedding~$I_n$
of the product of the zero section $N_0$ of $T^\ast S^1$ with
the Lagrangian submanifold $L$. If $L$ is monotone in $\mathbb{R}^{2n}$,
then~$\Theta(L)$ is monotone in $\mathbb{R}^{2n+2}$ with the same
monotonicity constant.
We are here particularly interested in the case $n=1$.
We give the expression in
coordinates of the map $I = I_1$ as it will be useful in the
following.
If $\theta \in \mathbb{S}^1, \tau \in T^\ast_{\theta} \mathbb{S}^1, x \in \mathbb{R}$ and $y \in
T^\ast_x \mathbb{R}$,
then
$$I ((\tau,t) , (y,x)) = (p_0, p_1, q_0, q_1)$$
with
$(q_0, q_1) \in \mathbb{R}^2$ and $(p_0, p_1) \in T_{(q_0, q_1)} \mathbb{R}^2$ such
that:
$$ \left \{ {
\begin{array}{ccc}
q_0 & = & e^x \cos (\theta) \\
q_1 & = & e^x \sin (\theta) \\
p_0 & = & e^{-x} ( - \tau \sin (\theta) + y \cos(\theta) ) \\
p_1 & = & e^{-x} ( \tau \cos (\theta) + y \sin(\theta) ).
\end{array}
} \right.
$$
Identifying $T^\ast \mathbb{R}^2$ with $\mathbb{C}^2$ via the symplectomorphism
$$
\begin{array}{ccc}
T^\ast \mathbb{R}^2 & \longrightarrow & \mathbb{C}^2 \\
(q_0, q_1, p_0, p_1) & \longmapsto & (q_0 + i p_0, q_1 + i p_1),
\end{array}
$$
the map $I$ can be written as the following embedding of $T^\ast(\mathbb{S}^1
\times \mathbb{R})$ into $\mathbb{C}^2$:
$$\left({(\tau,\theta) , (y,x)} \right) \mapsto
\left ( {
\begin{array}{ccc}
z_0 & = & (e^{x} + i e^{-x} y) \cos(\theta) - i \tau e^{-x} \sin(\theta)\\
z_1 & = & (e^{x} + i e^{-x} y) \sin(\theta) + i \tau e^{-x} \cos(\theta)
\end{array}
} \right).
$$
For $n=1$, Chekanov's construction with $L$ the circle centered at the
origin of area $2 r^2$ can be parametrized in $\mathbb{C}^2$ by:
\begin{equation}
\label{eq:ParamChek}
\left \{ {
\begin{array}{ccc}
z_0 & = & (e^{x} + i e^{-x} y) \cos(\theta)\\% = (e^{r \cos(s)} + i e^{-r \cos(s)} r \sin(s)) \cos(\theta) \\
z_1 & = & (e^{x} + i e^{-x} y) \sin(\theta
\end{array}
} \right.
\end{equation}
with $(x,y) \in L$.
It is a monotone Lagrangian torus with monotonicity constant $r^2$.
In \cite[Theorem 4.2]{Chekanov96}, Chekanov proved that this torus
is not Hamiltonian isotopic to the Clifford torus (by versal
deformations, using Ekeland-Hofer capacities and the displeacement
energy). In the following, we will call it Chekanov's torus and denote
it~$\Theta_{\Ch} (r^2)$.
\subsection{The version by Eliashberg and Polterovich and its relation
with the previous construction}
Eliashberg and Polterovich have given in~\cite{El-Pol-1993} another
description of an exotic monotone torus in $\mathbb{R}^4 \simeq \mathbb{C}^2$.
If $D$ is a disk of $\mathbb{C}$ of area $r^2$, $r>0$, which does
not contain the origin, and $c = \{c(s) \; | \; s \in [0, 2 \pi]\}$ is its
boundary, parametrised by a smooth map
$c~:~[0,2\pi]~\rightarrow~\mathbb{C}$, then the torus defined as:
$$\left\{{ \left. \left({c(s) e^{i\theta}, c(s) e^{-i\theta} }\right)
\right| \theta \in [0, 2 \pi], s \in [0, 2 \pi] } \right\}$$
is a monotone torus of monotonicity constant $r^2$ denoted
$\Theta_{\EP}(r^2)$. Eliashberg and Polterovich have proved that this torus
is again not Hamiltonian isotopic to a split Lagrangian torus
(\cite[Proposition 4.2.B]{El-Pol-1993}) by counts of holomorphic
discs with boundary along this torus.
\begin{prop}
\label{prop:EPetCh}
For any positive radius $r$, the monotone torus $\Theta_{\EP}(r^2)$ is
Hamiltonian isotopic to $\Theta_{\Ch}(r^2)$ in $\mathbb{C}^2$.
\end{prop}
We give here a detailed proof of this well known result as we
will use it in the next sections. We fix a positive radius $r$ and to
simplify the notations, we will drop $r$ from the notations in the
proof.
This proposition can be deduced from a series of lemmata.\\
The exotic torus of Eliashberg and Polterovich is constructed as the
orbit of some circle under a Hamiltonian circle action, so that it
satisfies the following:
\begin{lem}
\label{action EP}
The torus $\Theta_{\EP}$ is stable under the following
action $\rho_{\EP}$ of the circle on $\mathbb{C}^2$: for $\theta \in [0;2 \pi]$ and
$(z_0,z_1) \in \mathbb{C}^2$,
$$\rho_{\EP}(e^{i \theta})(z_0,z_1) =
\left ( { \begin{array}{cc} e^{i \theta} & 0 \\ 0 & e^{-i \theta}
\end{array} } \right ) \left ( { \begin{array}{c} z_0 \\ z_1
\end{array} } \right )
= \left ( { \begin{array}{c} e^{i \theta} z_0 \\ e^{-i \theta} z_1
\end{array} } \right )
$$
of Hamiltonian
$$H(z_0,z_1) = \dfrac{1}{2 \pi} \left({ |z_0|^2-|z_1|^2 } \right).$$
\end{lem}
More precisely, the torus $\Theta_{\EP}$ is the orbit of the curve
$$C = \left \{ \left .{ \left ( { \begin{array}{c} c(s) \\ c(s)
\end{array} } \right ) } \right | s \in [0, 2 \pi] \right \}$$
under the action $\rho_{\EP}$.\\
The exotic torus $\Theta_{\Ch}$ satisfies a similar property:
\begin{lem}
\label{lem:actionCh}
The torus $\Theta_{\Ch}$ is stable under the following action $\rho_{\Ch}$ of the
circle on $\mathbb{C}^2$: for $\theta \in [0;2 \pi]$ and $(z_0,z_1) \in
\mathbb{C}^2$,
$$\rho_{\Ch}(e^{i \theta})(z_0,z_1) =
\left ( { \begin{array}{cc} \cos(\theta) & - \sin(\theta) \\
\sin(\theta) & \cos(\theta) \end{array} } \right ) \left ( {
\begin{array}{c} z_0 \\ z_1 \end{array} } \right )
$$
of Hamiltonian
$$H(z_0,z_1) = \dfrac{1}{\pi} \im \left({z_0 \bar{z}_1}\right).$$
\end{lem}
In particular, the parametrisation~(\ref{eq:ParamChek}) gives that the
torus $\Theta_{\Ch}$ is the orbit under the action $\rho_{\Ch}$
of the following curve of $\mathbb{C}^2$:
$$\left \{ { \left . { \left ( { \begin{array}{c} e^x+ i e^{-x} y \\ 0
\end{array} } \right ) } \right | (x,y) \in L } \right \}$$
where $L$ is the circle of $\mathbb{C}$ centered in the origin of
radius~$r$. But this torus can also be described as the
orbit of the curve:
$$\Lambda = \left \{ { \left . { \left ( { \begin{array}{c}
\frac{1}{\sqrt{2}}(e^x+ i e^{-x} y) \\
- \frac{1}{\sqrt{2}}(e^x+ i e^{-x} y)
\end{array} } \right ) } \right | (x,y) \in L } \right \}.$$
\ \\
\begin{lem}
\label{lem:conjug} The two Hamiltonian actions $\rho_{\EP}$ and $\rho_{\Ch}$ are
conjugate inside the special unitary group $SU(2)$.
\end{lem}
\begin{proof}
If we denote by $P$ the matrix
$$P =
\left ( { \begin{array}{cc} \frac{1}{\sqrt{2}}i & - \frac{1}{\sqrt{2}} \\
\frac{1}{\sqrt{2}} & - \frac{1}{\sqrt{2}}i \end{array} }
\right )$$
then $P$ is a matrix of the special unitary group such that for every
$\theta \in [0; 2 \pi]$,
$$\,^t \bar{P}
\left ( { \begin{array}{cc} e^{i \theta} & 0 \\ 0 & e^{-i \theta}
\end{array} } \right ) P
=
\left ( { \begin{array}{cc} \cos(\theta) & - \sin(\theta) \\
\sin(\theta) & \cos(\theta) \end{array} } \right ). $$
\end{proof}
Now let $\Theta_{\Ch}^P$ be the torus obtained as the orbit under the
action $\rho_{\EP}$ of the curve $P \Lambda$ of $\mathbb{C}^2$.
Then $\Theta_{\Ch}^P$ is the image by the map $P$ of the torus $\Theta_{\Ch}$.
Note moreover that the diffeomorphism $P$ of $\mathbb{C}^2$ is a Hamiltonian
diffeomorphism (because $P$ is a matrix of $SU(2)$). This means that
$\Theta_{\Ch}^P$ is Hamiltonian isotopic to $\Theta_{\Ch}$ and it is now sufficient
for Proposition~\ref{prop:EPetCh} to prove that $\Theta_{\Ch}^P$ is
Hamiltonian isotopic to $\Theta_{\EP}$. \\
\begin{lem}
\label{lem:ChPetEP}
The tori $\Theta_{\Ch}^P$ and $\Theta_{\EP}$ are Hamiltonian isotopic inside $\mathbb{C}^2$.
\end{lem}
To prove this, we will use the following lemma:
\begin{lem}[See \cite{Polt-Geom-2001}]
\label{lem:lemmePolterovich}
A Lagrangian isotopy is exact if and only if it can be extended to
an ambient Hamiltonian isotopy.
\end{lem}
Recall (\cite[Section 6.1]{Polt-Geom-2001}) that given a Lagrangian
isotopy of a closed manifold $N$ into a symplectic manifold
$(W,\omega)$:
$$\Phi : N \times [0;1] \rightarrow W,$$
the pull-back $\Phi^\star \omega$ is of the form $\alpha_s \wedge ds$
where $\{\alpha_s\}$ is a family of closed $1$-forms on $N$. The
Lagrangian isotopy is said to be exact if $\alpha_s$ is exact for
all~$s$. \\
\noindent \emph{Proof of Lemma~\ref{lem:ChPetEP}.} \;
Thanks to Lemma~\ref{lem:lemmePolterovich}, it is enough to prove
that the tori $\Theta_{\Ch}^P$ and $\Theta_{\EP}$ are exact Lagrangian isotopic.
We note first that
$$\Lambda^P = P \Lambda =
\left \{ { \left . \frac{1}{\sqrt{2}}(e^x+ i e^{-x} y)
\frac{1}{\sqrt{2}} (1 + i)
{ \left ( { \begin{array}{c}
1 \\
1
\end{array} } \right ) } \right | (x,y) \in L } \right \},$$
so that $C$ and $\Lambda^P$ are both curves lying in the
diagonal of $\mathbb{C}^2$.
Moreover, as $c$ is the boundary of the disc $D$, the integral of
the Liouville form of $\mathbb{C}$ is equal to $r^2$ on the curve $c$.
On the other hand, let $f$ be the map
$$
\begin{array}{cccc}
f: & T^\ast \mathbb{R} & \longrightarrow & \mathbb{C} \\
& (x,y) & \longmapsto & e^{x}+ i e^{-x} y.
\end{array}
$$
The map $f$ is an exact symplectomorphism (i.e. it preserves the
Liouville form $p dq$) from $T^\ast \mathbb{R}$ onto its image
$$\{z \in \mathbb{C} \; | \; \Re e(z) > 0 \}.$$
In particular, the integral of the Liouville form on the curve $f(L)$
is equal to the integral of the Liouville form on $L$. The rotation
$z~\mapsto~\frac{1}{\sqrt{2}} (1+i) z$ preserves the Liouville
form of~$\mathbb{C}$ so that the integral of the Liouville form is also equal
to~$r^2$ on the curve
$$L^P=\frac{1}{\sqrt{2}} f(L) \frac{1}{\sqrt{2}} (1 + i).$$
As $L^P$ and $c$ are two closed curves in $\mathbb{C}$ on which the integral of
the Liouville form takes the same value (that is they are the boundary of
domains of the same area), these two curves can be Hamiltonianly
isotoped one into the other in $\mathbb{C}$ (this can also be seen by applying
Lemma~\ref{lem:lemmePolterovich}).
We denote by $\varphi: \mathbb{C} \times [0;1] \longrightarrow \mathbb{C}$ a Hamiltonian
isotopy of $\mathbb{C}$ such that:
$$\varphi_0 = \Id \mbox{ and } \varphi_1\left({L^P}\right) = c.$$
As $L^P$ and $c$ are boundary of domains which do not contain
the origin, the Hamiltonian isotopy can be chosen so that:
\begin{equation}
\label{eq:nonsing}
\forall \; t \in [0;1], \; \varphi_t(0) = 0,
\end{equation}
(in other words, $\varphi_t\left({L^P}\right)$ never
crosses the origin).
The two curves $\Lambda^P$ and $C$ are then Hamiltonian isotopic
inside the diagonal $\Delta_{\mathbb{C}^2}$ of $\mathbb{C}^2$ via the isotopy
$(\varphi, \varphi)$, and we use this Hamiltonian isotopy and the
circle action $\rho_{\EP}$ to construct an exact Lagrangian isotopy
from~$\Theta_{\Ch}^P$ to~$\Theta_{\EP}$. This Lagrangian isotopy
$\Phi: \Theta_{\Ch}^P \times [0;1] \longrightarrow \mathbb{C}^2$ is
defined for $t \in [0;1]$ by:
$$
\begin{array}{cccc}
\Phi_t: & \Theta_{\Ch}^P & \longrightarrow & \mathbb{C}^2 \\
& \left ( { \begin{array}{c}
e^{i \theta} z\\
e^{-i \theta} z
\end{array} } \right ) &
\longmapsto & \left ( { \begin{array}{c}
e^{i \theta} \varphi(z,t)\\
e^{-i \theta} \varphi(z,t)
\end{array} } \right )
\end{array}
$$
and it is a well defined Lagrangian isotopy because of
property~(\ref{eq:nonsing}).
As $\varphi$ is a Hamiltonian isotopy and $\rho_{\EP}$ is a
Hamiltonian circle action, one can check that $\Phi$ is
exact. Thanks to Lemma~\ref{lem:lemmePolterovich}, this
ends the proof of Lemma~\ref{lem:ChPetEP} and proves
Proposition~\ref{prop:EPetCh}. \hfill \qedsymbol
\section{In the complex projective plane}
In this section, the projective complex plane $\mathbb{C} \mathbb{P}^2$ will be
endowed with the Fubini-Study symplectic form normalized so
that the area of any complex projective line is
$$\int_{\mathbb{C} \mathbb{P}^1} \omega_{\FS} = 2.$$
This implies in particular that:
\begin{itemize}
\item The monotonicity constant of $\mathbb{C} \mathbb{P}^2$ is $\frac{2}{3}$ and
any monotone Lagrangian submanifold in $\mathbb{C} \mathbb{P}^2$ has monotonicity
constant $\frac{1}{3}$.
\item With the normalisation (\ref{normalisation}) of the symplectic
form of $\mathbb{R}^4 \simeq \mathbb{C}^2$, the open ball $B(2)$ of radius
$\sqrt{2}$ is symplectically embedded into $\mathbb{C} \mathbb{P}^2$ via the canonical
embedding:
\begin{equation}
\label{embball}
\begin{array}{cccc}
E_2:& B(2) & \longrightarrow & \mathbb{C} \mathbb{P}^2 \\
&(z_0,z_1)& \longmapsto &
\left[ \, {z_0:z_1:\sqrt{2-|z_0|^2-|z_1|^2} } \, \right]
\end{array}
\end{equation}
\end{itemize}
Because of the symplectic embedding~(\ref{embball}), a way to
construct a monotone torus into $\mathbb{C} \mathbb{P}^2$ is to begin with a monotone
torus of monotonicity constant~$\frac{1}{3}$ in $B(2)$ and embed
this torus into $\mathbb{C} \mathbb{P}^2$ via $E_2$. This is for example a possible
construction of the Clifford torus: one begins with the product of
the two circles centered in the origin and of
radius~$\sqrt{\frac{2}{3}}$ in each factor of~$\mathbb{C}^2 = \mathbb{C} \times \mathbb{C}$.
One could try to embed also Chekanov's torus as a monotone Lagrangian
torus of $\mathbb{C} \mathbb{P}^2$ via $E_2$. In order to get the right monotonicity
constant, one should begin with $L$ a circle of area $\frac{2}{3}$.
Unfortunately, one can check that the torus $\Theta_{\Ch}(\frac{2}{3})$
does not sit in the open ball $B(2)$ (for example with the
parametrisation~(\ref{eq:ParamChek}), for $s=0$ and any $\theta$,
one has $\|z\| > \sqrt{2}$).
Eliashberg and Polterovich's torus $\Theta_{\EP}$ as it has been defined in
Section~\ref{sec:C2} is also not lying in the ball $B(2)$, so that it
cannot be embedded inside the complex projective plane $\mathbb{C} \mathbb{P}^2$. But this
construction can be modified in order to define an exotic torus in $\mathbb{C} \mathbb{P}^2$.
\subsection{Chekanov and Schlenk's torus in the complex projective
plane}
\label{sec:CS-torus}
Instead of defining a torus as the orbit under $\rho_{\EP}$ of the boundary
of a disk which does not contain the origin, one can also define a
torus as the orbit of any closed embedded curve which is the boundary
of a domain which does not contain the origin of $\mathbb{C}$. The first
torus of the family of monotone tori in~$\mathbb{C} \mathbb{P}^n$ defined by Chekanov
and Schlenk~(\cite{Chek-SchI}, \cite{Chek-Sch-MSRI}) is constructed this way.
To be precise, replace the circle $c$ in the Eliashberg-Polterovich
construction by the curve $\gamma$ which is the boundary
of a domain of area $\frac{1}{3}$ sitting inside the disk of
radius~$1$ centered in the origin of $\mathbb{C}$ and inside the half-plane
of complex numbers of positive real part (see Figure \ref{fig:gamma}).
\begin{figure}[htbp]
\label{fig:gamma}
\begin{center}
\psfrag{C}{$\mathbb{C}$}
\psfrag{1}{$1$}
\psfrag{0}{$0$}
\psfrag{gamma}{$\color{red} \mathbf{\gamma}$}
\includegraphics{Gamma.eps}
\caption{The curve $\gamma$}
\end{center}
\end{figure}
Then the
curve
$$\Gamma = \left ( {\begin{array}{c}
\gamma\\
\gamma
\end{array} } \right )$$
is lying inside the ball $B(2)$ and as the action of the cercle $\rho_{\EP}$
is by multiplication by matrices of $SU(2)$, the orbit of this curve
is also entirely contained in $B(2)$. The torus constructed this way
can then be embedded in~$\mathbb{C} \mathbb{P}^2$ through the canonical
embedding~(\ref{embball}) and will be denoted~$\Theta_{\CS}$. Chekanov and
Schlenk have proved (see \cite{Chek-SchII}, \cite{Chek-Sch-MSRI})) that
this torus is not Hamiltonian isotopic to the Clifford torus in~$\mathbb{C} \mathbb{P}^2$
and is not Hamiltonianly displaceable in $\mathbb{C} \mathbb{P}^2$.
\subsection{The torus of Biran and Cornea}
The Biran-Cornea torus is defined using the Lagrangian circle bundle
construction of Biran (see~\cite{Biran-Ciel-Subcrit}
and~\cite{Biran-NonIntersections}).
Let $\Sigma$ be the quadric of $\mathbb{C} \mathbb{P}^2$, given by the homogeneous
equation:
$$z_0^2+z_1^2+z_2^2=0.$$
Then $(\mathbb{C} \mathbb{P}^2, \omega, J ; \Sigma)$ (where $J$ is the standard complex
structure on $\mathbb{C} \mathbb{P}^2$) is a polarized K\"ahler manifold of degree $1$ in
the sense of~\cite{Biran-Barriers} and~\cite{Biran-NonIntersections},
which means that $\Sigma$ is a smooth and reduced complex hypersurface
of the K\"ahler manifold $(W, \omega, J)$ (with $W = \mathbb{C} \mathbb{P}^2$) whose
homology class $[\Sigma]~\in~H_2(W,\mathbb{Z})$ represents the Poincar\'e
dual of~$[\omega]$.
In the case of such a polarisation, Biran proved
in~\cite{Biran-Barriers} that there exists an isotropic $CW$-complex
$\Delta~\in~(W,\omega)$ whose complement $(W \symbol{92} \Delta, \omega)$
is symplectomorphic to a standard symplectic disc bundle
$(E,\omega_0)$ modeled on the normal bundle $N_\Sigma$ of $\Sigma$ in
$W$ and whose fibres have area $1$.
The standard symplectic disc bundle is by definition the open unit
disc subbundle of the complex line bundle $\pi: N_\Sigma \rightarrow
\Sigma$ (with respect to some Hermitian metric), endowed with the
symplectic structure $\omega_0$ given by the following formula:
$$\omega_0 = \pi^\star \omega_{|\Sigma} + d(r^2 \alpha),$$
where $r$ is the radial coordinate on the fibres of $E_\Sigma
\rightarrow \Sigma$ defined using the Hermitian metric and $\alpha$ is
a connection $1$-form on $E_\Sigma \symbol{92} \Sigma$ with curvature
$$d \alpha = - \pi^\star(\omega_\Sigma).$$
In the case $\Sigma$ is the quadric above, $\Delta$ is $\mathbb{R}
\mathbb{P}^2$.\\
Biran and Cornea's torus is then constructed the following way
(see~\cite[Section 6.4.1]{Bir-Cor-uniruling}): the quadric $\Sigma$ is
topologically a sphere, so that we can consider an
equator~$\mathcal{E}$ of~$\Sigma$, that is a circle which divides the
sphere into two discs of equal areas. Let now $\Theta_{\BC}$ be the
restriction to the equator $\mathcal{E}$ of the circle subbundle of
radius $\sqrt{\frac{2}{3}}$ inside the disc bundle $E_\Sigma
\rightarrow \Sigma$. This defines an exotic monotone torus in
$\mathbb{C} \mathbb{P}^2$ whose Floer homology (with $\mathbb{Z}/2$ coefficients) is trivial.
The following remark of Biran on the description of $E_\Sigma$ for the
quadric will be crucial in the proof that $\Theta_{\BC}$ is Hamiltonian
isotopic to $\Theta_{\CS}$ in~$\mathbb{C} \mathbb{P}^2$. It was also used recently by Opshtein
to prove a symplectic embedding Theorem (see~\cite[Theorem 4]{Ophstein}).
\begin{prop}
\label{prop:normal}
Let $a, b,$ and $c$ be three real numbers, not identically zero, and
let $D_{a,b,c}$ denote the projective line of $\mathbb{C} \mathbb{P}^2$ defined by the
homogeneous equation
\begin{equation}
\label{D}
a z_0 + b z_1 + c z_2 = 0.
\end{equation}
The line $D_{a,b,c}$ intersects the quadric $\Sigma$ in two complex
conjugate points $x$ and $\bar{x}$ and $\mathbb{R} \mathbb{P}^2$ cuts this line along
a circle in two complex conjugate discs which are the fibers of
$E_\Sigma$ in $x$ and $\bar{x}$, respectively.
\end{prop}
\begin{proof}
The projective line $D_{a,b,c}$ is a curve of degree $1$ and
intersects the curve $\Sigma$ of degree $2$ in two points.
As the line and the quadric are defined by equations with real
coefficients, their intersection points are complex conjugate.
Moreover, they are distinct because the quadric has no real point.
Let $x = [x_0, x_1, x_2]$ be an intersection point of $D_{a,b,c}$
and $\Sigma$. The tangent space $T_x$ at the quadric in this point $x$
is defined by the homogeneous equation:
\begin{equation}
\label{T}
x_0 z_0 + x_1 z_1 + x_2 z_2 = 0.
\end{equation}
In the complex linear space $\mathbb{C}^3$, the two equations~(\ref{D})
and~(\ref{T}) define two complex hyperplanes $\tilde{D}$ and
$\tilde{T}$ respectively. They are distinct as $(x_0, x_1, x_2)$
is not a multiple of a real vector. As a consequence, they have
an intersection of complex dimension~$1$.
But the vector $(x_0, x_1, x_2)$ of $\mathbb{C}^3$ is a non-zero vector
in the intersection of $\tilde{D}$ and $\tilde{T}$ so that
$$\tilde{D} \cap \tilde{T} = \mbox{Vect}_\mathbb{C} (x_0, x_1, x_2).$$
Moreover the vector $(\bar{x}_0, \bar{x}_1, \bar{x}_2)$ belongs
to $\tilde{D}$ and is orthogonal to $(x_0, x_1, x_2)$ so that
$\tilde{D}$ is the direct orthogonal sum of the two vector
spaces spanned by $(x_0, x_1, x_2)$ and
$(\bar{x}_0, \bar{x}_1, \bar{x}_2)$, respectively.
Note also that the vector $(\bar{x}_0, \bar{x}_1, \bar{x}_2)$ is
in the Hermitian-ortogonal to the vector space $\tilde{T}$.
Consequently, the complex projective lines $D_{a,b,c}$ and $T_x$ are
orthogonal for the induced Hermitian product on $\mathbb{C} \mathbb{P}^2$.
The complex projective line $D_{a,b,c}$ is stable under complex
conjugation and its invariant submanifold $D_{a,b,c} \cap \mathbb{R} \mathbb{P}^2$
is a circle which divides $D_{a,b,c}$ in two discs which are
images of each other by the complex conjugation, and thus have
the same area (equal to $1$ with our conventions). Moreover,
each disc contains only one of the intersection points with the
quadric $x$ or~$\bar{x}$.
To be sure that the two discs above are exactly the fibers
of~$E_\Sigma$, one should be sure that the discs are ``centered''
in the intersection points.
One way to prove this is to see that it is true in the case $a=1$,
$b=0$ and $c=0$ and it can be extended to the other cases thanks
to a transformation of~$U(3)$. \end{proof}
\begin{rem}
Through any point of~$\mathbb{C}^3$ passes a complex line
defined by an equation with real coefficients
for dimensional reason, the Hermitian orthogonal supplement
of a point must intersect $\mathbb{R}^3$). As a consequence,
Proposition~\ref{prop:normal} describes the bundle $E_\Sigma$
in any point of the quadric $\Sigma$.
\end{rem}
\subsection{Biran and Cornea's torus is Hamiltonian isotopic to Chekanov and Schlenk's torus}
\begin{thm}
The Biran and Cornea's torus is Hamitonian isotopic to Chekanov and
Schlenk's torus in $\mathbb{C} \mathbb{P}^2$.
\end{thm}
\begin{proof}
First we also define a modified Chekanov torus using the
curve~$\gamma$. Let~$\widetilde{\Gamma}$ be the curve
$$\widetilde{\Gamma} = \left ( {\begin{array}{c}
\gamma\\
- \gamma
\end{array} } \right )$$
and $\widetilde{\Theta}_{\Ch}$ the torus defined as the orbit of the curve $\widetilde{\Gamma}$
under the action of~$\rho_{\Ch}$.
Note that as the curve $\gamma$ is Hamiltonian isotopic to the
curve~$L$ in $\mathbb{C}$, the same proof as the one of
Lemma~\ref{lem:ChPetEP} proves
that $\widetilde{\Theta}_{\Ch}$ is Hamiltonian isotopic to $\Theta_{\Ch}$ in $\mathbb{C}^2$.
The same way as for $\Theta_{\CS}$, because $\widetilde{\Gamma}$ is contained in $B(2)$
and because the action of the cercle $\rho_{\Ch}$ is by multiplication by
matrices of $SU(2)$, the torus $\widetilde{\Theta}_{\Ch}$ is entirely contained in $B(2)$
and can be embedded in $\mathbb{C} \mathbb{P}^2$ as a monotone torus.
Let also $\widetilde{\Theta}_{\Ch}^P$ be the torus defined as the orbit under the action
$\rho_{\EP}$ of the curve $P \Gamma$. As before,
$$\widetilde{\Theta}_{\Ch}^P = P \widetilde{\Theta}_{\Ch},$$
so that $\widetilde{\Theta}_{\Ch}^P$ and $\widetilde{\Theta}_{\Ch}$ are Hamiltonian isotopic. Moreover,
the Hamiltonian isotopy is given by a family of matrices of $U(2)$ so
that this isotopy is preserving the ball $B(2)$.
As a consequence, this isotopy enables to define an isotopy between
$\widetilde{\Theta}_{\Ch}^P$ and $\widetilde{\Theta}_{\Ch}$ inside the complex projective plane.
Now notice that in $B(2)$,
$$\Theta_{\CS} = R_{-\frac{\pi}{4}} \widetilde{\Theta}_{\Ch}^P$$
where $R_{-\frac{\pi}{4}}$ is the rotation $e^{- i \frac{\pi}{4}}$ in
each factor of $\mathbb{C}^2$, so that $\Theta_{\CS}$ is the image of $\widetilde{\Theta}_{\Ch}^P$
by a Hamiltonian diffeomorphism.
To prove the theorem, we will prove that $\widetilde{\Theta}_{\Ch}$ is Hamiltonian
isotopic to~$\Theta_{\BC}$.
The torus~$\widetilde{\Theta}_{\Ch}$ can be defined in homogeneous coordinates
of~$\mathbb{C} \mathbb{P}^2$ as the set:
$$\widetilde{\Theta}_{\Ch} = \left \{ \left. [\cos(\theta) \sqrt{2} \gamma(s): \sin(\theta) \sqrt{2} \gamma(s) : \sqrt{2 - 2 |\gamma(s)|^2}] \; \right| \; \theta, s \in [0, 2 \pi] \right\}.$$
It is Hamiltonian isotopic in $\mathbb{C} \mathbb{P}^2$ to any torus obtained
rotation inside the plane of the torus~$\widetilde{\Theta}_{\Ch}$, and in
particular to the torus
$$e^{i \frac{\pi}{2}} \widetilde{\Theta}_{\Ch}$$
described in $\mathbb{C} \mathbb{P}^2$ as the set
$$\widetilde{\Theta}_{\Ch}' = \left \{ \left. [\cos(\theta) \sqrt{2} \gamma(s): \sin(\theta) \sqrt{2} \gamma(s) : i \sqrt{2 - 2 |\gamma(s)|^2}] \; \right| \; \theta, s \in [0, 2 \pi] \right\}.$$
The latter can also be seen as the image of $\widetilde{\Theta}_{\Ch}$ by the
following modified embedding of the ball $B(2)$ inside $\mathbb{C} \mathbb{P}^2$:
\begin{equation}
\label{embball2}
\begin{array}{cccc}
\widetilde{E}_2:& B(2) & \longrightarrow & \mathbb{C} \mathbb{P}^2 \\
& (z_0,z_1)& \longmapsto &
\left[ \, {z_0:z_1: i \sqrt{2-|z_0|^2-|z_1|^2} } \, \right].
\end{array}
\end{equation}
It is thus enough to prove that the torus
$\Theta_{\BC}$ is Hamiltonian isotopic to $\widetilde{\Theta}_{\Ch}'$ in $\mathbb{C} \mathbb{P}^2$ and
for that purpose, it will be more convenient to work with the
embedding~$\widetilde{E}_2$.
Let $Z$ be the cylinder of $T^\ast \mathbb{S}^1 \times \{0\} \subset \ T^\ast (\mathbb{S}^1 \times \mathbb{R})$ defined as:
$$Z = \left \{ {(\theta,\tau,0,0) \in T^\ast (\mathbb{S}^1 \times \mathbb{R}^n) \;|\; |\tau| < 1 } \right \}.$$
The image of this cylinder by Chekanov's map $I$ can be parametrized
in~$T^\ast \mathbb{R}^2$ by:
$$ \left \{ {
\begin{array}{ccc}
q_0 & = & \cos (\theta) \\
q_1 & = & \sin (\theta) \\
p_0 & = & - \tau \sin (\theta) \\
p_1 & = & \tau \cos (\theta)
\end{array}
} \right.
$$
for $|\tau| < 1$ and in $\mathbb{C}^2$ by:
$$\left ( {
\begin{array}{ccc}
z_0 & = & \cos(\theta) - i \tau \sin(\theta)\\
z_1 & = & \sin(\theta) + i \tau \cos(\theta)
\end{array}
} \right).
$$
As $|\tau| < 1$, any point in the image by $I$ of the cylinder $Z$ is
lying inside the ball $B(2)$. Its image $I(Z)$ under the embedding
(\ref{embball2}) is parametrized by:
$$\{ [\cos(\theta) - i \tau \sin(\theta): \sin(\theta) + i \tau \cos(\theta) : i \sqrt{1-\tau^2}] \; | \; s \in [0, 2 \pi], \tau \in (-1,1) \}$$
\begin{lem}
The quadric $\Sigma$ is the union of the image $I(Z)$ and the two points
``at infinity'' $[1:i:0]$ and $[1:-i:0]$.
\end{lem}
\begin{proof}
Any point of $I(Z)$ satisfies the equation of the quadric. Moreover, when
$\tau$ tends to $\varepsilon \in \{-1;1\}$, the point of homogeneous
coordinates
$$[\cos(\theta) - i \tau \sin(\theta): \sin(\theta) + i \tau \cos(\theta) :
i \sqrt{1-\tau^2}]$$
tends to the point $[1 : \varepsilon : 0]$, so that $I(Z)$ can be
compactified in the closed surface $\Sigma$.
\end{proof}
\begin{lem}
The image by $I$ of the zero section $\tau = 0$ is an equator of the
quadric.
\end{lem}
Indeed, the zero section $\tau = 0$ of $T^\ast \mathbb{S}^1$ cuts the cylinder
$Z$ in two pieces of equal area so that its image cuts the quadric
in two discs of same area $1$, one disc containing the point
$[1 : i : 0]$ and the other one the point $[1 : -i : 0]$. In the
homogeneous coordinates of $\mathbb{C} \mathbb{P}^2$, this equator can be parametrised by:
$$[\cos(\theta):\sin(\theta):i],\; \theta \in [0 ; 2 \pi].$$
\begin{lem}
For a fixed $\theta \in \mathbb{R}$, the curve
$$\{[\cos(\theta) \sqrt{2} \gamma(s): \sin(\theta) \sqrt{2} \gamma(s) : i \sqrt{2 - 2 |\gamma(s)|^2}]\; | \; s \in [0, 2 \pi] \}$$
lies inside the fiber of $E_\Sigma$ at the point
$[\cos(\theta):\sin(\theta):i]$.\\
Moreover, this curve enclose a domain of area $\frac{2}{3}$ inside the
fiber at the base point $[\cos(\theta):\sin(\theta):i]$.
\end{lem}
\begin{proof}
Thanks to Proposition~\ref{prop:normal}, we know how to describe the
disc bundle~$E_\Sigma$ at a point of the quadric as soon as we have
the equation of a real projective line containing this point.
In the case of a point $[\cos(\theta):\sin(\theta):i]$ on the equator
of the $\Sigma$, it lies on the real line of homogeneous equation
$$\sin(\theta) z_0 - \cos(\theta) z_1 = 0.$$
Because of Proposition~\ref{prop:normal}, this means that the
fibers of the disc bundle $E_\Sigma$ in the points
$$[\cos(\theta):\sin(\theta):i] \mbox{ and }
[\cos(\theta):\sin(\theta):-i]=[\cos(\theta+\pi):\sin(\theta+\pi):i]$$
are the two discs of the projective line $N_\theta$ of equation
$$\sin(\theta) z_0 - \cos(\theta) z_1 = 0$$
separated by the curve $N_\theta \cap \mathbb{R} \mathbb{P}^2$.
Any point of the curve
$$\{[\cos(\theta) \sqrt{2} \gamma(s): \sin(\theta) \sqrt{2} \gamma(s) : i \sqrt{2 - 2 |\gamma(s)|^2}] \; | \; s \in [0, 2 \pi] \}$$
belongs to the line $N_\theta$.
Moreover, this curve does not intersect $\mathbb{R} \mathbb{P}^2$ as the curve $\gamma$
does not intersect neither the imaginary axes nor the circle of radius~$1$.
Therefore, this curve is entirely contained in one of the fiber of
the disc bundle $E_\Sigma$.
Actually, the image of the half disc
$$\{z \in \mathbb{C} \; | \; \Re e (z) > 0 \mbox{ and } |z|< 1\}$$
by the map
$$z \mapsto \{[\cos(\theta) \sqrt{2} z: \sin(\theta) \sqrt{2} z : i \sqrt{2 - 2 |z|^2}]\}$$
is contained in $N_\theta$ and does not intersect $\mathbb{R} \mathbb{P}^2$. Moreover the
area of the image of the half disc is equal to $1$ (that is twice the
area of the original half disc) so that it is the entire fiber of $E_\Sigma$
at the point $[\cos(\theta):\sin(\theta):i]$, image of the complex number
$\frac{1}{\sqrt{2}}$.
The curve $\gamma$ is enclosing a domain of area~$\frac{1}{3}$, and
consequently the curve
$$\{[\cos(\theta) \sqrt{2} \gamma(s): \sin(\theta) \sqrt{2}
\gamma(s) : i \sqrt{2 - 2 |\gamma(s)|^2}] \; | \; s \in [0, 2 \pi] \}$$
is enclosing a domain of area $\frac{2}{3}$.
\end{proof}
To end the proof, one can see that in any fiber of the normal bundle,
the curve
$$\{[\cos(\theta) \sqrt{2} \gamma(s): \sin(\theta) \sqrt{2} \gamma(s) : i \sqrt{2 - 2 |\gamma(s)|^2}] \; | \; s \in [0, 2 \pi] \}$$
is isotopic to the circle of area $\frac{2}{3}$ (used
to construct the torus of Biran and Cornea). In order to construct
a Lagrangian isotopy between $\widetilde{\Theta}_{\Ch}'$ and $\Theta_{\BC}$, one can use
the isotopy defined for the fiber of the point corresponding to
$\theta = 0$ and then extend this isotopy by the action of the
circle~$\rho_{\Ch}$. Lemma~\ref{lem:lemmePolterovich} enables then to
extend this exact Lagrangian isotopy into a Hamiltonian isotopy
of~$\mathbb{C} \mathbb{P}^2$.
\end{proof}
\section{In the product of two two-dimensional spheres}
Let us now consider $W = \mathbb{S}^2 \times \mathbb{S}^2 = \mathbb{C} \mathbb{P}^1 \times \mathbb{C} \mathbb{P}^1$
endowed with the symplectic form
$$\omega_W = \omega_{\FS} \oplus \omega_{\FS}$$
where $\omega_{\FS}$ is normalized such that
$$\int_{\mathbb{C} \mathbb{P}^1} \omega_{\FS} = 1,$$
which means that $\mathbb{C} \mathbb{P}^1$ is symplectomorphic to the sphere of
radius~$\frac{1}{2}$ in~$\mathbb{R}^3$ with our conventions.
With this normalisation, the product of the two spheres is monotone
with monotonicity constant
$$K_W = \frac{1}{2}.$$
As a consequence, the monotonicity constant of any monotone
Lagrangian submanifold in this product is
$$K_L = \frac{1}{4}.$$
In this symplectic manifold, one has also the two corresponding constructions
of exotic monotone Lagrangian torus.
\subsection{Chekanov and Schlenk's torus in the product of spheres}
\label{sec:CS-torus-insphere}
For the construction by Chekanov and Schlenk, one has this time to
begin with a curve $\gamma$ enclosing a domain of
area~$\frac{1}{4}$ in the complex
plane inside the half-disc of radius~$1$ and positive real part.
With this choice of curve $\gamma$ the torus of $\mathbb{C}^2$ obtained
by action of $\rho_{\EP}$
$$\left\{{ \left. \left({\gamma(s) e^{i\theta}, \gamma(s)
e^{-i\theta} }\right) \right| \theta \in [0, 2 \pi],
s \in [0, 2 \pi] } \right\}$$
is contained into the product $B(1) \times B(1)$, where $B(1)$ is
the ball of radius~$1$ in $\mathbb{C}$.
As this product can be symplectically embedded into
the product $\mathbb{C} \mathbb{P}^1 \times \mathbb{C} \mathbb{P}^1$ via
\begin{equation}
\label{embprodball}
\begin{array}{cccc}
E_{1,1}: & B(1) \times B(1) & \longrightarrow & \mathbb{C} \mathbb{P}^1 \times \mathbb{C} \mathbb{P}^1 \\
& (z_0,z_1) & \longmapsto &
\left({ \left[ \, {z_0:\sqrt{1-|z_0|^2}} \, \right] , \left[ \, {z_1:\sqrt{1-|z_1|^2}} \, \right] } \right),
\end{array}
\end{equation}
this construction produces a monotone Lagrangian torus
still denoted~$\Theta_{\CS}$ in~$\mathbb{C} \mathbb{P}^1 \times \mathbb{C} \mathbb{P}^1$ which is not Hamiltonian
isotopic to the Clifford torus and
non displaceable (see \cite{Chek-Sch-MSRI}, \cite{Chek-SchII}).
\subsection{The torus constructed by Biran's circle bundle construction}
The construction of Biran involves again a polarisation. This
time, the symplectic hypersurface $\Sigma$ is the diagonal
sphere described as
$$\Sigma = \{ (x,x) \in \mathbb{S}^2 \times \mathbb{S}^2 | \; x \in \mathbb{S}^2\}$$
in $\mathbb{S}^2 \times \mathbb{S}^2$, or as the hypersurface of $\mathbb{C} \mathbb{P}^1 \times \mathbb{C} \mathbb{P}^1$ satisfying
the homogeneous equation:
$$z_0 w_1 = z_1 w_0,$$
where $[z_0:w_0]$ are the homogeneous coordinates on the first
copy of $\mathbb{C} \mathbb{P}^1$ and $[z_1:w_1]$ on the second.
The total space of the standard symplectic disc
bundle~$(E_\Sigma,\omega_0)$ modeled on the normal
bundle~$N_\Sigma$ of $\Sigma$ in $W$ and whose fibres have
area $1$ is in the case of $\mathbb{C} \mathbb{P}^1 \times \mathbb{C} \mathbb{P}^1$ symplectomorphic
to the complement of the antidiagonal:
$$\Delta = \{ (x,-x) \in \mathbb{S}^2 \times \mathbb{S}^2 | \; x \in \mathbb{S}^2\}$$
described in $\mathbb{C} \mathbb{P}^1 \times \mathbb{C} \mathbb{P}^1$ as the Lagrangian submanifold
of homogeneous equation:
$$z_0 \bar{z}_1 + w_0 \bar{w}_1 = 0.$$
The exotic torus, still denoted $\Theta_{\BC}$, is now defined as the
restriction of the circle subbundle of radius $\frac{1}{\sqrt{2}}$
over an equator of $\Sigma$.
This torus is known to be equal to the exotic monotone Lagrangian
torus of Entov and Polterovich (\cite[Exemple 1.22]{En-Pol-Rigid},
see also the study in~\cite[Section~2]{El-Polt-QS}):
$$K = \left \{ {\left(x,y) \right) \in \mathbb{S}^2 \times \mathbb{S}^2 \; \left|{ \; x_3 + y_3 = 0, x_1 y_1 + x_2 y_2 + x_3 y_3 = - \frac{1}{2} }\right. } \right \},$$
with $\mathbb{S}^2$ being in their conventions the sphere of radius $1$ of
$\mathbb{R}^3$, the symplectic form being rescaled.
It is also equal to the torus defined in the cotangent bundle of
$\mathbb{S}^2$ by the geodesic flow (see~\cite{MR2417887}) and embedded
in a suitable way in a Weinstein's neighbourhood of the Lagrangian
sphere $\Delta$.
\subsection{The two constructions are Hamiltonian isotopic}
\begin{thm}
Biran's exotic torus is Hamitonian isotopic to Chekanov and
Schlenk's exotic torus in $\mathbb{S}^2 \times \mathbb{S}^2$.
\end{thm}
\begin{proof}
In the case of the product of spheres, we cannot use the
corresponding modified Chekanov's torus as in the case of the
complex projective plane because this torus does not embed into
the product of balls of radius $1$.
However, in $\mathbb{C}^2$, the modified Chekanov's torus $\widetilde{\Theta}_{\Ch}$ sits
in the normal bundle of the image $I(Z)$ of the cylinder
$Z$ of the
original description, and more precisely in the fibers over the
image of the zero section.
We also know from Section~\ref{sec:CS-torus}
that $\Theta_{\CS}$ is the image of $\widetilde{\Theta}_{\Ch}$ by the composition of the
rotations $R_{-\frac{\pi}{4}}$ and the Hamiltonian diffeomorphism
described by the matrix~$P$. Therefore it also sits in the normal
bundle of a surface $Q$, namely
the image of $I(Z)$ by $R_{-\frac{\pi}{4}} P$.
This surface $Q$ can be parametrized, with the
coordinates $(\tau, \theta)$ coming from the original
coordinates on $T^{\ast} \mathbb{S}^1$ by:
$$\left ( {
\begin{array}{ccc}
z_0 & = & \left(\frac{1}{\sqrt{2}} - \frac{1}{\sqrt{2}} \tau \right) e^{i (\theta + \pi/4)}\\
z_1 & = & \left(\frac{1}{\sqrt{2}} + \frac{1}{\sqrt{2}} \tau \right) e^{-i (\theta + \pi/4)}
\end{array}
} \right).
$$
This surface is, as expected, invariant by the action of
$\rho_{\EP}$ and the torus~$\Theta_{\CS}$ (for the moment considered
in~$\mathbb{C}^2$) lies in the normal bundle of the
surface along the curve $\tau = 0$. Moreover,
as $\tau \in (0,1)$, $Q$ is contained in the
product~$B(1) \times B(1)$ and can be embedded into
$\mathbb{S}^2 \times \mathbb{S}^2$.
We would like to relate this surface to the diagonal $\Sigma$ as
in the proof of the $\mathbb{C} \mathbb{P}^2$ case. However, the diagonal $\Sigma$
is not invariant by the action of the cercle $\rho_{\EP}$ but by the following action of the circle which can be written on~$\mathbb{C}^2$:
$$\rho_{\BC}(e^{i \theta})(z_0,z_1) =
\left ( { \begin{array}{cc} e^{i \theta} & 0 \\ 0 & e^{i \theta}
\end{array} } \right ) \left ( { \begin{array}{c} z_0 \\ z_1
\end{array} } \right )
$$
This action cannot be conjugate into $SU(2)$, but it will be enough
to conjugate the two circle actions inside $SO(3) \times SO(3)$ using
the description of $W$ as the product of two spheres of $\mathbb{R}^3$.
As the rotation $e^{i \theta}$ on $\mathbb{C} \mathbb{P}^1$ corresponds to the
rotation of matrix
$$ \left ( { \begin{array}{ccc}
\cos \theta & -\sin \theta & 0 \\
\sin \theta & \cos \theta & 0 \\
0 & 0 & 1
\end{array} } \right )$$
one sees that $\rho_{\EP}$ and $\rho_{\BC}$ are conjugate in
$SO(3) \times SO(3)$, for example by the pair $(P_1,P_2)$
where $P_1$ is the identity matrix and
$$ P_2 = \left ( { \begin{array}{ccc}
1 & 0 & 0 \\
0 &-1 & 0 \\
0 & 0 & -1
\end{array} } \right ).$$
In order to see the action of the diffeomorphism of matrix $P_2$
on $\mathbb{S}^2$ we need to embed symplectically the ball $B(1)$ in the
sphere of radius $\frac{1}{2}$ of $\mathbb{R}^3$. Such a map can be given
for example by
$$
\begin{array}{cccc}
E_1:& B(1) & \longrightarrow & S^2 \\
& z=a+ib & \longmapsto &
\left({ \begin{array}{ccc} \sqrt{1-|z|^2} \; a\\
\sqrt{1-|z|^2} \; b\\
\frac{1}{2} - |z|^2
\end{array}} \right).
\end{array}
$$
The image by $(P_1, P_2)$ of $(E_1,E_1)(Q)$
in $\mathbb{S}^2 \times \mathbb{S}^2$ is now invariant by the action~$\rho_{\BC}$, but is
not the diagonal $\Sigma$ (which was to be expected as the cylinder has an area equal to $4$):
$$\widetilde{Q}=(P_1,P_2)((E_1,E_1)(Q))= \left \{ (X,Y) \in \mathbb{S}^2 \times \mathbb{S}^2 \right \}$$
with $X$ parametrised by
$$\left ( {
\begin{array}{c}
\sqrt{\frac{1}{2}+\tau-\frac{1}{2} \tau^2} \left(\frac{1}{\sqrt{2}} - \frac{1}{\sqrt{2}} \tau \right) \cos (\theta + \pi/4)\\
\sqrt{\frac{1}{2}+\tau-\frac{1}{2} \tau^2} \left(\frac{1}{\sqrt{2}} - \frac{1}{\sqrt{2}} \tau \right) \sin (\theta + \pi/4)\\
\tau-\frac{1}{2} \tau^2
\end{array}
} \right)$$
and $Y$ parametrised by
$$\left ( {
\begin{array}{c}
\sqrt{\frac{1}{2}-\tau-\frac{1}{2} \tau^2} \left(\frac{1}{\sqrt{2}} + \frac{1}{\sqrt{2}} \tau \right) \cos (\theta + \pi/4)\\
\sqrt{\frac{1}{2}-\tau-\frac{1}{2} \tau^2} \left(\frac{1}{\sqrt{2}} + \frac{1}{\sqrt{2}} \tau \right) \sin (\theta + \pi/4)\\
\tau + \frac{1}{2} \tau^2
\end{array}
} \right)
$$
for $s \in [0, 2 \pi]$ and $\tau \in (-1,1)$.
Now note that the image of the zero section (the curve $\tau = 0$)
of the cylinder is an equator of the diagonal sphere.
Moreover, the tangent bundle of~$\widetilde{Q}$ along
the equator is equal to the tangent bundle of the sphere~$\Sigma$.
This means that the original curve~$\gamma$ which was sitting
in the normal bundle of~$Q$ is, after these Hamiltonian
isotopies, a curve on which the integral of the Liouville form
is equal to~$\frac{1}{2}$ sitting in the normal bundle of~$\Sigma$
along the equator. It is lying in the disc bundle of area one as it
does not intersect the anti-diagonal~$\Delta$ after the Hamiltonian
diffeomorphism.
One finishes the proof using Lemma~\ref{lem:lemmePolterovich} as in
the case of~$\mathbb{C} \mathbb{P}^2$.
\end{proof}
\bibliographystyle{plain}
|
\section{Introduction}
Pregeometries arise in model theory via the operation of algebraic closure on a strongly minimal set, or more generally, via forking on the realisations of a regular type (see \cite{AP2}, for example). In the 1980's it was conjectured by Zilber that any strongly minimal structure is geometrically equivalent to one of the `classical' strongly minimal structures in the appropriate language: a pure set, a vector space over a fixed division ring, or an algebraically closed field. Here, `geometrically equivalent' means that the (pre)geometries are isomorphic, possibly after adding a small set of parameters.
This conjecture was refuted by Hrushovski in \cite{EH}. Working with a language $L_3$ having a single 3-ary relation symbol $R$, Hrushovski constructs continuum-many non-isomorphic (countable, saturated) strongly minimal structures $D_\mu$, where $\mu$ is a function which controls the multiplicities of certain algebraic types. These are not geometrically equivalent to the classical strongly minimal structures: their geometries are not disintegrated as in the pure set case, nor do they embed a group configuration as in the other cases. Hrushovski refers to these structures as `flat' and asks whether there is more than one geometric equivalence type of flat strongly minimal structure; in particular, whether the $D_\mu$ are geometrically equivalent (cf. Section 5.2 of \cite{EH}). Questions of this form are reiterated in \cite{Hasson}. Hrushovski's question is the subject of this paper and its sequel \cite{MF}. Before describing our results it will be helpful to remind the reader of some of aspects of Hrushovski's construction, and to fix some notation.
In \cite{EH} the \textit{predimension} of a finite $L_3$-structure is defined to be the size of $A$ minus the number of triples from $A$ which are related by $R$. Of interest is the class $\mathcal{C}_3$ of finite structures where this is non-negative on all substructures. Associated to this is a notion $\leq$ of \textit{self-sufficient} embedding, and a \textit{dimension} $d$ which gives rise to a pregeometry on $A$ (all of this is reviewed in more detail and generality in Section \ref{sec2} here). The class $(\mathcal{C}_3, \leq)$ is an amalgamation class with respect to the distinguished embeddings, and so there is an associated generic structure $\mathcal{M}_3$ (-- sometimes called a \textit{Fra\"{\i}ss\'e limit}) which also inherits a dimension function $d$. It can be shown that $\mathcal{M}_3$ is $\omega$-stable of Morley rank $\omega$ and that the pregeometry $PG(\mathcal{M}_3)$ given by $d$ is exactly the geometry of forking given by the (unique) regular type of rank $\omega$. Similar statements hold if the language is replaced by a language $L_n$ with a single $n$-ary relation symbol, for $n \geq 3$, and we denote the corresponding generic structure by $\mathcal{M}_n$.
What we have just described is often referred to as the `\textit{uncollapsed} case' of Hrushovski's construction. The strongly minimal $D_\mu$ can be seen as self-sufficient homogeneous substructures of $\mathcal{M}_3$ in which certain rank-1 types of $\mathcal{M}_3$ have only finitely many realizations. Thus these rank-1 types are `collapsed' to algebraic types in $D_\mu$. The strongly minimal structure $D_\mu$ is also constructed as the generic structure of an amalgamation class $(\mathcal{C}_\mu, \leq)$ where $\mathcal{C}_\mu$ is a subclass of $\mathcal{C}_3$ in which the multiplicities of certain quantifier-free types (minimally simply algebraic extensions) are bounded by the function $\mu$.
We can now describe the structure and main results of this paper. Section 2 contains background material on the (uncollapsed) construction. In Section 3 we find embeddings between the pregeometries of the $\mathcal{M}_n$ and show that these pregeometries cannot be distinguished by finite subgeometries (Corollary \ref{cor39}). However, we show in Section 4 that if $m \neq n$, then the (pre)geometries of $\mathcal{M}_n$ and $\mathcal{M}_m$ are \textit{not} isomorphic (Theorem \ref{T527}). In Section 5 we examine the effect of localization on the pregeometies $PG(\mathcal{M}_n)$ and show (Theorem \ref{T532}) that the geometry after localizing over a finite subset is isomorphic to the original geometry. Combining this with the results of Section 3, we deduce that if $m \neq n$, then the (pre)geometries of $\mathcal{M}_n$ and $\mathcal{M}_m$ are not \textit{locally} isomorphic. The main result of Section 6 is to show that the pregeometry $PG(\mathcal{M}_n)$ can be viewed as the generic structure of an amalgamation class $(P_n, \unlhd_n)$ of pregeometries (Theorem \ref{T543}).
The main ingredient in the proof of these results is a series of `Changing Lemmas.' Typically, these describe the effect (on the pregeometry) of replacing part of a structure in some $\mathcal{C}_n$ by another structure in $\mathcal{C}_n$ with the same domain. Related results are used in \cite{DE1} and \cite{DE2}.
In the sequel \cite{MF} (and in \cite{MFThesis}) we show that the pregeometries of the strongly minimal sets $D_\mu$ are all isomorphic to the pregeometry of $\mathcal{M}_3$. Moreover, the variation on the strongly minimal set construction given in 5.2 of \cite{EH} produces pregeometries locally isomorphic to the pregeometry of $\mathcal{M}_3$. Parallel results hold if the construction is done with an $n$-ary relation in place of a $3$-ary relation (and more general languages), and it appears that the distinct geometric equivalence types of the flat countable saturated strongly minimal structures produced by the construction in \cite{EH} are given by the pregeometries $PG(\mathcal{M}_n)$ (for $n\geq 3$).
\medskip
\noindent\textit{Acknowledgement:\/} Most of the results of this paper were produced whilst the second Author was supported as an Early Stage Researcher by the Marie Curie Research Training Network MODNET, funded by grant MRTN-CT-2004-512234 MODNET from the CEC.
\section{The construction and notation}\label{sec2}
We start with a brief description of the objects of our study, following Hrushovski \cite{EH} and Wagner \cite{FW}. The article \cite{B&S} of Baldwin and Shi gives another axiomatization and generalization of this method. The book \cite{AP2} of Pillay contains all necessary background material on pregeometries and model theory.
\smallskip
We work in slightly more generality than that outlined in the Introduction.
\begin{defn}\rm\label{fnotation}
Let $I$ be a countable set and $f:I\to\mathbbm N\setminus\{0\}\times\mathbbm N\setminus\{0\}$ be a function. Write $f(i)=(n_i,\alpha_i)$. Let $L_f=\{R_i:i\in I\}$ be a language where $R_i$ is an $n_i$-ary relational symbol.
If $A$ is a finite $L_f$-structure, we define its \textit{predimension} $\delta_f(A)$ to be
$$\delta_f(A)=|A|-\sum_{i\in I}\alpha_i|R_i^A|$$
(or $-\infty$), where $R_i^A$ is the set of $n_i$-tuples from $A$ which satisfy the relation $R_i$. We let
$\mathcal C_f$ be the set of finite $L_f$-structures $A$ such that $\delta_f(A') \geq 0$ for all $A' \subseteq A$.
Suppose $A \subseteq B \in \mathcal{C}_f$. We write $A \leq B$ and say that $A$ is \textit{self-sufficient} in $B$ (or \textit{strong} in $B$) if for all $B'$ with $A \subseteq B' \subseteq B$ we have $\delta_f(A) \leq \delta_f(B')$. In particular, $\emptyset \leq B$ for $B \in \mathcal C_f$.
\end{defn}
A key property of $\delta_f$ is that it is \textit{submodular}: if $A, B \subseteq C \in \mathcal{C}_f$, then $\delta_f(A \cup B) \leq \delta_f(A) + \delta_f(B) - \delta_f(A \cap B)$. Using this, one shows that if $A \leq B \leq C$ then $A \leq C$, and if $A_1, A_2 \leq B$ then $A_1\cap A_2 \leq B$.
Let $\bar {\mathcal C_f}$ be the class of $L_f$-structures all of whose finite substructures lie in $\mathcal{C}_f$. We can extend the notion of self-sufficiency to this class in a natural way, and for $B \subseteq C \in \bar {\mathcal C_f}$ we have $B \leq C$ if and only if whenever $A \leq B$ is finite, then $A \leq C$.
Note that if $A \subseteq B \in \bar {\mathcal C_f}$ and $A$ is finite then there is a finite $A'$ with $A \subseteq A' \subseteq B$ and $\delta_f(A')$ as small as possible. In this case $A' \leq B$. It follows that there is a smallest finite set $C \leq B$ with $A \subseteq C$. We denote this by ${\rm cl}_B(A)$ and refer to it as the \textit{self-sufficient closure} of $A$ in $B$. Thus the self-sufficient closure (in some given structure) of a finite set is finite, and we can extend the definition to arbitrary subsets of a structure in $\bar{\mathcal{C}}_f$. It is easy to see that ${\rm cl}_B$ is indeed a closure operation on $B$. However it does not usually satisfy the exchange condition and it is \textit{not} the closure operation which produces our pregeometries. The latter is what is given in the following. Note that we write $\delta$ rather than $\delta_f$ where no confusion will arise.
\begin{defn} \rm
Let $\mathcal M\in\bar {\mathcal C_f}$ and $A$ be a finite subset of $\mathcal{M}$. We define the \textit{dimension} $d_{\mathcal M}(A)$ of $A$ (in $\mathcal{M}$) to be the minimum value of $\delta(A')$ for all finite subsets $A'$ of $\mathcal{M}$ which contain $A$.
We define the $d$-\textit{closure} of $A$ in $\mathcal{M}$ to be:
$${\rm cl}^{d_{\mathcal M}}(A)=\{c\in\mathcal M:d_{\mathcal M}(Ac)=d_{\mathcal M}(A)\}.$$
We can coherently extend the definition of $d$-closure to infinite subsets $A$ of $\mathcal M$ by saying that the $d$-closure of $A$ is the union of the $d$-closures of finite subsets of $A$.
\end{defn}
Note that the $d$-closure of a finite set does not need to be finite. Concerning the relation with the self-sufficient closure, we always have ${\rm cl}_{\mathcal M}(A)\subseteq {\rm cl}^{d_{\mathcal M}}(A)$. If $A$ is finite then $d_{\mathcal M}(A)=\delta({\rm cl}_{\mathcal M}(A))$. Also if $A$ is finite we have $A\leq\mathcal M\text{ if and only if }d_{\mathcal M}(A)=\delta(A)$.
\begin{thm}
Let $\mathcal M\in\bar {\mathcal C_f}$. Then $(\mathcal M,{\rm cl}^{d_{\mathcal M}})$ is a pregeometry. Moreover, the dimension function (as cardinality of a basis) equals $d_{\mathcal M}$ on finite subsets of $\mathcal M$. We may use the notation $PG(\mathcal M)$ instead of $(\mathcal M,{\rm cl}^{d_{\mathcal M}})$.
\end{thm}
The class $(\mathcal C_f,\leq)$ has the following (strong) $\leq$-\textit{free amalgamation property}.
\begin{prop} Let $A_0\subseteq A_1\in\mathcal C_f$ and $A_0\leq A_2\in\mathcal C_f$ and $A_1\cap A_2=A_0$. Let $F = A_1\amalg_{A_0}A_2$ be the structure with underlying set $A_1\cup A_2$ such that the only relations are the ones arising from $A_1$ and $A_2$. Then $F\in\mathcal C_f$ and $A_1\leq F.$
\end{prop}
\begin{defn}\rm Let $\mathcal M$ be an $L_f$-structure. We say that $\mathcal M$ is a \textit{generic structure for} $(\mathcal C_f,\leq)$ if it is countable and satisfies the following conditions:
\begin{enumerate}
\item [(F1)] $\mathcal M\in\bar{\mathcal C_f}$
\item [(F2)] (extension property) If $A\leq\mathcal M$ and $A\leq B\in\mathcal C_f$ then there exists an embedding $g:B\to\mathcal M$ such that $g_{|A}=Id_{|A}$ (where $Id_{|A}$ is the identity map) and such that $g(B)\leq\mathcal M$.
\end{enumerate}
\end{defn}
Here we could also replace (F1) by
\begin{enumerate}
\item[(F1$'$)] $\mathcal{M}$ is the union of a chain $A_0 \leq A_1 \leq A_2 \leq \cdots$ of structures in $\mathcal{C}_f$.
\end{enumerate}
A standard argument shows:
\begin{thm}
There is a generic model $\mathcal M_{\mathcal C_f}$ for $(\mathcal C_f,\leq)$. It is unique up to isomorphism and if $h : A_1 \to A_2$ is an isomorphism beteween finite self-sufficient substructures, then $h$ extends to an automorphism of $\mathcal M_{\mathcal C_f}$. We may use the notation $\mathcal M_f$ instead of $\mathcal M_{\mathcal C_f}$.
\end{thm}
\begin{prop}
\label{P514}
With the above notation, If $L_f$ is finite then $\mathcal M_f$ is saturated and $Th(\mathcal M_f)$ is $\omega$-stable.
\end{prop}
\begin{rem}\rm
There are examples of $f$ where $L_f$ is not finite and such that $\mathcal M_f$ is saturated and $Th(\mathcal M_f)$ is $\omega$-stable, for example when $\mathcal C_f$ has only finitely many isomorphism types of each size. However we believe that we do get saturation and $\omega$-stability for every choice of $f$ because $f$ is integer valued. But this would require further analysis.
\end{rem}
Among these examples we would like to distinguish the following ones. The first case is as in the Introduction:
\begin{nota}\rm
\begin{enumerate}
\item[(i)] Let $n \geq 3$ and $I$ consist of a singleton. The language $L_n$ consists of a single $n$-ary relation symbol $R$ and the predimension is given by $\delta_n(A) = \vert A \vert - \vert R^A\vert$. In this case we denote the class by $\mathcal{C}_n$ and the generic model by $\mathcal{M}_n$.
\item[(ii)] Let $I=\mathbbm N\setminus\{0\}$ and $f$ a function defined by $f(i)=(i,1)$ for each $i\in I$. In other words we are considering the predimension given by: $$\delta_f(A)=|A|-\sum_{i\in\mathbbm N\setminus\{0\}}|R_i^A|,$$
where $R_i$ is an $i$-ary relation.
We write $L_{\omega}$, $\mathcal C_{\omega}$, $\delta_{\omega}$, $d_{\omega}$ and $\mathcal M_{\omega}$ instead of $L_f$, $\mathcal C_f$, $\delta_f$, $d_f$ and $\mathcal M_f$, for this particular function $f$.
\end{enumerate}
\end{nota}
It will be convenient to fix a first order language for the class of pregeometries. A reasonable choice for this is the language $LPI=\{I_n:n\in\mathbbm N\setminus\{0\}\}$ where each $I_n$ is an $n$-ary relational symbol. A pregeometry $(P,{\rm cl}^P)$ will be seen as a structure in this language by saying $I_n^P=\{\bar a\in P^n:\bar a\textit{ is independent in }P\}$. Notice that we can recover a pregeometry just by knowing its finite independent sets. In particular, a pregeometry is completely determined by the dimension function on its finite subsets. Note that the isomorphism type of a pregeometry is determined by the isomorphism type of its associated geometry and the size of the equivalence classes of interdependence. In the case where these are all countably infinite, it therefore makes no difference whether we consider the geometry or the pregeometry.
\section{Pregeometries from different predimensions}
The first result of this section shows that sometimes the pregeometry on the generic model associated to a particular predimension is isomorphic to the pregeometry on the generic model associated to a `simpler' predimension. This is motivated by the observation that the pregeometry associated to the predimension $\delta(A)=|A|-|R_3^A|-|R_4^A|$ is isomorphic to the simpler one given by $\delta(A)=|A|-|R_4^A|$. Actually, the proof of the following general result follows the same idea as the proof of this particular result, but is considerably more technical.
\begin{thm}
\label{T516}
Let $I$ be a countable set. Let $f:I\to\mathbbm N\setminus\{0\}\times\mathbbm N\setminus\{0\}$ be a function and $f(i)=(n_i,\alpha_i)$. Let $\sim$ be an equivalence relation on $I$ defined by $i\sim j$ if and only if $f(i)=f(j)$. Define a partially ordered set $(\widetilde I,\leq)$ by saying $[i]_\sim\leq[j]_\sim$ if and only if $n_i\leq n_j$ and $\alpha_j|\alpha_i$. Let $J\subseteq I$ be such that $\widetilde J$ is cofinal in $\widetilde I$ (so this means that for every $i \in I$ there exists $j \in J$ with $n_i \leq n_j$ and $\alpha_j \vert \alpha_i$). Then $$PG(\mathcal M_f)\simeq PG(\mathcal M_{f_{|J}}).$$
\end{thm}
\begin{proof} We start by outlining the idea of the proof. We construct a countable structure $\mathcal M_f^{\pi}$ in the restricted language $L_{f_{|J}}$ with the same underlying set as $\mathcal M_f$ in such a way that $PG(\mathcal M_f^{\pi})=PG(\mathcal M_f)$ and such that $\mathcal M_f^{\pi}$ is isomorphic to $\mathcal M_{f_{|J}}$. We construct $\mathcal M_f^{\pi}$ in such a way that for any finite subset $A$ the predimension value remains the same as the predimension value of $A$ calculated in $\mathcal M_f$, in particular the pregeometry remains unchanged.
To achieve this we replace each tuple in $R_i^{\mathcal M_f}$, where $i\in I\setminus J$, by some new $R_j$-tuples, where $j\in J$, and the number of replacement tuples for such a given tuple is exactly $\alpha_i/\alpha_j$, to compensate for the diferent weights. We do this replacement process in steps: first we need to decide which $j\in J$ to use for a given $i\in I\setminus J$, then in each step we fix $j$ and a finite subset $A$ of $\mathcal M_f$ then for each such $i$ corresponding to $j$, we replace each $R_i$-tuple with underlying set $A$ by $\alpha_i/\alpha_j$ new $R_j$-tuples, each one of them with underlying set $A$. An adequate $j$ will be such that $n_j\geq n_i$ and $\alpha_j\vert\alpha_i$. Moreover we need to check that we have enough new replacement tuples to choose from: basically permutation of coordinates will produce enough such tuples, together with the fact that $\emptyset\leq\mathcal M_f$, which assures that we do not have too many `old' tuples to start with. The details are as follows.
Let $h:I\to J$ be a function such that $[i]_\sim\leq[h(i)]_\sim$ and such that $j\in J\Rightarrow h(j)=j$. Such a function exists because $\widetilde J$ is cofinal in $\widetilde I$.
Let $g:\bigcup_{n=1}^{\infty}\mathcal M_f^n\to\mathbbm N\setminus\{0\}$ be defined by $g(a_1,\cdots,a_n)=|\{a_1,\cdots,a_n\}|$.
Observe that for each $A\subseteq\mathcal M_f$ and $n,m\in\mathbbm N\setminus \{0\}$, if $A^n\cap g^{-1}(m)$ is nonempty then, $$|A^n\cap g^{-1}(m)|\geq m!\geq m$$ because we can permute the $m$ distinct coordinates of an $n$-tuple in $A^n\cap g^{-1}(m)$ obtaining $m!$ tuples.
Now we fix $j\in J$. Let $A\subseteq\mathcal M_f$ be such that $|A|\leq n_j$. Notice that $R_i^A\cap g^{-1}(|A|)$ is exactly the set of $R_i^{\mathcal M_f}$ tuples with underlying set $A$. Our aim is to replace each tuple $\bar a\in R_i^A\cap g^{-1}(|A|)$ for all $i\in h^{-1}(j)\setminus\{j\}$ by one or more tuples in $(A^{n_j}\setminus R_j^A)\cap g^{-1}(|A|)$. We need to prove that we have enough space for this. If $\delta_f(A)\geq 0$ then $\sum_{i\in I}\alpha_i|R_i^A|\leq|A|$, hence $\sum_{i\in h^{-1}(j)}\alpha_i|R_i^A|\leq|A|$ and so $\sum_{i\in h^{-1}(j)\setminus\{j\}}\alpha_i|R_i^A|\leq|A|-\alpha_j|R_j^A|$. But then we have
\begin{enumerate}
\item []$\sum_{i\in h^{-1}(j)\setminus\{j\}}\alpha_i|R_i^A\cap g^{-1}(|A|)|\leq|A|-\alpha_j|R_j^A|$
\item []$\leq|A^{n_j}\cap g^{-1}(|A|)|-\alpha_j|R_j^A|$
\item []$\leq\alpha_j(|A^{n_j}\cap g^{-1}(|A|)|-|R_j^A\cap g^{-1}(|A|)|)$
\item []$=\alpha_j|(A^{n_j}\setminus R_j^A)\cap g^{-1}(|A|)|.$
\end{enumerate}
Thus we have $$\sum_{i\in h^{-1}(j)\setminus\{j\}}\frac{\alpha_i}{\alpha_j}|R_i^A\cap g^{-1}(|A|)|\leq|(A^{n_j}\setminus R_j^A)\cap g^{-1}(|A|)|.$$
Observe that the first term in the above inequality is the number of replacement tuples that are needed for this step and the second term is the number of available new tuples to use as replacements. So we do have space, that is, for each $i\in h^{-1}(j)\setminus\{j\}$ and $\bar a\in R_i^A\cap g^{-1}(|A|)$ we can replace $\bar a$ by $\frac{\alpha_i}{\alpha_j}$ distinct tuples in $(A^{n_j}\setminus R_j^A)\cap g^{-1}(|A|)$.
So now, if $i \in I\setminus J$ and $j = h(i)$ we can replace an $R_i$-tuple $\bar a \in R_i^{\mathcal{M}_f}$ by $\alpha_i/\alpha_j$ $R_j$-tuples $\bar{a}' \in \mathcal{M}_f^{n_j} \setminus R_j^{\mathcal{M}_f}$ with the same underlying set as $\bar{a}$. Doing this does not change the predimension of any subset of $\mathcal{M}_f$. The above calculation shows that we can do this (in an inductive fashion) for all steps corresponding to a choice of $j\in J$ and underlying set $A$, replacing all $\bar{a} \in R_i^{\mathcal{M}_f}$ for all $i\in I\setminus J$. We obtain a structure $\mathcal M_f^\pi$ in the restricted language $L_{f_{|J}}$ with the same underlying set as $\mathcal M_f$ and the same predimension function on finite subsets.
Now let $\pi:\mathcal M_f\to\mathcal M_f^{\pi}$ be the identity function. If $A\subseteq \mathcal M_f$ then $\pi(A)$ means the substructure of $\mathcal M_f^{\pi}$ with $A$ as the underlying set. Reciprocally if $A\subseteq \mathcal M_f^{\pi}$ then $\pi^{-1}(A)$ means the substructure of $\mathcal M_f$ with $A$ as the underlying set. The construction of $\mathcal M_f^{\pi}$ was made so that for each finite $A\subseteq\mathcal M_f$ we have: $$\delta_f(A)=\delta_{f_{|J}}(\pi(A)).$$
In other words, for each finite $A\subseteq\mathcal M_f^{\pi}$ we have: $$\delta_{f_{|J}}(A)=\delta_f(\pi^{-1}(A)).$$
In particular, for each finite $A\subseteq\mathcal M_f^{\pi}$ we have $\delta_{f_{|J}}(A)=\delta_f(\pi^{-1}(A))\geq 0$, thus $\mathcal M_f^{\pi}\in\bar {\mathcal C}_{f_{|J}}$. Also, because the predimension function is the same, so is the dimension function. Thus $$PG(\mathcal M_f^{\pi})\simeq PG(\mathcal M_f).$$
We now want to prove that $\mathcal M_f^{\pi}$ is the generic model for the class $\mathcal C_{f_{|J}}$, that is, isomorphic to $\mathcal M_{f_{|J}}$. It remains to show the extension property (the axiom $F2$). The diagram below will help us to follow the proof.
\[
\begin{xy}
\xymatrix{
& & & \mathcal M_f\ar[r]^{\pi} & \mathcal M_f^{\pi}\\
B\ar[r]^{l_1} & B'\ar[r]^{l_2} & l_2(B')\ar[r]^{\pi}\ar[ru]^{\leq} & \pi l_2(B')\ar[ru]^{\leq}\\
A\ar[r]\ar[u]^{\leq} & \pi^{-1}(A)\ar[r]\ar[u]^{\leq}\ar[ru]^{\leq} & A\ar[ru]^{\leq}\\
}
\end{xy}
\]
Let $A\leq B\in\mathcal C_{f_{|J}}\subseteq \mathcal C_f$ and $A\leq\mathcal M_f^{\pi}$. Let $B'$ be obtained from $B$ just by replacing $A$ by $\pi^{-1}(A)$ and nothing else. Let $l_1:B\to B'$ be the identity function. We still have $\pi^{-1}(A)\leq B'\in\mathcal C_f$ and $\pi^{-1}(A)\leq\mathcal M_f$. In fact we have that $A\leq B$ implies that $\pi^{-1}(A)\leq B'$ because the relations that we add when going from $\pi^{-1}(A)$ to $B'$ are the same we add when going from $A$ to $B$. In particular we get $\emptyset\leq\pi^{-1}(A)\leq B'$ so we may conclude that $B'\in\mathcal C_f$. Also $A\leq \mathcal M_f^{\pi}$ implies $\pi^{-1}(A)\leq\mathcal M_f$ because when going from $\mathcal M_f$ to $\mathcal M_f^{\pi}$ we do not change the predimension values of subsets.
So we have $\pi^{-1}(A)\leq B'\in\mathcal C_f$ and $\pi^{-1}(A)\leq\mathcal M_f$ and we can apply the extension property of $\mathcal M_f$. Let $l_2:B'\to\mathcal M_f$ be an embedding such that $l_2(B')\leq\mathcal M_f$ and $l_{2|\pi^{-1}(A)}=Id_{|\pi^{-1}(A)}$. Now we apply the $\pi$ function to the chain $\pi^{-1}(A)\leq l_2(B')\leq\mathcal M_f$ obtaining $A\leq\pi l_2 l_1(B)\leq\mathcal M_f^{\pi}$ and we are done.
It is not completely obvious that $\pi l_2l_1$ is an $L_{f_{|J}}$-embedding, so let us check that.
Let $j\in J$. We want to prove that $\bar a\in R_j^B$ if and only if $\pi l_2l_1(\bar a)\in R_j^{\mathcal M_f^{\pi}}$. For $\bar a\subseteq A$ this is clear. For $\bar a\nsubseteq A$ then we observe that:
\begin{enumerate}
\item [] $\bar a\in R_j^B\setminus R_j^A\Leftrightarrow l_1(\bar a)\in R_j^{B'}\setminus R_j^{\pi^{-1}(A)}$ (the changes are made inside $A$)
\item [] $\Leftrightarrow l_2l_1(\bar a)\in R_j^{\mathcal M_f}\setminus R_j^{\pi^{-1}(A)}$ ($l_2$ is an embedding)
\item [] $\Rightarrow\pi l_2l_1(\bar a)\in R_j^{\mathcal M_f^{\pi}}\setminus R_j^A$ (because $j\in J$).
\end{enumerate}
But the converse of the last implication also holds. This is because for $i\in I\setminus J$ we have by construction $R_i^B=\emptyset\Rightarrow R_i^{B'}\setminus R_i^{\pi^{-1}(A)}=\emptyset\Leftrightarrow R_i^{l_2(B')}\setminus R_i^{\pi^{-1}(A)}=\emptyset$ so when we apply $\pi$ no tuples are added from $R_j^{l_2(B')}\setminus R_j^{\pi^{-1}(A)}$ to $R_j^{\pi l_2(B')}\setminus R_j^A$. Thus $\pi l_2l_1$ is an embedding.
Finally, by the uniqueness of the generic model we have $\mathcal M_f^{\pi}\simeq\mathcal M_{f_{|J}}$. In particular $PG(\mathcal M_f^{\pi})\simeq PG(\mathcal M_{f_{|J}})$. But we already know that $PG(\mathcal M_f^{\pi})=PG(\mathcal M_f)$, thus we have $PG(\mathcal M_f)\simeq PG(\mathcal M_{f_{|J}})$.
\end{proof}
Note that the proof of this result depends on the fact that we are working with ordered tuples, however a similar result working with sets rather than tuples can be obtained (cf. Section 4.3 in \cite{MF}). Of course, we can also ask about the extent to which the hypotheses are necessary. We obtain the following consequences the previous result.
\begin{cor}
\label{C517}
Let $I$ be a countable set and $f:I\to\mathbbm N\setminus\{0\}\times\mathbbm N\setminus\{0\}$ be a function. Then $PG(\mathcal M_f)$ embeds in $PG(\mathcal M_{\omega})$. In other words, the pregeometry associated to the predimension $$\delta_f(A)=|A|-\sum_{i\in I}\alpha_i|R_i^A|$$ embeds into the pregeometry associated to the predimension $$\delta_{\omega}(A)=|A|-\sum_{n\geq 1}|R_n^A|.$$
\end{cor}
\begin{proof}
We extend $I$ to $\bar I=I\cup I_{\omega}$ with $I_{\omega}=\mathbbm N\setminus\{0\}$ (we can assume $I\cap I_{\omega}=\emptyset$). Then we extend $f$ to $\bar f:\bar I\to\mathbbm N\setminus\{0\}\times\mathbbm N\setminus\{0\}$ by saying $\bar f(i)=(i,1)$ for $i\in I_{\omega}$. Clearly we have that $\mathcal M_f$ embeds (as a closed substructure) in $\mathcal M_{\bar f}$. To be more precise $\mathcal M_f$ can be seen as a structure in $\bar{ \mathcal C}_{\bar f}$ if we interpret the extra relational symbols as the empty set, then we use the extension property of $\mathcal M_{\bar f}$ recursively because $\mathcal M_f$ is countable. Thus $PG(\mathcal M_f)$ embeds in $PG(\mathcal M_{\bar f})$. Now we observe that by Theorem \ref{T516} we have $PG(\mathcal M_{\bar f})\simeq PG(\mathcal M_{\bar f_{|I_{\omega}}})$. But $\mathcal M_{\bar f_{|I_{\omega}}}\simeq\mathcal M_{\omega}$, thus $PG(\mathcal M_{\bar f})\simeq PG(\mathcal M_{\omega})$ and $PG(\mathcal M_f)$ embeds in $PG(\mathcal M_{\omega})$.
\end{proof}
\begin{cor}
Consider the languages $L_{3,4}=\{R_3,R_4\}$ and $L_4=\{R_4\}$. Associated to the language $L_4$ we have the standard construction using the predimension $\delta_4(A)=|A|-|R_4^A|$ obtaining the class $\mathcal C_4$ and the generic $\mathcal M_4$. Associated to the language $L_{3,4}$ we proceed in the same manner as in the standard case using the predimension $\delta_{3,4}(A)=|A|-|R_3^A|-|R_4^A|$, obtaining the class $\mathcal C_{3,4}$ and the generic $\mathcal M_{3,4}$. Then, $$PG(\mathcal M_{3,4})\simeq PG(\mathcal M_4).$$
\end{cor}
\begin{proof}
In the notation of Theorem \ref{T516} we just observe that $J=\{4\}$ is cofinal in $I=\{3,4\}$.
\end{proof}
Now we want to prove further that $PG(\mathcal M_{\omega})$ embeds in $PG(\mathcal M_3)$. First we need the following lemma.
\begin{nota} \rm
Let $\mathcal M,\mathcal N$ be two $L$-structures and $A$ be a subset of both $\mathcal M$ and $\mathcal N$. We write $A[\mathcal M]$ to denote the substructure of $\mathcal M$ with underlying set $A$ and $A[\mathcal N]$ similarly.
\end{nota}
\begin{lem}
\label{L520}
Let $\mathcal M_{\omega(3)}$ be the generic model associated to the predimension $$\delta_{\omega(3)}(A)=|A|-\sum_{n\geq 3}|R_n^A|.$$ Then, $$PG(\mathcal M_{\omega(3)})\textit{ embeds in }PG(\mathcal M_3).$$
\end{lem}
\begin{proof}
Let $\mathcal M$ be an $L_{\omega(3)}$-structure and $\bar b\in T_4(\mathcal M):=\bigcup_{n\geq 4}R_n^{\mathcal M}$. We define an $L_{\omega(3)}$-structure $\mathcal M^{\bar b}$ obtained from $\mathcal M$ by replacing the tuple $\bar b=(b_1,\cdots,b_n)$ by the relations $$\{(b_1,x_1,b_2),(b_2,x_2,b_3),\cdots,(b_{n-1},x_{n-1},b_n)\}\cup\{(x_1,\cdots,x_{n-1})\}$$ where $\{x_1,\cdots,x_{n-1}\}$ are new distinct points specifically added for this task. So as a set $\mathcal M^{\bar b}=\mathcal M\cup\{x_1,\cdots,x_{n-1}\}$. We say that $\bar b'=(x_1,\cdots,x_{n-1})$ is the derivative of $\bar b$.
\begin{claim}
Let $\mathcal M$ be an $L_{\omega(3)}$-structure and $\bar b=(b_1,\cdots,b_n)\in T_4(\mathcal M)$. Let $A'$ be a finite subset of $\mathcal M^{\bar b}$ and let $\bar b'=(x_1,\cdots,x_{n-1})$ and $A'=A\cup X$ with $A=A'\cap\mathcal M$ and $X=A'\cap\{x_1,\cdots,x_{n-1}\}$. Then, $$\delta_{\omega(3)}(A')\geq\delta_{\omega(3)}(A[\mathcal M]).$$
In particular, if $\mathcal M\in\bar {\mathcal C}_{\omega(3)}$ then $\mathcal M^{\bar b}\in\bar {\mathcal C}_{\omega(3)}$.
\end{claim}
\begin{proof}
Suppose $X\neq\{x_1,\cdots,x_{n-1}\}$. Then, when going from $A[\mathcal M]$ to $A'$ we added $|X|$ points and added at most $|X|$ relations. This is because the relation $(x_1,\cdots,x_{n-1})$ is not added as $X\neq\{x_1,\cdots,x_{n-1}\}$. Thus, in this case $\delta_{\omega(3)}(A')\geq\delta_{\omega(3)}(A[\mathcal M])$.
Suppose that $X=\{x_1,\cdots,x_{n-1}\}$ and $\bar b\nsubseteq A$. Then, when going from $A[\mathcal M]$ to $A'$ we added $n-1$ points and at most $n-1$ relations because as $\bar b\nsubseteq A$, one of the relations $(b_k,x_k,b_{k+1})$ is not added. Thus in this case we have also $\delta_{\omega(3)}(A')\geq\delta_{\omega(3)}(A[\mathcal M])$.
Finally, suppose that $X=\{x_1,\cdots,x_{n-1}\}$ and $\bar b\subseteq A$. Then, when going from $A[\mathcal M]$ to $A'$ we added $n-1$ points and exactly $n$ relations, but we removed the relation $\bar b=(b_1,\cdots,b_n)$. Thus in this case $\delta_{\omega(3)}(A')=\delta_{\omega(3)}(A[\mathcal M])$.
Thus we conclude that $\delta_{\omega(3)}(A')\geq\delta_{\omega(3)}(A[\mathcal M])$ for all finite $A'\subseteq\mathcal M^{\bar b}$. In particular, if $\mathcal M\in\bar {\mathcal C}_{\omega(3)}$ then $\mathcal M^{\bar b}\in\bar {\mathcal C}_{\omega(3)}$.
\end{proof}
Now we need to prove another claim.
\begin{claim}
Let $\mathcal M\in\bar{\mathcal C}_{\omega(3)}$ and $\bar b\in T_4(\mathcal M)$. Then the inclusion $\mathcal M\to\mathcal M^{\bar b}$ gives an embedding of pregeometries (for example in the $LPI$ language).
\end{claim}
\begin{proof}
We are going to prove that for finite $A\subseteq\mathcal M$ we have $d_{\mathcal M}(A)=d_{\mathcal M^{\bar b}}(A)$.
Let $A\subseteq B'\subseteq\mathcal M^{\bar b}$ with $B'$ finite and $B'=B\cup X$ as in the previous claim. Then $A\subseteq B$ and by the previous claim we have $\delta_{\omega(3)}(B[\mathcal M])\leq\delta_{\omega(3)}(B')$. This shows that $d_{\mathcal M}(A)\leq d_{\mathcal M^{\bar b}}(A)$.
Conversely, let $A\subseteq B\subseteq\mathcal M$ with $B$ finite. If $\bar b\nsubseteq B$ then $\delta_{\omega(3)}(B[\mathcal M])=\delta_{\omega(3)}(B[\mathcal M^{\bar b}])$. If $\bar b\subseteq B$ then we put $B'=B[\mathcal M^{\bar b}]\cup\{x_1,\cdots,x_{n-1}\}$ where $\bar b'=(x_1,\cdots,x_{n-1})$. Then $\delta_{\omega(3)}(B[\mathcal M])=\delta_{\omega(3)}(B')$ as when going from $B[\mathcal M]$ to $B'$ we added $n-1$ points, $n$ relations and remove the relation $\bar b$.
The fact that we can find a substructure of $\mathcal M^{\bar b}$ containing $A$ with the $\delta_{\omega(3)}$ value equal to $\delta_{\omega(3)}(B[\mathcal M])$ shows that $d_{\mathcal M^{\bar b}}(A)\leq d_{\mathcal M}(A)$. Thus we have $d_{\mathcal M}(A)=d_{\mathcal M^{\bar b}}(A)$ and this proves that the inclusion $\mathcal M\to\mathcal M^{\bar b}$ is an embedding of pregeometries.
\end{proof}
Now we make a construction by recursion, of a sequence of $L_{\omega(3)}$-structures. We start by ordering the set of tuples $T:=T_4(\mathcal M_{\omega(3)})=\bigcup_{n\geq 4}R_n^{\mathcal M_{\omega(3)}}$.
Let $\mathcal M^{(0)}=\mathcal M_{\omega(3)}$. Let $\mathcal M^{(1)}=(\mathcal M^{(0)})^{\bar c_0}$ where $\bar c_0$ is the first element of $T$.
Suppose that we have constructed $\mathcal M^{(i+1)}=(\mathcal M^{(i)})^{\bar c_i}$. Now if $\bar c_i$ has $n$-arity greater or equal than $4$, then we set $\bar c_{i+1}=\bar c_i'$, otherwise we set $\bar c_{i+1}$ equal to the next unused element of $T$. Finally we put $\mathcal M^{(i+2)}=(\mathcal M^{(i+1)})^{\bar c_{i+1}}$.
We have constructed a sequence of $L_{\omega(3)}$-structures $\mathcal M^{(i)}$ with $i\in\omega$. Moreover, if we iterate this procedure $\omega$ steps we end up with a structure $\mathcal M^{(\omega)}$ only with $3$-tuples. By the previous claims $\mathcal M^{(i)}\in\bar{\mathcal C}_{\omega(3)}$ and for each $i\in\omega$ the inclusion function $\mathcal M^{(i)}\to\mathcal M^{(i+1)}$ gives an embedding of pregeometries. Finally we can define a pregeometry $P^{(\omega)}=\bigcup_{i\in\omega}PG(\mathcal M^{(i)})$. Let $d^{(\omega)}$ be the dimension function of $P^{(\omega)}$.
We need to prove the following claim.
\begin{claim}
Given a finite $A\subseteq\mathcal M^{(\omega)}$ then for all $i$ large enough we have $A[\mathcal M^{(i)}]=A[\mathcal M^{(\omega)}]$. In particular $\mathcal M^{(\omega)}\in\bar{\mathcal C}_3\subseteq\bar{\mathcal C}_{\omega(3)}$.
\end{claim}
\begin{proof}
Let $i_0$ be large enough so that $A\subseteq\mathcal M^{(i_0)}$. Now $A[\mathcal M^{(i_0)}]\in\mathcal C_{\omega(3)}$ so $|A|-\sum_{n\geq 4}|R_n^{A[\mathcal M^{(i_0)}]}|\geq 0$, in particular $\bigcup_{n\geq 4}R_n^{A[\mathcal M^{(i_0)}]}$ is finite. Thus, by our construction there is $i_1\geq i_0$ such that $A[\mathcal M^{(i_1)}]$ has only $3$-tuples. Now it is clear that for all $i\geq i_1$ we have $A[\mathcal M^{(i)}]=A[\mathcal M^{(\omega)}]$. In particular $\mathcal M^{(\omega)}\in\bar{\mathcal C}_3\subseteq\bar{\mathcal C}_{\omega(3)}$.
\end{proof}
We need to prove one last claim.
\begin{claim}
$P^{(\omega)}=PG(\mathcal M^{(\omega)})$.
\end{claim}
\begin{proof}
Let $A\subseteq\mathcal M^{(\omega)}$ and $A$ finite. We want to prove that $d^{(\omega)}(A)=d_{\mathcal M^{(\omega)}}(A)$. By our last claim there is $i$ such that ${\rm cl}_{\mathcal M^{(\omega)}}(A)[\mathcal M^{(\omega)}]={\rm cl}_{\mathcal M^{(\omega)}}(A)[\mathcal M^{(i)}]$. Then,
\begin{enumerate}
\item[] $d^{(\omega)}(A)=d_{\mathcal M^{(i)}}(A)$
\item[] $\leq d_{\mathcal M^{(i)}}({\rm cl}_{\mathcal M^{(\omega)}}(A))$
\item[] $\leq\delta_{\omega(3)}({\rm cl}_{\mathcal M^{(\omega)}}(A)[\mathcal M^{(i)}])$
\item[] $=\delta_{\omega(3)}({\rm cl}_{\mathcal M^{(\omega)}}(A)[\mathcal M^{(\omega)}])$
\item[] $=d_{\mathcal M^{(\omega)}}({\rm cl}_{\mathcal M^{(\omega)}}(A))$
\item[] $=d_{\mathcal M^{(\omega)}}(A)$
\end{enumerate}
Thus $d^{(\omega)}(A)\leq d_{\mathcal M^{(\omega)}}(A)$.
Now we want to prove the inequality in the other direction.
Let $B$ be a finite set with $A\subseteq B\subseteq\mathcal M^{(i)}$. Let $B^{(i)}=B$. Suppose that we have constructed $B^{(j)}$ for some $j\geq i$, then we obtain $B^{(j+1)}$ from $B^{(j)}$ in the same manner as we obtain $\mathcal M^{(j+1)}$ from $\mathcal M^{(j)}$. More precisely, if $\bar c_j\nsubseteq B^{(j)}$ then $B^{(j+1)}=B^{(j)}$, if $\bar c_j\subseteq B^{(j)}$ then we remove $\bar c_j=(b_1,\cdots,b_n)$ and add the relations $\{(b_1,x_1,b_2),\cdots,(b_{n-1},x_{n-1},b_n)\}\cup\{x_1,\cdots,x_{n-1})\}$ where $\bar c_j'=(x_1,\cdots,x_{n-1})$. Note that $\{x_1,\cdots,x_{n-1}\}$ are really new points in $B^{(j+1)}\setminus B^{(j)}$ as $\{x_1,\cdots,x_{n-1}\}\cap\mathcal M^{(j)}=\emptyset$.
We have constructed $B^{(j)}$ with $i\leq j\in\omega$. Clearly the sequence stabilizes for some $k$ for which $B^{(k)}$ has only $3$-tuples (as in the proof of last claim). Moreover, $B^{(k)}$ is a substructure of $\mathcal M^{(\omega)}$.
It is now easy to see that for each $j\geq i$ we have $$\delta_{\omega(3)}(B^{(j)}[\mathcal M^{(j)}])=\delta_{\omega(3)}(B^{(j+1)}[\mathcal M^{(j+1)}])$$ in particular we have, $$\delta_{\omega(3)}(B[\mathcal M^{(i)}])=\delta_{\omega(3)}(B^{(i)}[\mathcal M^{(i)}])=\delta_{\omega(3)}(B^{(k)}[\mathcal M^{(k)}])=\delta_{\omega(3)}(B^{(k)}[\mathcal M^{(\omega)}]).$$
The fact that we were able to find a finite substructure $B^{(k)}$ of $\mathcal M^{(\omega)}$ with $A\subseteq B^{(k)}\subseteq\mathcal M^{(\omega)}$ and $\delta_{\omega(3)}(B^{(k)}[\mathcal M^{(\omega)}])=\delta_{\omega(3)}(B[\mathcal M^{(i)}])$ proves that $d_{\mathcal M^{(\omega)}}(A)\leq d_{\mathcal M^{(i)}}(A)=d^{(\omega)}(A)$.
We have proved that $d^{(\omega)}(A)=d_{\mathcal M^{(\omega)}}(A)$ for all finite $A\subseteq\mathcal M^{(\omega)}$. Thus, $P^{(\omega)}=PG(\mathcal M^{(\omega)})$.
\end{proof}
Now we observe that as $\mathcal M^{(\omega)}\in\bar{\mathcal C}_3$, then there exists a strong embedding $g:\mathcal M^{(\omega)}\stackrel{\leq}{\longrightarrow}\mathcal M_3$, (that is, $g$ is an embedding with $g(\mathcal M^{(\omega)})\leq\mathcal M_3$). For this we use the extension property (axiom $F2$) of $\mathcal M_3$ recursively ($\mathcal M^{(\omega)}$ is countable). But because $g$ is a strong embedding then $g$ is also an embedding of pregeometries, so we have that $PG(\mathcal M^{(\omega)})$ embeds in $PG(\mathcal M_3)$.
Finally, we recall all we have proved: $PG(\mathcal M_{\omega(3)})$ is a subpregeometry of $P^{(\omega)}$, $P^{(\omega)}=PG(\mathcal M^{(\omega)})$ and $PG(\mathcal M^{(\omega)})$ embeds in $PG(\mathcal M_3)$. Thus $PG(\mathcal M_{\omega(3)})$ embeds in $PG(\mathcal M_3)$.
\end{proof}
\begin{rem} \rm
Note that in contrast to the other results of this section, the previous lemma also holds if we work with unordered tuples.
\end{rem}
\begin{lem}
\label{L522}
Let $n\geq 3$ be a natural number. Then, $$PG(\mathcal M_3)\textit{ embeds in }PG(\mathcal M_n).$$
\end{lem}
\begin{proof}
Let $\mathcal M_3^h$ be the $L_n$-structure obtained from $\mathcal M_3$ by replacing each relation $(a_1,a_2,a_3)\in R_3^{\mathcal M_3}$ by the relation $(a_1,a_2,a_3,a_4,\cdots,a_n)$ where $a_i=a_3$ for all $i\geq 4$. Then, clearly, $\mathcal M_3^h\in\bar{\mathcal C}_n$ and $PG(\mathcal M_3^h)=PG(\mathcal M_3)$, because the predimension function is the same in both structures. But as $\mathcal M_3^h\in\bar{\mathcal C}_n$ we can use recursively the extension property to build a strong embedding $\mathcal M_3^h\stackrel{\leq}{\longrightarrow}\mathcal M_n$, that is, with $g(\mathcal M_3^h)\leq\mathcal M_n$. But then $g$ is also an embedding of pregeometries, thus $PG(\mathcal M_3)=PG(\mathcal M_3^h)$ embeds in $PG(\mathcal M_n)$.
\end{proof}
We obtain the following theorem.
\begin{thm}
\label{T523}
Let $I$ be a countable set and $f:I\to\mathbbm N\setminus\{0\}\times\mathbbm N\setminus\{0\}$ be a function and $n\geq 3$ a natural number. Then there exists a chain of embeddings of pregeometries as in the following diagram: $$PG(\mathcal M_f)\to PG(\mathcal M_{\omega})\to PG(\mathcal M_3)\to PG(\mathcal M_n)\to PG(\mathcal M_{\omega}).$$
\end{thm}
\begin{proof}
By Corollary \ref{C517} we get $PG(\mathcal M_f)\to PG(\mathcal M_{\omega})$. By Theorem \ref{T516} we have $PG(\mathcal M_{\omega})\simeq PG(\mathcal M_{\omega(3)})$. By Lemma \ref{L520} we have that $PG(\mathcal M_{\omega(3)})$ embeds in $PG(\mathcal M_3)$. By Lemma \ref{L522} we get $PG(\mathcal M_3)\to PG(\mathcal M_n)$. Finally, $\mathcal M_n\in\bar{\mathcal C}_n\subseteq\bar{\mathcal C}_{\omega}$ and $\mathcal M_n$ is countable, so we can use the extension property recursively to build an strong embedding $g:\mathcal M_n\to\mathcal M_{\omega}$. Then $g:PG(\mathcal M_n)\to PG(\mathcal M_{\omega})$ is an embedding of pregeometries. This conclude the proof.
\end{proof}
\begin{cor}\label{cor39}
Let $m,n\geq 3$ be natural numbers. Then $PG(\mathcal M_m)$ and $PG(\mathcal M_n)$ embed in each other. In particular, in the $LPI$ language of pregeometries we have that $PG(\mathcal M_m)$ and $PG(\mathcal M_n)$ have the same isomorphism types of finite subpregeometries.
\end{cor}
\begin{proof}
We use Theorem \ref{T523} to build a chain of embeddings $$PG(\mathcal M_m)\to PG(\mathcal M_{\omega})\to PG(\mathcal M_3)\to PG(\mathcal M_n).$$
\end{proof}
\section{Pregeometries and different arities}
In the previous section we have seen that for natural numbers $n,m\geq 3$ the pregeometries $PG(\mathcal M_n)$ and $PG(\mathcal M_m)$ embed in each other. The main result of this section is that if $m \neq n$, then $PG(\mathcal M_n)$ and $PG(\mathcal M_m)$ are not isomorphic. In order to prove this we need two technical lemmas. The first of these lemmas is in fact almost trivial.
\begin{lem}[First Changing Lemma]
\label{L525}
Keep the notation of the previous sections. In particular let $I$ be a countable set and $f:I\to\mathbbm N\setminus\{0\}\times\mathbbm N\setminus\{0\}$. Suppose $\mathcal M\in\bar{\mathcal C}_f$, $A\leq\mathcal M$ and $A'\in\bar{\mathcal C}_f$ where $A'$ has the same underlying set as $A$. Let $\mathcal M'$ be the structure obtained from $\mathcal M$ by replacing $A$ by $A'$, where by this we mean for each $i\in I$ we have $R_i^{\mathcal M'}=(R_i^{\mathcal M}\setminus R_i^A)\cup R_i^{A'}$. Then
$$A'\leq\mathcal M'\in\bar{\mathcal C}_f.$$
\end{lem}
\begin{proof}
Let $X$ be a finite subset of $\mathcal M'$. We want to prove that $\delta(X[\mathcal M'])\geq 0$.
We have that $A\leq\mathcal M\Rightarrow A\cap X[\mathcal M]\leq X[\mathcal M]$. Thus $\delta(X[\mathcal M])\geq\delta(A\cap X[\mathcal M])$. Notice that the points and relations added when going from $A\cap X[\mathcal M]$ to $X[\mathcal M]$ are the same points and relations added when going from $A'\cap X[\mathcal M']$ to $X[\mathcal M']$. Thus we also have $\delta(X[\mathcal M'])\geq\delta(A'\cap X[\mathcal M'])$. But as $A'\cap X[\mathcal M']$ is a substructure of $A'\in\bar{\mathcal C}_f$, we have $\delta(A'\cap X[\mathcal M'])\geq 0$. Thus $\delta(X[\mathcal M'])\geq 0$. This proves that $\mathcal M'\in\bar{\mathcal C}_f$. To see that $A'\leq\mathcal M'$ we use a similar argument.
\end{proof}
\begin{lem}[Second Changing Lemma]
\label{L526}
With the above notation, suppose $\mathcal M\in\bar{\mathcal C}_f$, $A\leq\mathcal M$ and $A'\in\bar{\mathcal C}_f$ where $A'$ has the same underlying set as $A$. Let $\mathcal M'$ be the structure obtained from $\mathcal M$ by replacing $A$ by $A'$ as we have done in Lemma \ref{L525}. Then $A'\leq\mathcal M'\in\bar{\mathcal C}_f$ and
if $PG(A)=PG(A')$, then $PG(\mathcal M)=PG(\mathcal M').$
\end{lem}
\begin{proof}
Assume that $PG(A)=PG(A')$, that is, $d_A=d_{A'}$. We want to prove that $PG(\mathcal M)=PG(\mathcal M')$, that is, $d_{\mathcal M}=d_{\mathcal M'}$. The idea of this proof is to reconstruct in small steps $\mathcal M$ starting with $A$ and $\mathcal M'$ starting with $A'$, and at each one of this steps the dimension function will be the same in both of these parallel constructions (reconstructions). More precisely, in the first step we add the remaining points, and in each subsequent step we add one of the remaining relations.
Let $R:=\bigcup_{i\in I}R_i^{\mathcal M}\setminus R_i^A=\bigcup_{i\in I}R_i^{\mathcal M'}\setminus R_i^{A'}$.
Let $\mathcal M_0$ be the structure obtained from $\mathcal M$ by removing all the relations in $R$, that is, the relations contained in $\mathcal M$ not entirely contained in $A$. Similarly, let $\mathcal M_0'$ be obtained from $\mathcal M'$ by removing all the relations in $R$, that is, the relations contained in $\mathcal M'$ not entirely contained in $A'$.
Clearly we have $d_{\mathcal M_0}=d_{\mathcal M_0'}$. In fact let $B$ be a finite subset of both $\mathcal M$ and $\mathcal M'$ (which share the same underlying set) and let $X=B\setminus(A\cap B)$, then we have
\begin{enumerate}
\item [] $d_{\mathcal M_0}(B)=d_{\mathcal M_0}(A\cap B)+|X|$
\item [] $=d_{A}(A\cap B)+|X|$
\item [] $=d_{A'}(A\cap B)+|X|$
\item [] $=d_{\mathcal M_0'}(A\cap B)+|X|=d_{\mathcal M_0'}(B)$
\end{enumerate}
thus $d_{\mathcal M_0}=d_{\mathcal M_0'}$.
Now we add the relations that we have removed one by one in both $\mathcal M_0$ and $\mathcal M_0'$. Let $R=(r_i)_{i\in\kappa}$ be an enumeration of $R$, with $\kappa$ some cardinal.
For each $j\in\kappa$ let $\mathcal M_j$ be obtained from $\mathcal M_0$ by adding the relations $\{r_i:i\in j\}$ and $\mathcal M_j'$ be obtained from $\mathcal M_0'$ by adding the same set of relations $\{r_i:i\in j\}$. Clearly $\mathcal M_{\kappa}=\mathcal M$ and $\mathcal M_{\kappa}'=\mathcal M'$. We want to prove that $d_{\mathcal M_{\kappa}}=d_{\mathcal M_{\kappa}'}$. The proof is by transfinite induction.
Let $j=0$. In this case we already proved that $d_{\mathcal M_0}=d_{\mathcal M_0'}$.
Now we want to prove that $d_{\mathcal M_j}=d_{\mathcal M_j'}\Rightarrow d_{\mathcal M_{j+1}}=d_{\mathcal M_{j+1}'}$. This follows directly from the following claim:
\begin{claim}
Let $A_1,A_2\in\bar{\mathcal C}_f$ have the same underlying set and that $PG(A_1)=PG(A_2)$. Let $B_1$ and $B_2$ be the structures obtained by adding the same relation $r$ to $A_1$ and $A_2$. Assume further that $B_1,B_2\in\bar{\mathcal C}_f$. Then $PG(B_1)=PG(B_2)$.
\end{claim}
\begin{proof}
Let $X$ be a finite subset of $B_1=B_2$ (as sets). We know that $d_{A_1}(X)=d_{A_2}(X)$ and we want to prove that $d_{B_1}(X)=d_{B_2}(X)$.
Fix $i\in\{1,2\}$. We have either $$d_{B_i}(X)=d_{A_i}(X)\textit{ or }d_{B_i}(X)=d_{A_i}(X)-1$$ because we just add one relation.
Suppose we have $d_{B_i}(X)=d_{A_i}(X)-1$. Then there is a finite $Y$ with $X\subseteq Y\subseteq B_i$ such that $\delta(Y[A_i])=d_{A_i}(X)$ and $Y\supseteq r$. In particular $d_{A_i}(X)\leq d_{A_i}(Y)\leq\delta(Y[A_i])=d_{A_i}(X)$. Thus $d_{A_i}(Y)=d_{A_i}(X)$.
Conversely, if there exists a finite $Y$ with $X\subseteq Y\subseteq B_i$ such that $d_{A_i}(Y)=d_{A_i}(X)$ and $Y\supseteq r$, then there exists a finite $Z\supseteq Y\supseteq X$ such that $\delta(Z[A_i])=d_{A_i}(X)$. But as $Z\supseteq r$ we have $\delta(Z[B_i])=d_{A_i}(X)-1$, thus $d_{B_i}(X)\leq d_{B_i}(Y)\leq\delta(Z[B_i])=d_{A_i}(X)-1$ so we get $d_{B_i}(X)=d_{A_i}(X)-1$.
But now we have
\medskip
\begin{enumerate}
\item[] $d_{B_1}(X)=d_{A_1}(X)-1\Leftrightarrow$
\item[] $\Leftrightarrow\exists Y\supseteq X\cup r\text{ such that } d_{A_1}(X)=d_{A_1}(Y)$
\item[] $\Leftrightarrow\exists Y\supseteq X\cup r\text{ such that } d_{A_2}(X)=d_{A_2}(Y)$ (because $d_{A_1}=d_{A_2}$)
\item[] $\Leftrightarrow d_{B_2}(X)=d_{A_2}(X)-1$.
\end{enumerate}
But consequently, we also have $d_{B_1}(X)=d_{A_1}(X)$ if and only if $d_{B_2}(X)=d_{A_2}(X)$. Now as we have $d_{A_1}(X)=d_{A_2}(X)$, the previous argument proves that either way we have $d_{B_1}(X)=d_{B_2}(X)$. Thus $d_{B_1}=d_{B_2}$, that is, $PG(B_1)=PG(B_2)$.
\end{proof}
Now we return to our recursion argument considering the case when $j$ is a limit ordinal. Let $j$ be a limit ordinal. Assume that $d_{\mathcal M_i}=d_{\mathcal M_i'}$ for all $i\in j$. We want to prove that $d_{\mathcal M_j}=d_{\mathcal M_j'}$.
Let $X$ be a finite subset of $\mathcal M_j=\mathcal M_j'$ (as sets).
Let $i_0\in j$ be such that all the relations in ${\rm cl}_{\mathcal M_j}(X)$ are in $\mathcal M_{i_0}$, this is possible because ${\rm cl}_{\mathcal M_j}(X)$ contains only finitely many relations. Then,
$$d_{\mathcal M_{i_0}}(X)\leq d_{\mathcal M_{i_0}}({\rm cl}_{\mathcal M_j}(X))\leq\delta({\rm cl}_{\mathcal M_j}(X)[\mathcal M_{i_0}])=\delta({\rm cl}_{\mathcal M_j}(X)[\mathcal M_j])=d_{\mathcal M_j}(X)$$ thus $d_{\mathcal M_{i_0}}(X)=d_{\mathcal M_j}(X)$.
We have proved that for all $i\in j$ big enough (with $X$ fixed) we have $d_{\mathcal M_i}(X)=d_{\mathcal M_j}(X)$. Similarly, for all $i\in j$ big enough we have $d_{\mathcal M_i'}(X)=d_{\mathcal M_j'}(X)$. Let $i\in j$ be big enough for both cases, then using the inductive hypothesis we get
$$d_{\mathcal M_j}(X)=d_{\mathcal M_i}(X)=d_{\mathcal M_i'}(X)=d_{\mathcal M_j'}(X).$$
This means that we proved that $d_{\mathcal M_j}=d_{\mathcal M_j'}$ for all $j\in\kappa+1$. Finally we have $PG(\mathcal M)=PG(\mathcal M_{\kappa})=PG(\mathcal M_{\kappa}')=PG(\mathcal M')$, as desired.
\end{proof}
We now have the tools necessary to prove that the isomorphism types of pregeometries arising from different arities are different.
\begin{thm}
\label{T527}
Let $n,m\in\mathbbm N\setminus\{0\}$ with $n<m$. Then, $$PG(\mathcal M_n)\ncong PG(\mathcal M_m).$$
\end{thm}
\begin{rem}\rm In fact the proof we shall give shows that if $A \in \bar{\mathcal{C}}_m$ and $B \in \bar{\mathcal{C}}_n$ and there is $X \leq A$ consisting of $m$ points related by a single relation, then $PG(A)$ and $PG(B)$ are non-isomorphic.
\end{rem}
\begin{proof}
Suppose that there is an isomorphism of pregeometries from $PG(\mathcal M_n)$ to $PG(\mathcal M_m)$. We first observe that without loss of generality, we can assume that $\mathcal M_n$ and $\mathcal M_m$ have the same underlying set and that the isomorphism of pregeometries is the identity map. We can write $d$ instead of $d_{\mathcal M_n}$ and $d_{\mathcal M_m}$. Let ${\rm cl}^d$ be the closure operator of the pregeometry $PG(\mathcal M_n)=PG(\mathcal M_m)$.
Let $X=\{a_1,\cdots,a_m\}$ be a subset of $\mathcal M_m$ such that $\widetilde X:=X[\mathcal M_m]\leq\mathcal M_m$ and $R_m^{\widetilde X}=\{(a_1,\cdots,a_m)\}$ with $a_1,\cdots,a_m$ distinct. This is possible by the extension property of $\mathcal M_m$. Let $\widehat X:=X[\mathcal M_n]$.
Now consider $Y={\rm cl}_{\mathcal M_n}(X)$ and $\widehat Y:=Y[\mathcal M_n]$ and $\widetilde Y:=Y[\mathcal M_m]$. We have $$d(Y)=d_{\mathcal M_n}({\rm cl}_{\mathcal M_n}(X))=d_{\mathcal M_n}(X)=d(X)=d_{\mathcal M_m}(X)=\delta_m(X[\mathcal M_m])=m-1$$ thus $$d(Y)=d(X)=m-1.$$
Let $(a_{11},\cdots,a_{1n}),\cdots,(a_{k1},\cdots,a_{kn})$ be a list of the relations in $R_n^{\widehat Y}$. We have $\widehat Y={\rm cl}_{\mathcal M_n}(X)\leq\mathcal M_n$, thus $$d(Y)=\delta_n(\widehat Y)=|Y|-k.$$
Now, for each $1\leq i\leq k$ let $Z_i:={\rm cl}^d(\{a_{i1},\cdots,a_{in}\})$ and let $\widehat Z_i:=Z_i[\mathcal M_n]$ and $\widetilde Z_i:=Z_i[\mathcal M_m]$. Note that $Z_i$ can be infinite and that ${\rm cl}^d(Z_i)=Z_i$. Also we have $\widetilde Z_i\subseteq {\rm cl}_{\mathcal M_m}(\widetilde Z_i)\subseteq {\rm cl}^d(Z_i)$ and $\widehat Z_i\subseteq {\rm cl}_{\mathcal M_n}(\widehat Z_i)\subseteq {\rm cl}^d(Z_i)$. Thus we have $\widetilde Z_i={\rm cl}_{\mathcal M_m}(\widetilde Z_i)$ and $\widehat Z_i={\rm cl}_{\mathcal M_n}(\widehat Z_i)$, that is, $\widetilde Z_i\leq\mathcal M_m$ and $\widehat Z_i\leq\mathcal M_n$. In particular, the pregeometries $PG(\widetilde Z_i)$ and $PG(\widehat Z_i)$ are those induced from $PG(\mathcal M_m)=PG(\mathcal M_n)$ so we get $PG(\widetilde Z_i)=PG(\widehat Z_i)$.
Now observe that $Z_i\nsupseteq X$. In fact, if $X\subseteq Z_i$ then we would have $X\subseteq {\rm cl}^d(\{a_{i1},\cdots,a_{in}\})$, in particular $d(X)\leq d(\{a_{i1},\cdots,a_{in}\})$. But we have
\begin{enumerate}
\item[] $d(\{a_{i1},\cdots,a_{in}\})\leq\delta_n(\{a_{i1},\cdots,a_{in}\})[\mathcal M_n])$
\item[] $=|\{a_{i1},\cdots,a_{in}\}|-|R_n^{\{a_{i1},\cdots,a_{in}\}[\mathcal M_n]}|$
\item[] $\leq n-1<m-1=d(X).$
\end{enumerate}
Thus $Z_i\nsupseteq X$.
Let $Z_i^*\in\bar{\mathcal C}_m$ be the structure obtained from $\widehat Z_i$ by replacing each relation $(x_1,\cdots,x_n)\in R_n^{\widehat Z_i}$ by the relation $(x_1,\cdots,x_n,\cdots,x_n)$ with $m$ coordinates. Then we have $$PG(Z_i^*)=PG(\widehat Z_i)=PG(\widetilde Z_i).$$
Now we want to construct a sequence of $L_m$-structures $$\mathcal M_m=\mathcal M_m^{(0)},\cdots,\mathcal M_m^{(k)}$$ all with the same underlying set and with the following properties:
\begin{enumerate}
\item $\mathcal M_m^{(i)}\in\bar{\mathcal C}_m$
\item $PG(\mathcal M_m^{(i)})=PG(\mathcal M_m)$
\item $R_m^{\mathcal M_m^{(i)}}\supseteq\{(a_1,\cdots,a_m)\}\cup\{(a_{j1},\cdots,a_{jn},\cdots,a_{jn}):1\leq j\leq i\}.$
\end{enumerate}
Let $\mathcal M_m^{(0)}=\mathcal M_m$. Then for $i=0$ the properties are satisfied, note that $(a_1,\cdots,a_m)\in R_m^{\mathcal M_m}$. Now suppose we have constructed $\mathcal M_m^{(i)}$ satisfying the properties and $i<k$. Then we obtain $\mathcal M_m^{(i+1)}$ from $\mathcal M_m^{(i)}$ by replacing $Z_{i+1}[\mathcal M_m^{(i)}]$ by $Z_{i+1}^*$. We need to prove that $\mathcal M_m^{(i+1)}$ satisfies the properties.
Observe that $Z_{i+1}[\mathcal M_m^{(i)}]\leq\mathcal M_m^{(i)}$. In fact we have $PG(\mathcal M_m^{(i)})=PG(\mathcal M_m)$ so we have $Z_{i+1}\subseteq {\rm cl}_{\mathcal M_m^{(i)}}[Z_{i+1}]\subseteq {\rm cl}^d(Z_{i+1})$ and as $Z_{i+1}={\rm cl}^d(Z_{i+1})$ we get that $Z_{i+1}={\rm cl}_{\mathcal M_m^{(i)}}(Z_{i+1})$, that is, $Z_{i+1}[\mathcal M_m^{(i)}]\leq\mathcal M_m^{(i)}$.
Now we apply the First Changing Lemma: $Z_{i+1}^*\in\bar {\mathcal C}_m$ and $Z_{i+1}[\mathcal M_m^{(i)}]\leq\mathcal M_m^{(i)}$ so $Z_{i+1}^*\leq\mathcal M_m^{(i+1)}\in\bar{\mathcal C}_m$, proving property $1)$.
For property $2)$ we use the Second Changing Lemma: all we need to prove is that $PG(Z_{i+1}^*)=PG(Z_{i+1}[\mathcal M_m^{(i)}])$. But in fact, $$d_{Z_{i+1}[\mathcal M_m^{(i)}]}=(d_{\mathcal M_m^{(i)}})_{|Z_{i+1}}=(d_{\mathcal M_m})_{|Z_{i+1}}=d_{Z_{i+1}[\mathcal M_m]}=d_{\widetilde Z_{i+1}}$$ thus $$PG(Z_{i+1}[\mathcal M_m^{(i)}])=PG(\widetilde Z_{i+1})=PG(\widehat Z_{i+1})=PG(Z_{i+1}^*).$$ Now we apply the Second Changing Lemma and we get $PG(\mathcal M_m^{(i+1)})=PG(\mathcal M_m^{(i)})=PG(\mathcal M_m)$, proving property $2)$.
Finally we need to prove property $3)$. First we observe that when going from $\mathcal M_m^{(i)}$ to $\mathcal M_m^{(i+1)}$ we add the relation $(a_{j1},\cdots,a_{jn},\cdots,a_{jn})$ for $j=i+1$, as this relation is in $R_m^{Z_{i+1}^*}$.
Observe also that $(a_1,\cdots,a_m)$ is not removed because $\{a_1,\cdots,a_m\}=X\nsubseteq Z_{i+1}$.
Now we need to see that for $j\leq i$ the relation $(a_{j1},\cdots,a_{jn},\cdots,a_{jn})$ is not removed when going from $\mathcal M_m^{(i)}$ to $\mathcal M_m^{(i+1)}$. If $\{a_{j1},\cdots,a_{jn}\}\nsubseteq Z_{i+1}$ this is clear. If $\{a_{j1},\cdots,a_{jn}\}\subseteq Z_{i+1}$ then
\begin{enumerate}
\item[] $(a_{j1},\cdots,a_{jn})\in R_n^{\mathcal M_n}\Leftrightarrow (a_{j1},\cdots,a_{jn})\in R_n^{Z_{i+1}[\mathcal M_n]}=R_n^{\widehat Z_{i+1}}$
\item[] $\Leftrightarrow (a_{j1},\cdots,a_{jn},\cdots,a_{jn})\in R_m^{Z_{i+1}^*}$
\end{enumerate}
thus $(a_{j1},\cdots,a_{jn},\cdots,a_{jn})$ is not removed. This proves that $\mathcal M_m^{(i+1)}$ satisfies property $3)$.
Finally we get the desired contradiction. $\mathcal M_m^{(k)}$ satisfies properties $2)$ and $3)$, that is, $PG(\mathcal M_m^{(k)})=PG(\mathcal M_m)$ and $$R_m^{Y[\mathcal M_m^{(k)}]}\supseteq\{(a_1,\cdots,a_m)\}\cup\{(a_{j1},\cdots,a_{jn},\cdots,a_{jn}):1\leq j\leq k\}.$$ Thus we can make the following calculation (remember that $d(Y)=|Y|-k$): $$d(Y)\leq\delta_m(Y[\mathcal M_m^{(k)}])=|Y|-|R_m^{Y[\mathcal M_m^{(k)}]}|\leq |Y|-(k+1)=(|Y|-k)-1=d(Y)-1.$$ Thus we get the desired contradiction $d(Y)\leq d(Y)-1$. We may then conclude that $PG(\mathcal M_n)\ncong PG(\mathcal M_m)$.
\end{proof}
\section{The local isomorphism type}
We have seen in the previous section that $PG(\mathcal M_n)$ is not isomorphic to $PG(\mathcal M_m)$. Here we show that $PG(\mathcal M_n)$ and $PG(\mathcal M_m)$ are not even locally isomorphic. First we need some more changing lemmas.
\begin{lem} (Third Changing Lemma)
\label{L528}
Let $f$ be as in Lemma \ref{L525} and let $\mathcal M_f$ be the generic model for $(\mathcal C_f,\leq)$. Let $Z\leq\mathcal M_f$ with $Z$ finite. Let $Z'\in\mathcal C_f$ be a structure with the same underlying set as $Z$ and $\mathcal M_f'$ be obtained from $\mathcal M_f$ by replacing $Z$ by $Z'$, as we have done in Lemma \ref{L525}. Then we have $Z'\leq\mathcal M_f'\in\bar{\mathcal C}_f$ and
$$\mathcal M_f'\simeq\mathcal M_f.$$
\end{lem}
\def\mathcal{M}{\mathcal{M}}
\begin{proof}
The only thing we have to show is that $\mathcal M_f'$ satisfies the extension property. For any set $X\subseteq\mathcal M_f$ we write $X$ instead of $X[\mathcal M_f]$ and $X'$ instead of $X[\mathcal M_f']$. Let $A\subseteq\mathcal M_f$ with $A'\leq\mathcal M_f'$ and $A'\leq B'\in\mathcal C_f$ ($B'$ arbitrary). We want to prove that there exists an embedding $g':B'\to\mathcal M_f'$ fixing the elements of $A'$ and with $g'(B')\leq\mathcal M_f'$.
Let $W' = {\rm cl}_{\mathcal{M}_f'}(Z'A')$ and let $W$ be the corresponding structure in $\mathcal{M}_f$. Note that because $W' \leq \mathcal{M}_f'$ and $Z' \subseteq W'$ we have $W \leq \mathcal{M}_f$. Let $C'$ be the free amalgam of $W'$ and $B'$ over $A'$. Then $W', B' \leq C'$ and so we also have $Z', A' \leq C'$. Let $C$ be the result of changing $Z'$ to $Z$ in $C'$. Then we have $Z \leq C$ and (as above) $W \leq C$. We have $A \subseteq C$, but $A$ is not necessarily self-sufficient in $C$.
By the extension property for $\mathcal{M}_f$ there is an embedding $g : C \to \mathcal{M}_f$ with $g(C) \leq \mathcal{M}_f$ and $g$ fixing each element of $W$. Now consider the effect of changing $Z$ to $Z'$ in this embedding. The result is an embedding $g' : C' \to \mathcal{M}_f'$ (same map, different structures) with $g'(C') \leq \mathcal{M}_f'$ and which is the identity on $W'$. As $B' \leq C'$ we have $g'(B') \leq \mathcal{M}_f'$ (by transitivity), and $g'$ is the identity on $A'$, as required.
\end{proof}
Now we need a Fourth Changing Lemma concerning pregeometries and localization. For our purpose it would be enough to have a finite version of this lemma, however we decide to present a proof that works for the infinite case. We still show the finite case separately inside of the proof, the reason for this is that for the infinite case the proof requires the predimension to be of a particular form (but covering all the ones that we introduced), while the finite case could be more easily generalized to other contexts.
First some notation.
\begin{nota} \rm
Let $\mathcal M$ be a structure and $PG(\mathcal M)$ a pregeometry on $\mathcal M$ with closure operation $c$. Let $Z\subseteq\mathcal M$. We write $PG_Z(\mathcal M)$ to denote the localization of the pregeometry $PG(\mathcal M)$ to the set $Z$, that is, the pregeometry with closure operation $c_Z(A) = c(Z \cup A)$ for $A \subseteq \mathcal{M}$. If $c$ is given by the dimension function $d$ and $Z$ is finite, then $c_Z$ is given by the dimension function $d(X/Z) = d(X\cup Z) - d(Z)$.
\end{nota}
\begin{lem} (Fourth Changing Lemma)
\label{L530}
Let $f$ be as in Lemma \ref{L525} and let $\mathcal M\in\bar{\mathcal C}_f$. Let $Z\leq\mathcal M$ and $Z'\in\bar{\mathcal C}_f$ be a structure with the same underlying set as $Z$. Let $\mathcal M'$ be obtained from $\mathcal M$ by replacing $Z$ by $Z'$, as we have done in Lemma \ref{L525}. Then $Z'\leq\mathcal M'\in\bar{\mathcal C}_f$ and
$$PG_Z(\mathcal M)=PG_{Z'}(\mathcal M').$$
\end{lem}
\begin{proof}
Let us set up some notation. The value of the predimension depends on the structure. However, instead of distinguishing the structure from the underlying set (as in most of the paper), here we use different notation for the predimension. Depending on whether we are working inside of $\mathcal M$ or $\mathcal M'$, given a finite subset $A\subseteq\mathcal M$ we write
\begin{enumerate}
\item[] $\delta (A)=|A|-\sum_{i\in I}\alpha_i|R_i^{A[\mathcal M]}|$
\item[] $\delta'(A)=|A|-\sum_{i\in I}\alpha_i|R_i^{A[\mathcal M']}|$
\end{enumerate}
with $f(i)=(n_i,\alpha_i)$, where $n_i$ is the arity of $R_i$. Also we write $d$ instead of $d_{\mathcal M}$ and $d'$ instead of $d_{\mathcal M'}$.
First we give a separate proof for the case when $Z$ is finite.
\begin{claim}
Assume $Z$ is finite. Then $PG_Z(\mathcal M)=PG_{Z'}(\mathcal M')$.
\end{claim}
\begin{proof}
First observe that for every finite $A\subseteq\mathcal M$ we have $$\delta(A)-\delta(A\cap Z)=\delta'(A)-\delta'(A\cap Z)$$ because all the changes are done inside $Z$.
All we have to prove is that $d(A/Z)=d'(A/Z)$. We do the following calculation:
\begin{enumerate}
\item[] $d(A/Z)=d(AZ)-d(Z)$
\item[] $=\delta({\rm cl}^{\mathcal M}(AZ))-\delta(Z)$ (because $Z\leq\mathcal M$)
\item[] $=\delta'({\rm cl}^{\mathcal M}(AZ))-\delta'(Z)$ (because ${\rm cl}^{\mathcal M}(AZ)\cap Z=Z$)
\item[] $\geq d'({\rm cl}^{\mathcal M}(AZ))-d'(Z)$ (because $Z'\leq\mathcal M'$)
\item[] $\geq d'(AZ)-d'(Z)$
\item[] $=d'(A/Z)$
\end{enumerate}
We get $d(A/Z)\geq d'(A/Z)$. Proceeding analogously we get $d'(A/Z)\geq d(A/Z)$. We then obtain $d(A/Z)=d'(A/Z)$, concluding the proof of the claim.
\end{proof}
Before proving the infinite case in full generality we need to prove the following weaker result, using an extra assumption.
\begin{claim}
Let $Z$ be possibly infinite, but $Z'$ obtained from $Z$ by just adding relations. Then $PG_Z(\mathcal M)=PG_{Z'}(\mathcal M')$.
\end{claim}
\begin{proof}
Our extra hypothesis means that for every set $B\subseteq\mathcal M$ we have ${\rm cl}_{\mathcal M}(B)\subseteq {\rm cl}_{\mathcal M'}(B).$ This is because from $\mathcal M$ to $\mathcal M'$ we are just adding relations, so if we could decrease the predimension value of $B$ by taking an superset in $\mathcal M$, then the same superset would work as well in $\mathcal M'$.
Let $A$ be a finite subset of $\mathcal M$. Again we have $$\delta(A)-\delta(A\cap Z)=\delta'(A)-\delta'(A\cap Z).$$ Also, all we need to prove is that $d(A/Z)=d'(A/Z)$.
Let $Z_0$ be a finite subset of $Z$ such that $d(A/Z)=d(A/Z_0)$ and $d'(A/Z)=d'(A/Z_0)$: this can be done by choosing $Z_0$ large enough. Let $Z_1={\rm cl}_{\mathcal M'}(AZ_0)\cap Z$. Then $Z_1\leq\mathcal M'$ and it is easy to see that ${\rm cl}_{\mathcal M'}(AZ_1)\cap Z={\rm cl}_{\mathcal M}(AZ_1)\cap Z=Z_1$.
In fact we have that $$Z_1\subseteq {\rm cl}_{\mathcal M}(AZ_1)\cap Z\subseteq {\rm cl}_{\mathcal M'}(AZ_1)\cap Z=Z_1.$$
Notice that we still have $d(A/Z)=d(A/Z_1)$ and $d'(A/Z)=d'(A/Z_1)$ as $Z_0\subseteq Z_1\subseteq Z$.
Now we are going to prove that we have $d(A/Z)=d'(A/Z)$ by proving the two inequalities separately.
\begin{enumerate}
\item[] $d(A/Z)=d(A/Z_1)$
\item[] $=d(AZ_1)-d(Z_1)$
\item[] $=\delta({\rm cl}_{\mathcal M}(AZ_1))-\delta(Z_1)$ (because $Z_1\leq\mathcal M$)
\item[] $=\delta'({\rm cl}_{\mathcal M}(AZ_1))-\delta'(Z_1)$ (because ${\rm cl}_{\mathcal M}(AZ_1)\cap Z=Z_1$)
\item[] $\geq d'(AZ_1)-d'(Z_1)$ (because $Z_1\leq\mathcal M'$)
\item[] $=d'(A/Z_1)$
\item[] $=d'(A/Z)$
\end{enumerate}
The same argument will prove the opposite inequality. Thus $d(A/Z)=d'(A/Z)$ for every finite $A\subseteq\mathcal M$. This proves that $PG_Z(\mathcal M)=PG_{Z'}(\mathcal M')$, proving the claim.
\end{proof}
Now we need to remove the extra assumption made in the previous claim in order to prove the full lemma.
For this we need to consider an intermediate structure $Z''$ with the same underlying set as $Z$ and $Z'$, with relations given by $R_i^{Z''}=R_i^Z\cap R_i^{Z'}$. Notice that $Z''\in\bar {\mathcal C}_f$ because we are just removing relations. Let $\mathcal M''$ be obtained from $\mathcal M$ by replacing $Z$ by $Z''$.
From $Z''$ to $Z$ we are just adding relations, thus by the previous claim we have $PG_{Z''}(\mathcal M'')=PG_Z(\mathcal M)$. But from $Z''$ to $Z'$ are also just adding relations, so again by the previous claim we have $PG_{Z''}(\mathcal M'')=PG_{Z'}(\mathcal M')$. Finally we can conclude that $PG_Z(\mathcal M)=PG_{Z'}(\mathcal M')$, proving the full lemma.
\end{proof}
\medskip
\medskip
\begin{cor}
\label{C531}
Let $f$ be as in Lemma \ref{L525} and let $\mathcal M\in\bar{\mathcal C}_f$. Let $Z\leq\mathcal M$ and $Z'\in\bar{\mathcal C}_f$ be a structure with the same underlying set as $Z$. Let $\mathcal M'$ be obtained from $\mathcal M$ by replacing $Z$ by $Z'$. Then $Z'\leq\mathcal M'\in\bar{\mathcal C}_f$ and
$$\textit{If }{\rm cl}^{d_{Z'}}(\emptyset)=Z'\textit{ then }PG_Z(\mathcal M)=PG(\mathcal M').$$
\end{cor}
\begin{proof}
Let $A$ be a finite subset of $\mathcal M'$. We have that ${\rm cl}^{d_{Z'}}(\emptyset)=Z'$ and $Z'\leq\mathcal M'$. Thus $Z'={\rm cl}^{d_{Z'}}(\emptyset)={\rm cl}^{d_{\mathcal M'}}(\emptyset)\cap Z'\subseteq {\rm cl}^{d_{\mathcal M'}}(A)$. So we have ${\rm cl}^{d_{\mathcal M'}}(A)={\rm cl}^{d_{\mathcal M'}}(AZ')$, this means that $PG_{Z'}(\mathcal M')=PG(\mathcal M')$. But by the Fourth Changing Lemma we have $PG_Z(\mathcal M)=PG_{Z'}(\mathcal M')$, thus $PG_Z(\mathcal M)=PG(\mathcal M')$.
\end{proof}
Now we can say something about the localization in $PG(\mathcal M_n)$. More precisely we show that localizing over a finite subset does not change the isomorphism type of the pregeometry.
\begin{thm}
\label{T532}
Let $Z$ be a finite subset of $\mathcal M_n$. Then $$PG_Z(\mathcal M_n)\simeq PG(\mathcal M_n).$$
\end{thm}
\begin{proof} If $Y, Z \subseteq \mathcal{M}_n$ have the same $d$-closure then clearly $PG_Z(\mathcal{M}_n) = PG_Y(\mathcal{M}_n)$. Thus we may assume that $Z \leq \mathcal{M}_n$.
Let $Z'$ be a structure in $\mathcal C_n$ with the same underlying set as $Z$ and $\delta(Z')=0$. This can be done: take for example $R_n^{Z'}=\{(c,\cdots,c):c\in Z'\}$. Let $\mathcal M_n'$ be obtained from $\mathcal M_n$ by replacing $Z$ by $Z'$. By the Third Changing Lemma we have $\mathcal M_n\simeq\mathcal M_n'$, in particular $PG(\mathcal M_n)\simeq PG(\mathcal M_n')$. But as $d_{Z'}(Z')=\delta(Z')=0$, by the Corollary \ref{C531} of the Fourth Changing Lemma we have $PG_Z(\mathcal M_n)=PG(\mathcal M_n')$, thus $PG_Z(\mathcal M_n)\simeq PG(\mathcal M_n)$.
\end{proof}
We finish this section with a consequence of the last theorem.
\begin{thm}
Let $m\neq n$ be natural numbers $\neq 0$. Then $PG(\mathcal M_m)$ and $PG(\mathcal M_n)$ are not locally isomorphic.
\end{thm}
\begin{proof}
By the last theorem, if they are locally isomorphic they must be isomorphic, which we know that is not true by Theorem \ref{T527}. Thus they are not locally isomorphic.
\end{proof}
\section{A generic structure for pregeometries}
The original motivation for this section was an attempt to prove that $PG(\mathcal M_n)$ and $PG(\mathcal M_m)$ \textit{are} isomorphic for $m,n\geq 3$. The idea was to try to recognize both pregeometries as a generic model of the same class of pregeometries and use the uniqueness of the generic model up to isomorphism: this idea was motivated by the fact that the isomorphism types of finite subpregeometries are the same in both pregeometries. Of course this idea cannot work as we have seen in Theorem \ref{T527} that $PG(\mathcal M_n)$ and $PG(\mathcal M_m)$ are not isomorphic. However we can in fact see $PG(\mathcal M_n)$ as a generic model of a class of pregeometries, but the appropriate \textit{embeddings} of the class change with $n$.
\begin{defn} \rm
Let $n$ be a natural number greater than zero and let $(\mathcal C_n,\leq_n)$ be the usual amalgamation class corresponding to arity $n$. Here we use the notation $\leq_n$ instead of $\leq$ to distinguish self-sufficiency corresponding to different arities. We now consider $(\mathcal C_n,\leq_n)$ as a category where the objects are the elements of $\mathcal C_n$ and where the morphisms between objects $A,B\in\mathcal C_n$ are the strong embeddings between the $A$ and $B$, more precisely the embeddings $f:A\to B$ such that $f(A)\leq_n B$.
Now we apply a functor to $(\mathcal C_n,\leq_n)$ that forgets the structure and remembers only the associated pregeometries and we obtain in this way a class of pregeometries $(P_n,\unlhd_n)$. More precisely, given two finite pregeometries $A\subseteq B$ we say that $A\unlhd_n B$ if and only if there are structures $\widetilde A,\widetilde B\in\mathcal C_n$ such that $PG(\widetilde A)=A$ and $PG(\widetilde B)=B$ and $\widetilde A\leq_n\widetilde B$. The objects are given by $P_n=\{A:A\mbox{ is a finite pregeometry and }\emptyset\unlhd_n A\}$. Also, we can see the category $(P_n,\unlhd_n)$ as a class of relational structures in the language $LPI$ of pregeometries, the morphisms are (some of the) embeddings in this language.
\end{defn}
\begin{rem} \rm
$P_n$ is closed under isomorphism, however not closed under substructures. Also $\unlhd_n$ is invariant under isomorphism, that is, if $f:B\to B'$ is an isomorphism of pregeometries and $A\subseteq B$ then we have $A\unlhd_n B$ if and only if $f(A)\unlhd_n f(B)$.
\end{rem}
\begin{prop}
$P_3\subsetneq P_4$.
\end{prop}
\begin{proof}
If $A\in P_3$ then there is a structure $\widetilde A\in\mathcal C_3$ with $PG(\widetilde A)=A$. Now we change the structure $\widetilde A$ to $\widehat A$ by replacing each relation $(a,b,c)$ by the relation $(a,b,c,c)$. Notice that $\widehat A\in\mathcal C_4$ and $PG(\widehat A)=PG(\widetilde A)=A$, thus $A\in P_4$. We have proved that $P_3\subseteq P_4$.
However this is a proper inclusion. Consider a structure $B\in\mathcal C_4$ consisting of $4$ distinct points $a,b,c,d$ and only one relation $(a,b,c,d)$. The pregeometry associated to $B$ can be described by saying that for $X\subseteq B$ and $|X|\leq 3$ we have $d_B(X)=|X|$ and $d_B(B)=3$. We have $PG(B)\in P_4$. However, this pregeometry is not in $P_3$ because there is no structure in $\mathcal C_3$ matching this pregeometry. In fact, suppose that there is a structure $B'\in\mathcal C_3$ with $PG(B')=PG(B)$. We have $d_{B'}(B')=\delta(B')=3$, thus there is exactly one relation in $B'$, say it is $(x,y,z)$ with $x,y,z\in B'$. We would have $d_{B'}(\{x,y,z\})\leq\delta(\{x,y,z\})=|\{x,y,z\}|-1<|\{x,y,z\}|$ which does not happen in $PG(B)$. We have proved that $P_3\subsetneq P_4$.
\end{proof}
\begin{prop}
In the class $(P_n,\unlhd_n)$ the relation $\unlhd_n$ is transitive.
\end{prop}
\begin{proof}
Suppose that we have $A\unlhd_n B$ and $B\unlhd_n C$. Suppose that $\widehat A,\widehat B,\widetilde B,\widetilde C\in\mathcal C_n$ such that $PG(\widehat A)=A$, $PG(\widehat B)=PG(\widetilde B)=B$, $PG(\widetilde C)=C$ and such that $\widehat A\leq_n\widehat B$ and $\widetilde B\leq_n\widetilde C$.
Now we construct a structure $\widehat C$ obtained from $\widetilde C$ by replacing $\widetilde B$ by $\widehat B$. By the First Changing Lemma we have $\widehat B\leq_n\widehat C\in\mathcal C_n$ and as $PG(\widetilde B)=PG(\widehat B)$ then by the Second Changing Lemma we have $PG(\widehat C)=PG(\widetilde C)=C$. Now we have $\widehat A\leq_n\widehat B$ and $\widehat B\leq_n\widehat C$. Thus by transitivity of $\leq_n$ we get $\widehat A\leq_n\widehat C$ with $PG(\widehat A)=A$ and $PG(\widehat C)=C$, thus $A\unlhd_n C$.
\end{proof}
\begin{prop}
The class $(P_n,\unlhd_n)$ is an amalgamation class. More precisely, if we have $A_0\unlhd_n A_1\in P_n$ and $A_0\unlhd_n A_2\in P_n$ then there exists $P\in P_n$ and embeddings of pregeometries $f_i:A_i\to P$, for $i\in\{1,2\}$, such that $f_i(A_i)\unlhd_n P$.
\end{prop}
\begin{proof}
Suppose that we have $A_0\unlhd_n A_1\in P_n$ and $A_0\unlhd_n A_2\in P_n$. Let $\widehat A_0\leq_n\widehat A_1\in\mathcal C_n$ and $\widetilde A_0\leq_n\widetilde A_2\in\mathcal C_n$ be the corresponding lifts to $(\mathcal C_n,\leq_n)$. Let $\widehat A_2$ be the structure obtained from $\widetilde A_2$ by replacing $\widetilde A_0$ by $\widehat A_0$ then by the First Changing Lemma we have $\widehat A_0\leq_n\widehat A_2\in\mathcal C_n$, moreover, as $PG(\widetilde A_0)=PG(\widehat A_0)$ then by the Second Changing Lemma we have $PG(\widehat A_2)=PG(\widetilde A_2)=A_2$. Now we use the fact that $(\mathcal C_n,\leq_n)$ is an amalgamation class and the fact that $\widehat A_0\leq_n\widehat A_1\in\mathcal C_n$ and $\widehat A_0\leq_n\widehat A_2\in\mathcal C_n$ to construct a structure $C\in\mathcal C_n$ and embeddings $f_1:\widehat A_1\to C$ and $f_2:\widehat A_2\to C$ fixing elements of $\widehat A_0$ such that $f_1(\widehat A_1)\leq_n C$ and $f_2(\widehat A_2)\leq_n C$. We have $PG(f_i(\widehat A_i))\unlhd_n PG(C)$, that is, $f_i(A_i)\unlhd_n PG(C)\in P_n$. In other words, $f_i:A_i\to PG(C)$ are strong embeddings of pregeometries fixing elements of $A_0$ and $PG(C)\in P_n$. This proves that $(P_n,\unlhd_n)$ is an amalgamation class.
\end{proof}
We would like to define a notion of generic model for the class $(P_n,\unlhd_n)$. However $P_n$ is not closed under substructures but this is not a real barrier, we just need to adapt the definition of generic model to this context.
\begin{defn} \rm
\label{D540}
Let $(\mathcal C,\leq)$ be a class of finite relational structures in a countable language $L$ with countably many isomorphism types. Let $\leq$ a binary relation such that $A\leq B$ implies that $A$ is a substructure of $B$. Assume that $\mathcal C$ is closed under isomorphism (but not necessarily under substructures) and that $\leq$ is invariant under isomorphism. We say that an $L$-structure $\mathcal M$ is a generic model for $(\mathcal C,\leq)$ if $\mathcal M$ is countable and satisfies $FC 1$ and $FC 2$ where:
\begin{enumerate}
\item[$FC1$] There is a chain $M_0\leq M_1\leq M_2\leq\cdots$ with $M_i\in\mathcal C$ and $\bigcup_{i\in\mathbbm N}M_i=\mathcal M$.
\item[$FC2$] (Extension property) If $A\leq M_i$ and $A\leq B\in\mathcal C$ then there are $j\in\mathbbm N$ and an embedding $f:B\to M_j$ such that $f_{|A}=Id_{|A}$ and $f(B)\leq M_j$.
\end{enumerate}
\end{defn}
\begin{defn} \rm
Let $\mathcal M$ be a generic model for a class $(\mathcal C,\leq)$ as in the above definition, with respect to a chain $M_0\leq M_1\leq M_2\leq\cdots$ Let $A$ be a finite substructure of $\mathcal M$ such that $A\in\mathcal C$. We say that $A\leq\mathcal M$ if $A\leq M_i$ for some $i\in\mathbbm N$. Notice that, at least apparently, the set of $\leq$-subsets of $\mathcal M$ depends not only on the structure of $\mathcal M$ but also on the choice of the chain (so for example, it is not \textit{a priori} necessarily preseved by automorphisms).
\end{defn}
Now we state the existence and uniqueness up to isomorphism of the generic model. The proof of this result is only a straightforward modification of the standard procedure.
\begin{prop}
Let $(\mathcal C,\leq)$ be as in definition \ref{D540}. Assume that $\emptyset\leq A$ for all $A\in\mathcal C$, that $\leq$ is transitive and that $(\mathcal C,\leq)$ is an amalgamation class. Then there is a generic model for $(\mathcal C,\leq)$ in the sense of definition \ref{D540} and it is unique up to isomorphism.
\end{prop}
Now we can prove the main result of this section.
\begin{thm}
\label{T543}
There is a generic model $\mathcal P_n$ (unique up to isomorphism) for the class $(P_n,\unlhd_n)$ and $$\mathcal P_n\simeq PG(\mathcal M_n)$$ where $\mathcal M_n$ is the generic model for the class $(\mathcal C_n,\leq_n)$.
\end{thm}
\begin{proof}
There is a unique generic model $\mathcal P_n$ of the class $(P_n,\unlhd_n)$ because we have seen that this class satisfies the conditions of last proposition. To prove that $\mathcal P_n\simeq PG(\mathcal M_n)$ we just need to prove that $PG(\mathcal M_n)$ is also a generic model for the class $(P_n,\unlhd_n)$.
Let $A_0\leq_n A_1\leq_n A_2\leq_n\cdots$ with $A_i\in\mathcal C_n$ and $\mathcal M_n=\bigcup_{i\in\mathbbm N}A_i$. In particular we have $PG(A_0)\unlhd_n PG(A_1)\unlhd_n PG(A_2)\unlhd_n\cdots$ with $PG(A_i)\in P_n$ and $PG(\mathcal M_n)=\bigcup_{i\in\mathbbm N}PG(A_i)$. We want to prove that $PG(\mathcal M_n)$ is a generic model for $(P_n,\unlhd_n)$ with respect to this chain. It remains to prove the extension property.
Suppose that $A\unlhd_n PG(A_i)$ and $A\unlhd_n B\in P_n$. We want to prove that there exist $j\in\mathbbm N$ and an embedding of pregeometries $f:B\to PG(\mathcal M_n)$ fixing the elements of $A$ and with $f(B)\unlhd_n PG(A_j)$. We have $A\unlhd_n PG(A_i)$, so there are structures $A',A_i'\in\mathcal C_n$ such that $PG(A')=A$ and $PG(A_i')=PG(A_i)$ and such that $A'\leq_n A_i'$. Let $\mathcal M_n'$ be the structure obtained from $\mathcal M_n$ by replacing $A_i$ by $A_i'$. Then by the First Changing Lemma we have $A_i'\leq_n\mathcal M_n'\in\bar{\mathcal C}_n$, by the Second Changing Lemma we have $PG(\mathcal M_n')=PG(\mathcal M_n)$ and by the Third Changing Lemma we have $\mathcal M_n'\simeq\mathcal M_n$.
We have $A\unlhd_n B\in P_n$, so there are structures $\widetilde A,\widetilde B\in\mathcal C_n$ such that $PG(\widetilde A)=A$ and $PG(\widetilde B)=B$ and $\widetilde A\leq_n\widetilde B\in\mathcal C_n$. Now we construct $B'$ obtained from $\widetilde B$ by replacing $\widetilde A$ by $A'$, by the the First Changing Lemma we get $A'\leq_n B'\in\mathcal C_n$, moreover, as $PG(A')=A=PG(\widetilde A)$ we can apply the Second Changing Lemma and we get $PG(B')=PG(\widetilde B)=B$.
Now we have $A'\leq_n A_i'\leq_n\mathcal M_n'$ and $A'\leq_n B'\in\mathcal C_n$. But we have $\mathcal M_n'\simeq\mathcal M_n$ so $\mathcal M_n'$ satisfies the extension property. We can apply the extension property to construct an embedding $f:B'\to\mathcal M_n'$ fixing the elements of $A'$ and such that $f(B')\leq_n\mathcal M_n'$. Now we apply the forgetful functor to this embedding and obtain an embedding $f:B\to PG(\mathcal M_n')=PG(\mathcal M_n)$ of pregeometries. It remains to prove that $f(B)\unlhd_n PG(\mathcal M_n)$.
For $j\geq i$ let $A_j'$ be the substructure of $\mathcal M_n'$ with the same underlying set as $A_j$. Notice that $A_i\leq_n A_{i+1}$ and $PG(A_i')=PG(A_i)$ imply by the first and second Changing Lemmas that $A_i'\leq_n A_{i+1}'$ and $PG(A_{i+1}')=PG(A_{i+1})$. Then in the next step $A_{i+1}\leq_n A_{i+2}$ and $PG(A_{i+1}')=PG(A_{i+1})$ imply that $A_{i+1}'\leq_n A_{i+2}'$ and $PG(A_{i+2}')=PG(A_{i+2})$... We repeat this procedure a countable number of times and we obtain a chain $A_i'\leq_n A_{i+1}'\leq_n A_{i+2}'\leq_n\cdots\leq_n A_j'\leq_n\cdots$ with $PG(A_j')=PG(A_j)$ and $A_j'\leq_n\mathcal M_n'$ because we have $\mathcal M_n'=\bigcup_{j\geq i}A_j'$.
Finally we have that $f(B')\leq_n\mathcal M_n'\Rightarrow f(B')\leq_n A_j'\mbox{ (for some } j\geq i)\Rightarrow f(B)=PG(f(B'))\unlhd_n PG(A_j')=PG(A_j)$. Thus $f(B)\unlhd_n PG(\mathcal M_n)$ as desired. We proved the extension property for $PG(\mathcal M_n)$ so we can conclude that $PG(\mathcal M_n)\simeq\mathcal P_n$.
\end{proof}
We end this section mentioning an alternative and more natural way of defining the generic model of $(P_n,\unlhd_n)$. However we do not know if such a generic model exists.
\begin{defn}\rm
Let $$\bar P_n:=\{A:A\mbox{ is pregeometry and there exists }A'\in\bar{\mathcal C}_n\mbox{ with }PG(A')=A\}$$ and let $P_n$ be as usual, that is, the class of finite pregeometries of $\bar P_n$.
Now we define an embedding relation $\sqsubseteq_n$ on $\bar P_n$ by saying: $A\sqsubseteq_n B$ if and only if there are $A',B'\in\bar{\mathcal C}_n$ such that $PG(A')=A,PG(B')=B$ and $A'\leq_n B'.$\end{defn}
It is clear that this is preseved by automorphisms of $B$.
Notice also that for finite pregeometries $\sqsubseteq_n$ coincides with $\unlhd_n$. In other words we have $(P_n,\sqsubseteq_n)=(P_n,\unlhd_n)$.
We can now give the alternative possible definition of generic model of $(P_n,\unlhd_n)$.
\begin{defn}\rm
Let $\mathcal M$ be a countable pregeometry. We say that $\mathcal M$ is a $\sqsubseteq_n$-generic model of $(P_n,\unlhd_n)$ if:
\begin{itemize}
\item $\mathcal M\in\bar P_n$
\item ($\sqsubseteq_n$-extension property) $A\sqsubseteq_n\mathcal M$, $A\sqsubseteq_n B\in P_n$ imply that there exists an embedding of pregeometries $f:B\to\mathcal M$ over $A$ such that $f(B)\sqsubseteq_n\mathcal M$.
\end{itemize}
\end{defn}
In fact, if the $\sqsubseteq_n$-generic of $(P_n,\unlhd_n)$ exists, then it is isomorphic to the $\unlhd_n$-generic, that is, isomorphic to $PG(\mathcal M_n)$. The following problem arises naturally:
\begin{problem}
Does the $\sqsubseteq_n$-generic model exists? In other words, does $PG(\mathcal M_n)$ satisfies the $\sqsubseteq_n$-extension property?
\end{problem}
|
\section{Introduction}
Suppose we are given a convex object $P$, and a multiset of discrete
translation vectors $\Lambda$. We wish to cover all of ${\mathbb R}^d$ by
translating $P$ using the translation vectors in $\Lambda$, such
that each point $x \in {\mathbb R}^d$ is covered exactly $k$ times. Along the
boundary points of $P$ there may be some technical lower-dimensional
problems, but if we require that each point which does not lie on the boundary of any translate of $P$
to be covered exactly $k$ times, then we call such a
covering of ${\mathbb R}^d$ a $k$-tiling. The traditional field of tilings of
Euclidean space by translates of a single convex object $P$ has a
long and rich history. The usual notion of a tiling by translations
is thus equivalent to the notion of a $1$-tiling. The reader is
invited to consult the books by Alexandrov \cite{Alex} and
Gruber~\cite{Gru} for a nice overview of the problem of tiling space
with translates of one convex body.
Tilings of ${\mathbb R}^d$ by translations of a single object have been
extensively studied from as early as 1881~\cite{Fed}, by the
mathematical crystallographer Fedorov, and are an active research
area today. For example, translational tilings of sets on the real line have
been studied in the $90$'s by Lagarias and Wang \cite{Jeff}.
There is also a beautiful recent survey article on tilings in various different mathematical contexts,
by Kolountzakis and Matolcsi~\cite{Kol3}.
We first note that if we have any $k$-tiling by a
convex object $P$, then it is an elementary fact that the convex
body $P$ must be a polytope, and we may therefore assume henceforth
that any convex object $P$ that $k$-tiles is a polytope.
Minkowski \cite{Min} has shown that if a convex body $P$ tiles
${\mathbb R}^d$ by a lattice, then it follows that $P$ is a centrally
symmetric polytope, with centrally symmetric facets. Venkov \cite{Ven} and
McMullen \cite{McM} proved that if a convex body $P$ tiles ${\mathbb R}^d$ by
translation, then for each of its codimension two faces $F$ there
are either four or six faces which are translates of $F$.
Here we find analogues of the necessary Minkowski conditions in the
case of general $k$-tilings, for any integer $k$ (see the main
Theorem \ref{thm_main} below).
Despite the beautiful characterization of $1$-tilers, given
collectively by Minkowski, Venkov, and McMullen, there is
still no known complete classification of polytopes that admit a
$k$-tiling, even in two dimensions. However, it is known that in
${\mathbb R}^2$, every $k$-tiling convex body has to be a centrally symmetric
polygon. Also, there exists a characterization by Bolle~\cite{Bol}
of all {\it lattice} $k$-tilings of convex bodies in ${\mathbb R}^2$.
Kolountzakis \cite{Kol2} proved that every $k$-tiling of ${\mathbb R}^2$ by a
convex polygon $P$, which is not a parallelogram, is a $k$-tiling
with a finite union of two-dimensional lattices.
\begin{figure}
\begin{center}
\include{Figures/fig_8gon}
\end{center}
\caption{ An octagon that $7$-tiles, but does not $1$-tile. Here,
each point in the interior of the octagon is covered exactly $7$
times, once we translate the octagon by all of the integer
translation vectors.} \label{octagon}
\end{figure}
A {\bf parallelotope} is, by definition, a convex polytope that
tiles (i.e. $1$-tiles) ${\mathbb R}^d$ facet-to-facet, with a lattice. That
is, its multiset of discrete translation vectors $\Lambda$ is in fact
given by a lattice in this case. It was proved by McMullen that if
a polytope tiles ${\mathbb R}^d$ with a discrete multiset of translations
$\Lambda$, then it must also admit a facet-to-facet tiling with a
{\it lattice}. In other words, McMullen showed that every
$1$-tiler must be a parallelotope. A very active area of current research
deals with the ``Voronoi conjecture'', which
A {\bf zonotope} in ${\mathbb R}^d$ is a polytope which can be represented as
a Minkowski sum of finitely many line segments. Equivalently, a
zonotope is a polytope in ${\mathbb R}^d$ with the property that all of its
$k$-dimensional faces are centrally symmetric, for all $1 \leq k
\leq d$. For example, the zonotopes in ${\mathbb R}^2$ are the centrally
symmetric polygons. A third equivalent definition for a zonotope is
that it is the projection of a $d$-dimensional cube, for some $d$.
For a good reference regarding these equivalences, and more about
polytopes, see the book by G. Ziegler \cite{Zie}.
It is clear that not all zonotopes are parallelotopes, an easy
example being furnished by the octagon (see fig.\,ref{octagon})
in two dimensions, which clearly does not tile by a lattice of translation vectors;
conversely, not all parallelotopes are zonotopes, as evidenced by
the example of the $24$-cell given below. In fact, McMullen has
given a beatiful characterization of those parallelotopes which are
zonotopes, in terms of unimodular systems (see \cite{McM2} for more
details).
Very little is known about the precise classification of polytopes
which $k$-tile ${\mathbb R}^d$ by translations. We outline some specific
open questions in the last section that pertain to the current state
of affairs along these lines. (see \cite{Gru}, pages 463-479 for
more details about $1$-tiling polytopes and some open problems).
\noindent We can now state the main result of this paper.
\begin{theorem}\label{thm_main}
If a convex polytope $k$-tiles ${\mathbb R}^d$ by translations, then it is
centrally symmetric and its facets are centrally symmetric.
\end{theorem}
We note that in ${\mathbb R}^3$ these two conditions are enough for a convex
body to necessarily be a zonotope. However, in dimension $4$ this is
no longer the case. A counterexample is furnished by the {\bf
$24$-cell}, which is a polytope in ${\mathbb R}^4$ which $1$-tiles ${\mathbb R}^4$, is
centrally symmetric, has centrally symmetric facets, but is not a
zonotope because it has $2$-dimensional faces that are triangles.
The $24$-cell is by definition the Voronoi region for the root
lattice $D_4$, and the reader may consult Coxeter \cite{Cox} for more
details.
Our proof of the main theorem above involves some new ideas that are quite different from Minkowski's proof for $1$-tilings.
We also prove the following counter-part to the main Theorem above.
\begin{theorem}\label{thm_main2}
Every rational polytope $P$ that is centrally symmetric and has
centrally symmetric facets must necessarily $k$-tile ${\mathbb R}^d$ with a
lattice, for some positive integer $k$.
\medskip \noindent
Moreover, the polytope $P$ must $k$-tile ${\mathbb R}^d$ with the rational
lattice $\frac{1}{N}{\mathbb Z}^d$, where $N$ is the lcm of the denominators
of all the vertex coordinates of $P$.
\end{theorem}
The paper is organized as follows. Section \ref{MAIN} is devoted to
the proof of the main result, namely Theorem \ref{thm_main}, and
comprises the main body of the paper. Section \ref{second main theorem} is short, and is devoted to the proof of Theorem
\ref{thm_main2}. In section \ref{FourierSection} we provide a more analytic approach of the main result, using Fourier techniques. Although it is
not crucial to supply another proof of the main result, this
approach provides a Fourier lens through which we can view our results.
Kolountzakis has also studied this problem using the
Fourier approach, and indeed our Fourier approach borrows some
techniques from his work. In section \ref{solid-angles} we give
another necessary and sufficient condition for a polytope to
$k$-tile ${\mathbb R}^d$, this time in terms of the solid angles of the
vertices of $P$. Finally, in section \ref{open} we mention some of
the important open problems concerning polytopes that $k$-tile
${\mathbb R}^d$.
\section{Definitions and preliminaries}
We adopt the usual conventions and notation from combinatorial
geometry. First, we recall that the {\bf Minkowski sum} of two
multisets $A \subset {\mathbb R}^d$ and $B \subset {\mathbb R}^d$ is the set
$A+B=\{a+b: a\in A, b\in B\}$, and that the Minkowski difference is
defined similarly by $A-B=\{a-b: a\in A, b\in B\}$.
For any set $A \subset {\mathbb R}^d$, its { \bf opposite set} is defined as
$-1 \cdot A=\{-a: a\in A\}$. We are particularly interested in the
case that both $A$ and $B$ are polytopes. We are also keenly
interested in the case that $A$ is a polytope and $B$ is a discrete
set of vectors, so that here $A+B$ is a set of translated copies of
the polytope $A$.
Given a convex body $P\subseteq{\mathbb R}^d$, $ \partial P$ denotes the
boundary of $P$. The standard convention for $\partial P$ includes
the fact that it has ($d$-dimensional) Lebesgue measure $0$, with
respect to the Lebesgue measure of ${\mathbb R}^d$. We let the interior of a
body $P$ be denoted by $\textup{Int}(P)$. Throughout the paper,
$\Lambda$ denotes an infinite discrete multiset of vectors in ${\mathbb R}^d$,
which is not necessarily a lattice.
We say that body $P$ {\bf $k$-tiles} ${\mathbb R}^d$ with the discrete
multiset $\Lambda$, if after translating $P$ by each vector
$\lambda\in\Lambda$, almost every point of ${\mathbb R}^d$ (except for the
boundary points of translated copies of $P$) is covered by exactly
$k$ of these translated copies of $P$. This condition can be written
more concisely as follows:
$$
\sum\limits_{\lambda\in\Lambda}1_{P+\lambda}(v)=k,
$$
for all ${v}\notin \partial P+\Lambda$.
We also recall that a { \bf facet} of a $d$-dimensional polytope is
any one of its $(d-1)$-dimensional faces. We let $V^{k}(F)$ denote
the $k$-dimensional volume of a $k$-dimensional object $F$, even if
$F$ resides in a higher dimensional ambient space, and sometimes we
simply write $V(F)$ for the $d$-dimensional volume of a
$d$-dimensional object $F \subset {\mathbb R}^d$. Finally, $\#(A)$ denotes
the cardinality of any finite multiset $A$.
\section{Proof of Theorem \ref{thm_main}} \label{MAIN}
To simplify the ensuing notation, we will assume that $-1 \cdot P$
$k$-tiles ${\mathbb R}^d$. We do not lose any generality, because $-1 \cdot
P$ $k$-tiles ${\mathbb R}^d$ if and only if $P$ also $k$-tiles ${\mathbb R}^d$.
We say that $v\in {\mathbb R}^d$ is in {\it general position} if there are no
points of $\Lambda$ on the boundary of $P+v$. In other words, $v
\notin \Lambda-\partial P$. We first prove the following elementary
but useful lemma, giving an equivalent condition for $k$-tiling in
terms of the number of $\Lambda$-points that lie in a 'typical'
translate of $P$.
\begin{lemma}\label{lm_restate}
A convex polytope $-1 \cdot P$ $k$-tiles ${\mathbb R}^d$ by translations with
a multiset $\Lambda$
if and only if\\
$\#(\Lambda\cap\{P+v\})=k$ for every $v$ in general position.
\end{lemma}
\begin{proof} Suppose that $-1 \cdot P$ $k$-tiles ${\mathbb R}^d$. Then for every
$v\notin \partial(-1 \cdot P)+\Lambda$ we can write
$$
k=\sum\limits_{\lambda\in\Lambda}1_{ \{-1 \cdot P+\lambda \} }(v)
=\sum\limits_{\lambda\in\Lambda}1_{P+v}(\lambda)=\#(\Lambda\cap\{P+v\}).
$$
It remains to mention that $\partial(-1 \cdot
P)+\Lambda=\Lambda-\partial P$. The proof in the other direction is
identical.
\end{proof}
\bigskip
We need to introduce some useful and natural notation for the
theorems that follow. Let ${\mathfrak P}$ be the vector space of the real
linear combinations of indicator functions of all convex polytopes
in ${\mathbb R}^d$. Thus, for example, if $P$ is any convex $k$-dimensional
polytope and $Q$ is any convex $m$-dimensional polytope, then
$\frac{1}{3} \cdot 1_P - 2 \cdot 1_Q \in {\mathfrak P}$.
One of the most important operators for us is the following boundary
operator, with respect to a vector $n\in{\mathbb R}^d$. It is the function
$\partial_n:{\mathfrak P}\rightarrow{\mathfrak P}$, defined as follows:
\[
\partial_{n}1_P=1_{F^{+}}-1_{F^{-}},
\]
where $F^{+}$ and $F^{-}$ are the (possibly degenerate) facets of
$P$ with outward pointing normals $n$ and $-n$, respectively. It
is a standard vector space verification that this operation is also
well-defined on ${\mathfrak P}$.
We also define this boundary operator on all of ${\mathfrak P}$, by letting it
act as a linear operator on the linear combinations of indicator
functions of polytopes. For example, another iteration of this
operator on $P \subset {\mathbb R}^3$ yields $\partial_{n_2} (\partial_{n_1}
P) = \partial_{n_2}( 1_{F^+}-1_{F^-} ) = (1_{E_1} - 1_{E_2})
-(1_{E_3} - 1_{E_4})$, where $E_1, E_2$ are the edges (which are by
definition the $1$-dim'l faces) of $F^+$, and $E_3, E_4$ are the
edges of $F^-$. In this case, each of the four edges is orthogonal
to both of the vectors $n_1$ and $n_2$, as is seen in Figure
$\ref{boundary.operator}$ below.
\begin{figure}
\begin{center}
\include{Figures/fig_2normals}
\caption{ The boundary operator with respect to $n_1$ picks out the
two facets $F^+$ and $F^-$, illustrating the definition of
$\partial_{n}1_P=1_{F^{+}}-1_{F^{-}}$. A second iteration of the
boundary operator, this time with respect to $n_2$, picks out the
four edge vectors $E_1, E_2, E_3$, and $E_4$, thus visually
illustrating the identity $
\partial_{n_2} (\partial_{n_1} P) = \partial_{n_2}( 1_{F^+}-1_{F^-} ) = (1_{E_1} - 1_{E_2}) -(1_{E_3} - 1_{E_4}).
$ } \label{boundary.operator}
\end{center}
\end{figure}
For the sake of convenience, we also define the action of the
boundary operator $\partial_n$ on convex polytopes $P$ as follows:
\[
\partial_{n}P=supp(\partial_{n}1_P)=\{v\in{\mathbb R}^d|\partial_{n}1_P(v)\neq 0\},
\]
so that the same symbol now acts on the subset $P$. However, we
note that the more salient operator for our discussions is still
$\partial_{n}1_P $.
It is useful to
utilize both of these actions, the first being an action on
indicator functions, and the second being an action on subsets of
points $P \subset {\mathbb R}^d$.
We call a sequence ${\mathfrak n}=(n_1,\ldots,n_m)$ of vectors in ${\mathbb R}^d$ an
orthogonal frame if they are pairwise orthogonal to each other. We
denote it by ${\mathfrak n}^{\bot}$ the subspace of ${\mathbb R}^d$ consisting of those
vectors which are orthogonal to every vector in the orthogonal frame
${\mathfrak n}$.
We define $\partial_{{\mathfrak n}} := \partial_{n_m}\ldots\partial_{n_1}$, a
composition of boundary operators that is read from right to left.
In case $m=0$, when an orthogonal frame ${\mathfrak n}$ is empty,
we define $\partial_{{\mathfrak n}}$ to be an identity operator. Similarly to $\partial_{n}P$ we define a boundary operator
relative to a whole frame ${\mathfrak n}=(n_1,\ldots,n_m)$:
\[
\partial_{{\mathfrak n}}P=supp(\partial_{{\mathfrak n}}1_P)=\{v\in{\mathbb R}^d|\partial_{{\mathfrak n}}1_P(v)\neq 0\}.
\]
Note that all the faces whose indicator functions appear in
$\partial_{{\mathfrak n}}1_P$ must have codimension $m$, must be parallel to
each other, and must have outward pointing normals $n_m$ or $-n_m$
in $\partial_{n_{m-1}}\ldots\partial_{n_1}P$.
We can now separate $\partial_{{\mathfrak n}}$ into two parts:
$\partial_{{\mathfrak n}}^{+}$ and $\partial_{{\mathfrak n}}^{-}$, corresponding to
faces with outward normals $n_m$ or $-n_m$, so that
$\partial_{{\mathfrak n}}1_P = \partial_{{\mathfrak n}}^{+}1_P - \partial_{{\mathfrak n}}^{-}1_P$.
In other words, if $\partial_n 1_P=1_{F^{+}}-1_{F^{-}}$, then by
definition $\partial_n^{+}1_P=1_{F^{+}}$, and
$\partial_n^{-}1_P=1_{F^{-}}$.
We say that $v \in {\mathbb R}^d$ is in general position w.r.t. the
orthogonal frame ${\mathfrak n}$, if there are no points of $\Lambda$ on any
boundary component of $\partial_{{\mathfrak n}}(P+v)$. A more formal
description which we will have occasion to use below is that
$v\notin \Lambda-\partial\partial_{{\mathfrak n}}P$.
Even though we only need to consider orthogonal frames of size
at most two in order to prove the theorem~\ref{thm_main}, we will prove
two following lemmas in general case, for an orthogonal frame of any size.
\begin{lemma}\label{lm_discrete}
Suppose $\#(\Lambda\cap\{P+v\})=k$ for every $v$ in general
position. Let ${\mathfrak n}=(n_1,\ldots,n_m)$ be an orthogonal frame in
${\mathbb R}^d$. Then for any $v$ in general position w.r.t ${\mathfrak n}$ the
following formula holds:
\begin{equation}
\sum\limits_{\lambda\in\Lambda}\partial_{{\mathfrak n}}1_{P+v}(\lambda)=0.
\label{fml_discrete}
\end{equation}
\end{lemma}
\begin{proof}
We proceed by induction on $m$. We remark that for $m=0$ the hypothesis tells us that
$\sum\limits_{\lambda\in\Lambda}\partial_{{\mathfrak n}}1_{P+v}(\lambda)=k$, and for $m=0$ this operator is by definition the identity operator.
However, for each
$m\geq 1$, we will show that $\sum\limits_{\lambda\in\Lambda}\partial_{{\mathfrak n}}1_{P+v}(\lambda)=0$.
Suppose that for an $(m-1)$-dimensional orthogonal frame
${\mathfrak n}^{\prime}=(n_1,\ldots,n_{m-1})$ and for every $v$ in general
position w.r.t. ${\mathfrak n}^{\prime}$ the formula holds:
$$\sum\limits_{\lambda\in\Lambda}\partial_{{\mathfrak n}^{\prime}}1_{P+v}(\lambda)=const$$
Now consider any $m$-dimensional orthogonal frame
${\mathfrak n}=(n_1,\ldots,n_m)$, and $v$ in general position w.r.t ${\mathfrak n}$. We
know that all $\Lambda$-points of $v+\partial_{{\mathfrak n}}P$ lie in
$v+Int(\partial_{{\mathfrak n}}P)$. Therefore, one can pick sufficiently small
$\epsilon'$, such that no $\epsilon'$-perturbation of $v$ by a
vector in ${\mathfrak n}^{\bot}$ removes or adds any $\Lambda$-points to
$v+\partial_{{\mathfrak n}}P$. Clearly, by doing so we do not change
$\sum\limits_{\lambda\in\Lambda}\partial_{{\mathfrak n}}1_{P+v}(\lambda)$. On
the other hand, we may choose an $\epsilon'$-perturbation,
$v_{\epsilon'}$, such that all $\Lambda$-points in
$v_{\epsilon'}+\partial\partial_{{\mathfrak n}^{\prime}}P$ get either inside
or outside of $v_{\epsilon'}+\partial_{{\mathfrak n}^{\prime}}P$ (see
fig.\,\ref{fig:boundary1}).
\begin{figure}
\begin{center}
\include{Figures/fig_boundary1}
\end{center}
\caption{ The $\epsilon'$ perturbation, along the $n^{\bot}$ direction, insures that all $\Lambda$ points have been removed from the four dotted edges on the upper facet and lower facet of $P + v$, giving us the set $\partial_{{\mathfrak n}^{\prime}}(P+v)$. Also, the $\epsilon$ perturbation, along the
$n_2$ direction, insures that all $\Lambda$ points on the right-hand bold edges, attached to the normal vector $-n_2$, will end up outside of the perturbed set, and that all $\Lambda$ points on the left-hand bold edges, attached to the normal vector $n_2$, will end up inside the perturbed set. }
\label{fig:boundary1}
\end{figure}
Then consider two small perturbations of $v_{\epsilon'}$ in the directions $n_m$ and
$-n_m$: $v_{\epsilon'}^{+}=v_{\epsilon'}+\epsilon n_m$ and $v_{\epsilon'}^{-}=v_{\epsilon'}-\epsilon n_m$, such that
$v_{\epsilon'}^{+}$ and $v_{\epsilon'}^{-}$ are in general position w.r.t ${\mathfrak n}^{\prime}$,
and $\epsilon$ small enough so that there are no points of $\Lambda$
that lie in $P+v_{\epsilon'}^{\pm}$ and do not lie in $P+v_{\epsilon'}$ (such an $\epsilon$ can
be found, because $\Lambda$ is discrete).
By induction,
$\sum\limits_{\lambda\in\Lambda}\partial_{{\mathfrak n}^{\prime}}1_{P+v_{\epsilon'}^{+}}(\lambda)=const=
\sum\limits_{\lambda\in\Lambda}\partial_{{\mathfrak n}^{\prime}}1_{P+v_{\epsilon'}^{-}}(\lambda)$.
On the other hand, recalling that by definition $\partial_{{\mathfrak n}}P=\partial_{n_m}\partial_{{\mathfrak n}'}P$,
\begin{align}
\sum\limits_{\lambda\in\Lambda}\partial_{{\mathfrak n}^{\prime}}1_{P+v_{\epsilon'}^{+}}(\lambda)-
\sum\limits_{\lambda\in\Lambda}\partial_{n_m}^{+}\partial_{{\mathfrak n}'} 1_{P+v_{\epsilon'}}(\lambda)
&=
\sum\limits_{\lambda\in\Lambda}\partial_{{\mathfrak n}^{\prime}}1_{P+v_{\epsilon'}}(\lambda)\cdot1_{Int(\partial_{{\mathfrak n}^{\prime}}P)}(\lambda) \\
&=
\sum\limits_{\lambda\in\Lambda}\partial_{{\mathfrak n}^{\prime}}1_{P+v_{\epsilon'}^{-}}(\lambda)-
\sum\limits_{\lambda\in\Lambda}\partial_{n_m}^{-}\partial_{{\mathfrak n}'} 1_{P+v_{\epsilon'}}(\lambda).
\end{align}
It follows that $\sum\limits_{\lambda\in\Lambda}\partial_{n_m}^{+}\partial_{{\mathfrak n}'}
1_{P+v}(\lambda)=\sum\limits_{\lambda\in\Lambda}\partial_{n_m}^{+}\partial_{{\mathfrak n}'}
1_{P+v_{\epsilon'}}(\lambda)= \sum\limits_{\lambda\in\Lambda}\partial_{n_m}^{-}\partial_{{\mathfrak n}'}
1_{P+v_{\epsilon'}}(\lambda)= \sum\limits_{\lambda\in\Lambda}\partial_{n_m}^{-}\partial_{{\mathfrak n}'}
1_{P+v}(\lambda)$, which gives us (\ref{fml_discrete}), since
$\partial_{{\mathfrak n}}=\partial_{n_m}^{+}\partial_{{\mathfrak n}'}-\partial_{n_m}^{-}\partial_{{\mathfrak n}'}$.
\end{proof}
For any polytope in ${\mathbb R}^d$ lying in an affine subspace parallel to
${\mathfrak n}^{\bot}$, we may consider a naturally defined lower dimensional
volume $\text{Vol}_{{\mathfrak n}}$. For example, if $d=3$ and ${\mathfrak n}=(n_1, n_2)$, we
get $\text{Vol}_{{\mathfrak n}}$ to be just a length of a line segment in ${\mathbb R}^3$. As
we know $\partial_{{\mathfrak n}}1_P$ is a finite sum of indicator functions of
polytopes lying in affine subspaces parallel to ${\mathfrak n}^{\bot}$ taken
with $+$ or $-$ signs. For each such indicator function $1_F$ let us denote
by $\text{Vol}_{{\mathfrak n}}(1_F)$ the volume of polytope $F$. Note that we can
take any measurable object in the affine subspace parallel to
${\mathfrak n}^{\bot}$ instead of $F$.
We now extend the notion of $\text{Vol}(S)$ to a more general notion of a signed linear combination of volumes.
We let $V_{{\mathfrak n}}(\partial_{{\mathfrak n}}1_P)$ denote the sum of the corresponding
volumes taken with different signs, and in a similar way we can write
$V_{{\mathfrak n}}$ for any sum of positive and negative indicator
functions. The next Lemma extends equality (\ref{fml_discrete}) in Lemma \ref{lm_discrete} from a discrete measure of facets to a continuous measure of facets.
\begin{lemma}\label{lm_integral}
Under the same assumptions of lemma~\ref{lm_discrete}, the following
formula holds:
$$ V_{\mathfrak n}( \partial_{{\mathfrak n}}1_P) =0.$$
\end{lemma}
\begin{proof}
Let us recall what we have so far. Lemma~\ref{lm_discrete} tells us
$\sum\limits_{\lambda\in\Lambda}\partial_{{\mathfrak n}}1_{P+v}(\lambda)=0$
for any $v$ with $v+\partial\partial_{{\mathfrak n}}P$ containing no $\Lambda$
points.
For each $\lambda\in\Lambda$ let us consider a set $S$ of vectors $v$ enjoying the property that
$\lambda\in v+\partial_{{\mathfrak n}}P$ and $\Lambda\cap
\{v+\partial\partial_{{\mathfrak n}}P\}=\emptyset$. We call the set $S$
${\mathfrak n}$-interior w.r.t. $\lambda$. We can also realize the set $S$ by excluding a finite number of lower dimensional polytopes (polytopes $F$ with
$V_{{\mathfrak n}}(F)=0$) from $\lambda-\partial_{{\mathfrak n}}P$. We call a vector ${\mathfrak n}$-internal if it belongs to
${\mathfrak n}$-interior for some $\lambda\in\Lambda$.
Assume now that $V_{{\mathfrak n}}(\partial_{{\mathfrak n}}1_P)=A_1\neq 0$. Let us
also write $V_{{\mathfrak n}}(|\partial_{{\mathfrak n}}1_P|)=A_2\ge|A_1|>0,$ where by
$|\partial_{{\mathfrak n}}1_P|$ we imply the sum of indicators of
$\partial_{{\mathfrak n}}1_P$ with all negative coefficients of indicators
switched to their absolute value.
For any $R>0$ we may consider a ball $B_R$ in ${\mathbb R}^d$ with the center
at origin and given radius $R$. Clearly, there is a constant
$C=C(P)$, such that $B_R + (-1)\partial_{{\mathfrak n}}P\subset B_{R+C}$ (see fig.\,\ref{fig:balls}). For
any positive real $R$ we define $N(R):=\#\{B_R\cap\Lambda\}$. For
each ${\mathfrak n}$-internal $v\in B_R$ we may rewrite the formula from lemma
\ref{lm_discrete} and get
$$\sum_{\lambda\in
B_{R+C}\cap\Lambda}\partial_{{\mathfrak n}}1_{\lambda-P}(v)=0.$$ This implies
$$V_{{\mathfrak n}}\left(1_{B_R}\cdot\sum_{\lambda\in
B_{R+C}\cap\Lambda}\partial_{{\mathfrak n}}1_{\lambda-P}\right)=0.$$
Also we know that $$\left|V_{{\mathfrak n}}\sum_{\lambda\in
B_{R+C}\cap\Lambda}\partial_{{\mathfrak n}}1_{\lambda-P}\right|=N(R+C)|A_1|.$$
\begin{eqnarray*}
\left|V_{{\mathfrak n}}\left(\sum_{\lambda\in
B_{R+C}\cap\Lambda}\partial_{{\mathfrak n}}1_{\lambda-P}\right)\right| &\leq&
\left|V_{{\mathfrak n}}\left(1_{B_R}\cdot\sum_{\lambda\in
B_{R+C}\cap\Lambda}\partial_{{\mathfrak n}}1_{\lambda-P}\right)\right|+
\\
&~&
\left|V_{{\mathfrak n}}\left(\left(1_{B_{R+2C}}-1_{B_{R}}\right)\cdot\sum_{\lambda\in
B_{R+C}\cap\Lambda}
\partial_{{\mathfrak n}}1_{\lambda-P}\right)\right|=
\\
&~&
\left|V_{{\mathfrak n}}\left(\left(1_{B_{R+2C}}-1_{B_{R}}\right)\cdot\sum_{\lambda\in\left(B_{R+C}\setminus
B_{R-C}\right)\cap\Lambda}\partial_{{\mathfrak n}}1_{\lambda-P}\right)\right|\leq
\\
&\leq& V_{{\mathfrak n}}\sum_{\lambda\in\left(B_{R+C}\setminus
B_{R-C}\right)\cap\Lambda}|\partial_{{\mathfrak n}}1_{\lambda-P}|=A_2\cdot\left(N(R+C)-N(R-C)\right).
\end{eqnarray*}
\begin{figure}
\begin{center}
\include{Figures/fig_balls_new}
\end{center}
\caption{None of the $\Lambda$-translates of $\partial_{{\mathfrak n}}1_{\lambda-P}$ can overlap more than two adjacent shells between the concentric balls. }
\label{fig:balls}
\end{figure}
Thus we get $(1-\frac{|A_1|}{A_2})N(R+C)\geq N(R-C)$, which
establishes an exponential grows of $N(R)$ in $R$. We can cover
$B_R$ by a disjoint union of $O(R^{2d})$ cubes whose side-length is $\frac{1}{R}$. Thus taking sufficiently large $R$ we can find a
cube $K$ with side-length $\frac{1}{R}$, which contains more than $k$ $\Lambda$-points. We can now translate $P$ so that the cube $K$ is contained in $P$, and therefore this translate of $P$ now contains more than $k$ $\Lambda$-points, a contradiction.
\end{proof}
In order to finish the proof of main theorem, we need the following
theorem by Minkowski~\cite{Min}.
\begin{theorem}[Minkowski]\label{thm_mink}
Convex polytope in ${\mathbb R}^d$ with given facet normals and facet
$(d-1)$-volumes is unique up to translation.
\end{theorem}
\bigskip
\begin{proof}[Proof of Theorem \ref{thm_main}]
We will first prove that $P$ is centrally symmetric. Take any pair of
facets of $P$, $F^{+}$ and $F^{-}$, with outward normals $n$ and
$-n$ respectively. Applying lemma~\ref{lm_integral} to ${\mathfrak n}=(n)$ we
get $V_{{\mathfrak n}}( \partial_{{\mathfrak n}}1_P) =0$, which means that
$V(F^{+})=V(F^{-})$. Since $n$ can be chosen arbitrarily, polytopes
$P$ and $(-1)\cdot P$ have equal codimension $1$ volumes of facets
in every direction. By theorem~\ref{thm_mink} we get that
$P=(-1)\cdot P+v$ for some translation vector $v$, so $P$ is
centrally symmetric.
Similarly we prove that everly facet of $P$ is centrally symmetric.
Given a pair of opposite facets $F_1$ and $F_2$ of $P$ with outward
normals $n_1$ and $-n_1$ respectively, consider any direction
$n_2\in (n_1)^{\bot}$ and two pairs of corresponding faces of
codimension $2$: $F_1^{+}$ and $F_1^{-}$ are facets of $F_1$ with
outward normals $n_2$ and $-n_2$ respectively, $F_2^{+}$ and
$F_2^{-}$ are facets of $F_2$ with outward normals $n_2$ and $-n_2$
respectively. Applying lemma~\ref{lm_integral} to ${\mathfrak n}=(n_1, n_2)$
we get $V_{{\mathfrak n}}( \partial_{{\mathfrak n}}1_P) =0$, which means that
$(V(F_1^{+})-V(F_1^{-}))-(V(F_2^{+})-V(F_2^{-}))=0$. But since $P$
is centrally symmetric, $F_1^{+}$ and $F_2^{-}$ are symmetric to
each other as well as $F_1^{-}$ and $F_2^{+}$, so
$V(F_1^{+})=V(F_2^{-})$ and $V(F_1^{-})=V(F_2^{+})$.
Combining the last
three equations we get an equality for codimension $2$ faces of $P$: $V(F_1^{+})=V(F_1^{-})$. It follows that as $(d-1)$-dimensional objects, $F_1$ and $(-1)\cdot F_1$, themselves have equal facets in every direction (in their affine span), and again by
theorem~\ref{thm_mink} we get that $F_1$ is centrally symmetric. But
since $F_1$ could be chosen arbitrarily among the facets of $P$, every facet of $P$ is
centrally symmetric, which concludes the proof of
theorem~\ref{thm_main}.
\end{proof}
\begin{remark}
We note that Lemma \ref{lm_discrete} gives us interesting information about the relationship between the $\Lambda$ points that lie in various faces, for any frame that has more than $2$ vectors. In contrast, Lemma \ref{lm_integral} does not give us any additional information about the codimension $3$ volumes (or higher codimension volumes). It is for this reason that we cannot conclude that codimension $3$ faces of a $k$-tiling polytope are centrally symmetric, and in fact they are not in general centrally symmetric, as the example of the $24$-cell shows.
\end{remark}
\section{Proof of Theorem \ref{thm_main2} }\label{second main theorem}
\begin{proof}
We may assume, without loss of generality, that our rational
polytope $P$ is an integer polytope, by dilating it by the lcm of
the denominators of all of the rational coordinates of its vertices.
Now, given that $P$ has integer vertices, we will show that the polytope $P$ $k$-tiles ${\mathbb R}^d$ with $\Lambda={\mathbb Z}^d$.
We claim that in every general position $P$ has an equal number of
integer points on every pair of opposite facets. Indeed, since it is
centrally symmetric and has centrally symmetric facets (and integer
vertices), any two opposite facets are translations of one another
by some integer vector. It follows that for every integer point on a
facet there is a corresponding integer point on an
opposite facet, so their numbers are equal.
Now, consider any two general positions of $P$, say $P+u$ and $P+v$.
There exists some path from $u$ to $v$ such that when we translate $P$ along this
path, no integer point of ${\mathbb Z}^d$ collides with any co-dimension $2$ face of the translates of $P$ along this path (see fig.\,\ref{moving.polygon}).
But
since in any general position the number of integer points on two
opposite facets of $P$ are equal, it follows that the number of points
inside $P$ along this path is constant. We conclude that any two
general positions of $P$ have the same number of interior integer points,
say $k$. Thus, $-P$ $k$-tiles ${\mathbb R}^d$ with the lattice ${\mathbb Z}^d$.
\end{proof}
\begin{figure}
\begin{center}
\include{Figures/fig_path}
\end{center}
\caption{ This polygon illustrates that fact that there is always a
continuous path that a polygon $P$ may take so that the vertices of
$P$ (and in general the codimension $2$ faces of $P$) never pass
through the discrete set of translations vectors $\Lambda$, shown
here as a lattice.} \label{moving.polygon}
\end{figure}
\newpage
\section{An analytic approach, using Fourier techniques}\label{FourierSection}
In this section we give another proof of the main result, Theorem
\ref{thm_main}, but this time from the Fourier perspective, so that
we may employ the language of generalized functions. The reader may
consult the classic reference \cite{Stein} for more information about Fourier analysis on Euclidean spaces. We begin once again with
the definition of a $k$-tiling. Thus, we suppose that a polytope
$P$
$k$-tiles ${\mathbb R}^d$ with some discrete multiset $\Lambda$. In other
words, we assume that
$$
\sum\limits_{\lambda\in\Lambda}1_{P+\lambda}(v)=k,
$$
for all ${v}\notin \partial P+\Lambda$. We can rewrite this
condition as a convolution of generalized functions, as follows:
\begin{equation}\label{convolution}
1_P * \delta_\Lambda = k,
\end{equation}
where $1_P$ is the indicator function of $P$, and where
$\delta_\Lambda := \sum_{\lambda \in \Lambda} \delta_\lambda$,
where $\delta_\lambda$ is the unit point mass for the point $\lambda
\in \Lambda$. That is, $\delta_\lambda$ equals $1$ at the point
$\lambda$ and zero elsewhere. We first differentiate both sides
of (\ref{convolution}), with respect to any $\xi \in {\mathbb R}^d$,
obtaining
\begin{equation}\label{Step1}
\frac{d}{d \xi} \left( 1_P * \delta_\Lambda \right)= (\frac{d}{d
\xi} 1_P) * \delta_\Lambda = 0.
\end{equation}
\noindent Next, we take the Fourier transform of both sides of
(\ref{Step1}), obtaining
\begin{equation}\label{Step2}
\left( \xi \hat 1_P \right) \, \hat \delta_\Lambda = 0,
\end{equation}
where the last step uses the standard Fourier identities $\widehat{
\left( \frac{d}{dx} F \right)}(\xi) = \xi \hat F(\xi)$, and
$\widehat{F*G} = \hat F \hat G$. If we now have some more detailed
knowledge about $\hat 1_P$, then we can use (\ref{Step2}) to proceed
further.
The next result is a useful combinatorial version of Stokes'
formula, which holds for the Fourier transform of the indicator
function of any polytope. This is a result about $\hat 1_P$ that
appears to be not as well-known, so we prove it in complete detail.
For the transform of a function on ${\mathbb R}^d$, we use the standard
definition:
\[
\hat 1_P(\xi) := \int_P {\textup{exp}}(2\pi i \langle \xi, x \rangle ) dx,
\]
valid for any $\xi \in {\mathbb R}^d$, because $P$ is compact.
\begin{theorem}\label{Stokes}
Let $F$ be a $k$-dimensional polytope in $\mathbb R^d$, for any $k
\leq d$. Let $Proj_F(\xi)$ denote the orthogonal projection of $\xi$
onto the $k$-dimensional subspace of $\mathbb R^d$ that is parallel
to $F$. Moreover, for each $(k-1)$-dimensional face $G \in \partial
F$, let $n_G$ be its outward pointing normal vector. Then the
Fourier transform of the indicator function of $F$ can be written as
follows:
\medskip
Case I. \ \ If $Proj_F(\xi) = 0$, then
\[
\hat 1_F(\xi) = V^k(F) {{\textup{exp}}}(2 \pi i \Phi ),
\]
where $\Phi$ is the constant value of the function $\phi(x) :=
\langle \xi, x \rangle$ on $F$.
\medskip
Case II. \ \ If $Proj_F(\xi) \not= 0$, then
\[
\hat 1_F(\xi) = -\frac{1}{2\pi i} \sum_{G \in \partial F}
\frac{\langle Proj_F(\xi), n_F \rangle}{|| Proj_F(\xi)||^2 } \hat
1_G(\xi).
\]
\end{theorem}
\begin{proof}
We note that the gradient of $\phi(x):= \langle \xi, x \rangle$,
with respect to the Riemannian structure of the submanifold $F
\subset {\mathbb R}^d$, is simply the projection of the $d$-dimensional
Euclidean gradient of $\phi$ onto $F$. We denote this projection by
$\textup{grad}_F\phi$ in the argument that follows. Fix any $\xi
\in {\mathbb R}^d$.
Case I. \ If $Proj_F(\xi) = 0$, then $\textup{grad}_F\phi = 0$, so
that $\phi$ is constant on $F$. The Fourier integral defining $\hat
1_F(\xi)$ in this case degenerates into an integral of a constant
function on $F$, hence the conclusion of the theorem for this case.
Case II. \ If $Proj_F(\xi) \not= 0$, then from the linearity of
$\phi$ it follows that $\textup{grad}_F\phi(x) = 2\pi Proj_F(\xi)$
is a constant vector field on $F$. The identity
\[
\textup{div}_F \textup{grad}_F exp(2\pi i \phi(x)) = (2\pi i)^2
||\textup{grad}_F \phi(x)||^2 exp(2\pi i \phi(x))
\]
shows us that $\textup{exp}(2\pi i \phi(x))$ is an eigenfunction of
the Laplacian, with eigenvalue
\[
\lambda := (2\pi i)^2 ||\textup{grad}_F\phi||^2 \not=0.
\]
Hence
\begin{align*}
\hat 1_F(\xi) &= \int_F e^{2\pi i \phi(x)} \\
&= \frac{1}{\lambda}
\int_F \textup{div} \left(\textup{grad}_F e^{2\pi i \phi(x)} \right) dF \\
&= \frac{1}{\lambda} \sum_{G \in \partial F} \int_G
\langle \ \textup{grad}_F e^{2\pi i \phi(x)}, n_G \ \rangle dG,
\end{align*}
where we've used the identity for the Laplacian above in the second
equality, and Stokes' theorem for the polytope $F$ and its finite
collection of boundary polytope components $G \in \partial F$ in the
third equality. Unravelling the remaining definitions, we get:
\begin{align*}
\hat 1_F(\xi) &= \frac{2\pi i}{\lambda} \sum_{G \in \partial F}
\langle \textup{grad} \phi(x), n_G \ \rangle \int_G e^{2\pi i \phi(x)} dG \\
&= -\frac{1}{2\pi i} \sum_{G \in \partial F}
\frac{ \langle Proj_F (\xi), n_G \rangle}{ ||Proj_F(\xi)||^2 } \hat 1_G(\xi).
\end{align*}
\end{proof}
The result above uses functions, as opposed to generalized
functions, but we may indeed pass to generalized functions, abusing
the notation $\hat 1_P$ only slightly.
Applying Theorem (\ref{Stokes}) above to the generalized function
$1_P$, we may continue from (\ref{Step2}) to get the identity
\begin{equation}\label{Step3}
\left( \sum_{F \in \partial P} \xi \frac{ \langle \xi, n_F \rangle
}{\langle \xi, \xi \rangle} \hat 1_F \right) \, \hat \delta_\Lambda
= 0,
\end{equation}
valid for any nonzero $\xi \in {\mathbb R}^d$. We also note that the sum
runs over all the (codimension $1$) facets $F$ of the boundary
$\partial P$. It now follows, upon taking the inner product with
$\xi$, that
\begin{equation}\label{Step4}
\left( \sum_{F \in \partial P} \langle \xi, n_F \rangle \, \hat
1_F \right) \, \hat \delta_\Lambda = 0.
\end{equation}
Taking Fourier transforms again, we may rewrite the last equation as
\begin{equation}\label{Step5}
\left( \sum_{F \in \partial P} \left( \frac{d}{d \,n_F} \right)
( 1_F ) \right) * \delta_\Lambda = 0.
\end{equation}
We now focus our attention on each {\it pair} of facets of $P$, as
in the first section. Thus, we consider a facet $F^+$ with its
outward pointing normal $n(F)$, and a parallel facet $F^-$, with its
outward pointing normal $-n(F)$.
\begin{lemma}\label{FacetPairs}
For each facet $F$ of $P$, we have the identity
\[
\left( 1_{F^+} - 1_{F^-} \right) * \delta_\Lambda = 0.
\]
\end{lemma}
\begin{proof}
We assume that $ \left( 1_{F^+} - 1_{F^-} \right) *
\delta_\Lambda \not= 0$. Therefore there exists a small ball $B_r$,
of radius $r$, such that for any nonnegative, nonzero test
function $f$ whose support is contained in $B_r$, we have $ \langle
\left( 1_{F^+} - 1_{F^-} \right) * \delta_\Lambda, f \rangle \not=
0$. We may further assume that the support of $f$ is disjoint from
the support of $ \left( 1_{G^+} - 1_{G^-} \right) *
\delta_\Lambda$, for any facet $G$ of $P$ where $G \not= F$. Indeed,
the discreteness of $\Lambda$ guarantees that we can find such a
ball $B_r$ on which $f$ satisfies the above conditions.
Now we construct a test function $g$ whose support is contained in
$B_r$, with positive derivative $\left( \frac{d}{d n_F} \right)$
along the direction $n_F$, in a small $\epsilon$ vicinity of $B_r
\cap Supp( \left( 1_{F^+} - 1_{F^-} \right) * \delta_\Lambda ) :=
D_\epsilon$. To construct such a $g$, we first restrict $f$ to
$D_\epsilon$, call it $f_0$. We now multiply $f_0$ by a one dimensional smooth bump function $b$ whose derivative on $[-\epsilon, \epsilon]$ is
positive, and whose support lives in $[-2\epsilon, 2\epsilon]$. Thus $g := f_0 \cdot b$ has positive derivative on $D_\epsilon$. When we
insert this $g$ into (\ref{Step5}), we arrive at a contradiction. Indeed $<\left( 1_{G^+} - 1_{G^-} \right) * \delta_\Lambda,\frac{d}{d n_G} g>=0$ for $G \not= F$ because the choice of the support of $g$. On the other hand $<\left( 1_{F^+} - 1_{F^-} \right) * \delta_\Lambda, \frac{d}{d n_F} g>\not=0$ by the construction, since $\frac{d}{d n_F}g$ is positive in the vicinity of the support of $\left( 1_{F^+} - 1_{F^-} \right) * \delta_\Lambda$.
\end{proof}
We finish this section by remarking that iteration of Lemma \ref{FacetPairs} allows us to establish the same conclusion as Lemma \ref{lm_discrete}. The next iteration would be applied to a normal vector to a facet of $F^+$ within the affine span of the facet $F^+$. The Fourier analogue of Lemma \ref{lm_integral} involves the scalar product of $ \left( 1_{F^+} - 1_{F^-} \right) * \delta_\Lambda$ against an ``approximate identity'' function, compactly supported on a large ball. The main Theorem \ref{thm_main} now follows in a similar manner as in the previous section.
\section{Another equivalent condition for $k$-tiling, using solid angles}\label{solid-angles}
Here we show that it is possible to reinterpret the condition that a
polytope $k$-tiles ${\mathbb R}^d$ by considering all of the solid angles
$\omega_{P}(\lambda)$ of the $d$-dimensional convex polytope $P$, at
each point $\lambda \in \Lambda$. For any point $\lambda \in {\mathbb R}^d$, we define
the solid angle at $\lambda$ to be the proportion of a small sphere of radius $R$, centered at $\lambda$, which intersects $P$.
More precisely, the solid angle is defined by
\[
\omega_{P}(\lambda)=\lim\limits_{R\rightarrow 0}
\frac{V\left(\{\lambda+B_R)\}\cap P\right)}{V(B_R)},
\]
where $V(S)$ is the $d$-dimensional volume of $S$. The following
Theorem is of independent interest, showing another interesting
equivalent condition for $k$-tiling Euclidean space.
\begin{theorem}
A polytope $P$ $k$-tiles ${\mathbb R}^d$ with the multiset $\Lambda$ if and only if
\[
\sum\limits_{\lambda\in{\Lambda}}\omega_{P+v}(\lambda)=k,
\]
for every $v \in {\mathbb R}^d$.
\end{theorem}
\begin{proof}
Suppose that $P$ $k$-tiles ${\mathbb R}^d$ with the multiset $\Lambda$. We know from Theorem \ref{thm_main} that $P$ must be centrally symmetric, and therefore
$-P$ $k$-tiles as well, with the multiset $\Lambda$. By Lemma \ref {lm_restate} $\#(\Lambda\cap\{P+x\})=k$ for
almost every $x\in{\mathbb R}^d$. We can therefore integrate this equality in the variable $x$,
over a $d$-dimensional ball $ B_R(v)$ with center in $v$ and
radius $R$, as follows:
\begin{align*}
k \cdot V(B_R(v))=\int\limits_{B_R(v)}k \,dx &=
\int\limits_{B_R(v)}\#(\Lambda\cap\{P+x\})dx \\
&= \int\limits_{B_R(v)}\sum\limits_{\lambda\in\Lambda}1_{\lambda-P}(x)dx \\
&= \sum\limits_{\lambda\in\Lambda}\int\limits_{B_R(v)}1_{\lambda-P}(x)dx \\
&=\sum\limits_{\lambda\in\Lambda}V(B_R(v)\cap\{\lambda-P\}) \\
&=\sum\limits_{\lambda\in\Lambda}V(\{\lambda-B_R\}\cap\{P+v\})
\end{align*}
\noindent
It follows that
$k=\sum\limits_{\lambda\in\Lambda}\frac{V(\{\lambda-B_R\}\cap\{P+v\})}{V(B_R(v))}$,
which approaches
$\sum\limits_{\lambda\in\Lambda}\omega_{P+v}(\lambda)$ as $R$ goes
to $0$.
\medskip
\noindent
In the other direction, the assumption that $\sum\limits_{\lambda\in{\Lambda}}\omega_{P+v}(\lambda)=k$ is, in general position, equivalent to
the statement that $\#(\Lambda\cap\{P+x\})=k$. By Lemma \ref {lm_restate} we conclude that $-P$ $k$-tiles with the multiset $\Lambda$.
Finally, by Theorem \ref{thm_main} we know that $P$ is centrally symmetric, so that $P$ $k$-tiles with the same multiset $\Lambda$.
\end{proof}
\medskip
We note that a particularly interesting choice of $v$ in this
Theorem is the value $v=0$, so that we can in fact have points in
$\Lambda$ coincide with vertices of $P$. This equivalent condition
allows us to consider such coincidences without having to translate
$P$ into general position.
\section{Some open questions}\label{open}
\bigskip
We conclude our paper with some fascinating open questions which the main results of the present paper suggest as a natural
research direction for $k$-tilings, a relatively new area.
\bigskip
\bigskip
\noindent
1. \ Recall that the Venkov-McMullen condition for the existence of belts consisting of $4$ or $6$ parallel codimension $2$ faces allowed an
``if and only if" characterization for $1$-tiling polytopes. Find the analogous additional condition that would give a complete characterization for
$k$-tiling polytopes.
\medskip
\noindent
2. \ Classify the combinatorial types of all polytopes which $k$-tile ${\mathbb R}^d$ by translations.
\medskip
We note that for the classical question of $1$-tiling ${\mathbb R}^d$ by
parallelotopes (and parallelotopes are the only objects that can
tile ${\mathbb R}^d$, by McMullen's theorem), there are exactly $5$
combinatorially distinct parallelotopes in ${\mathbb R}^3$, and exactly $52$
distinct parallelotopes in ${\mathbb R}^4$. It is still not known how many
combinatorially distinct parallelotopes there are in dimensions $5$
and higher. It is also not known how many facets a parallelotope
may have in general (see \cite{Gru} for references).
\medskip
\noindent
3. \ Prove or disprove that if any polytope $k$-tiles ${\mathbb R}^d$ by
translations, then it also $m$-tiles ${\mathbb R}^d$ by a lattice, for a
possibly different $m$.
\medskip
This would give an analogue of the McMullen Theorem for $1$-tiling parallelotopes in ${\mathbb R}^d$, but appears to be a very difficult problem.
\bigskip
\noindent
4. \ Prove or disprove that if a $3$-dimensional polytope, which is not a prism, $k$-tiles ${\mathbb R}^3$ by translations with a
multiset $\Lambda$, then $\Lambda$ is a union of a finite number of $3$-dimensional lattices.
\medskip
This would prove the $3$-dimensional analogue of Kolountzakis' $2$-dimensional result \cite{Kolsurvey}.
\bigskip
\noindent
5. \ Is it always true that whenever $P$ $k$-tiles with a multiset $\Lambda$, it follows that $\Lambda + v = \Lambda$ for some
$v \in {\mathbb R}^2$ ? (This is one of Kolountzakis' open questions in \cite{Kol2})
\bigskip \noindent
6. \ Find, or estimate, the smallest $k$ for which a given polytope can $k$-tile ${\mathbb R}^d$. This problem is open even in two dimensions.
\bigskip
|
\section{}
Recently the gauge-gravity duality \cite{maldacena98, witten98} has provided a new means by which to explore strongly interacting theories in condensed matter systems. One of the earliest and most successful of such models has been shown to exhibit the key properties of superconductivity: a phase transition at a critical temperature, where a spontaneous symmetry breaking of a $U(1)$ gauge symmetry in the bulk gravitational theory corresponds to a broken global $U(1)$ symmetry on the boundary, and the formation of a charged condensate. Gravity duals have so far been found for $s$-wave superconductors \cite{herzog08, hartnoll08} (in which the Cooper electron pairs have angular momentum $l=0$), as well as for $p$-wave ($l=1$) \cite{gubser08, roberts08, gubser08b} and $d$-wave ($l=2$) \cite{chen10, benini10} superconductors. The field of holographic superconductivity has since grown rapidly (see Ref.\ \cite{reviews} for reviews). Such a dual description provides a window through which we might hope to obtain insight into the properties and behaviors of superconductors and superfluids that defy description by more traditional approaches.
The basic recipe for creating a holographic superconductor involves the introduction of a black hole and a charged scalar field into anti-de Sitter (AdS) spacetime in $d$+1 dimensions. According to the AdS/CFT correspondence, this theory is dual to a $d$-dimensional field theory that exists on the boundary of this space, where the boundary value of the bulk field is related to the expectation value of an operator, which is interpreted as the superconducting order parameter, in the boundary theory, and the temperature of the boundary theory is given by the Hawking temperature of the black hole. Below a critical temperature $T_c$ the field condenses, and the operator on the boundary acquires a nonzero expectation value, which corresponds to a superconducting phase transition.
The simplest case in the above scenario is that in which the electromagnetic and scalar fields are a function only of the radial AdS coordinate, having a boundary value which is spatially uniform. Such solutions were soon extended to include solutions for isolated vortices, which feature a spatially non-uniform order parameter, in $s$-wave superconductors \cite{montull09, albash09, domenech10} and vortex lattices in $s$-wave \cite{maeda10} and $d$-wave \cite{zeng10} superconductors. Extending these results to $p$-wave superconductors is of interest for a number of reasons. The order parameter in these systems has multiple components and breaks time-reversal symmetry, which leads to a richer set of possibilities than is possible in the simpler $s$-wave superconductors. In addition, the fact that the gravity dual is a $SU(2)$ gauge theory makes its development more straightforward than for $d$-wave superconductors, where the dual theory involves a spin-2 field in a gravitational background. The theory of holographic $p$-wave superconductors is also attractive because it has fewer free parameters than the charged scalar field theories that describe $s$- and $d$-wave superconductors.
The action proposed in Ref.\ \cite{gubser08b} to describe a $p$-wave holographic superconductor is
\be
\label{action}
S = \frac{1}{\kappa^2} \int d^4x \sqrt{-g} \left[ R + \frac{6}{L^2} - \frac{1}{4 q^2} (F^a_{\mu \nu})^2 \right],
\ee
where $\kappa$ is the gravitational coupling, $R$ is the Ricci scalar curvature, $L$ is the radius of AdS space and $F^a_{\mu \nu} = \partial_\mu A^a_\nu - \partial_\nu A^a_\mu + \epsilon^{abc} A^b_\mu A^c_\nu$ is the $SU(2)$ Yang-Mills field strength. The bulk gravitational theory is described by the AdS-Schwarzschild metric:
\be
ds^2 = \frac{L^2}{z^2} \left[ -g(z) dt^2 + \frac{dz^2}{g(z)} + dx^2 + dy^2 \right] ,
\ee
where $z$ is the radial AdS coordinate and $g(z) = 1-(z/z_0)^3$. The black hole horizon at $z=z_0$ is related to the Hawking temperature of the black hole, which is equal to the temperature of the boundary theory, by $z_0 = 3 / (4 \pi T)$. Our calculations will be performed in the probe limit, in which there is no back-reaction of the gauge field on the metric \cite{hartnoll08}.
As a starting point we consider the following ansatz for the gauge field:
\be
\label{ansatz}
\begin{split}
A =& \tau^3 ( \Phi dt + A^3_x dx + A^3_y dy )+ w_+ (\tau^1 dx + \tau^2 dy)
\\ & + w_- (\tau^1 dx - \tau^2 dy).
\end{split}
\ee
Here $\tau^a$ are the generators of $SU(2)$, which obey the relation $[\tau^a,\tau^b] = \epsilon^{abc} \tau^c$ and are related to the Pauli matrices by $\tau^a = \sigma^a / 2i$. Following Refs.\ \cite{gubser08, roberts08, gubser08b}, we interpret the $U(1)$ subgroup generated by $\tau^3$ as the group of electromagnetism, so that $\Phi (x,y,z)$ and $A^3_{x,y} (x,y,z)$ are the electromagnetic scalar and vector potentials, respectively. Because $\tau^3$ generates a rotation in the 1-2 ``plane," which is the analogue of a rotation in the complex plane in an ordinary Ginzburg-Landau theory of superconductivity, the (real) scalar fields $w_\pm (x,y,z)$ are charged under this $U(1)$, and they represent the amplitudes of the $p_x \pm i p_y$ components of the superconducting order parameter, respectively.
To study the case of an isolated vortex, we switch boundary coordinates from $(x,y)$ to $(r,\phi)$. In an ordinary Ginzburg-Landau theory with a complex order parameter, the vortex solution is found by replacing $\psi (r, \phi) \to e^{i n \phi} \psi (r)$, such that the phase changes by $2 \pi n$ as one goes around the vortex core, and $n$ is known as the winding number. By analogy, the vortex ansatz for the $p_x \pm i p_y$ superconductor is given by the replacements $w_\pm (r,\phi,z) \to \mathrm{exp}(2 n_\pm \phi \tau^3) w_\pm (r,z)$ in Eq.\ \eqref{ansatz}, where $n_\pm$ are the (integer) winding numbers for the two components of the superconducting order parameter. With this modification, the gauge field ansatz becomes
\be
\begin{split}
A^1_x &= w_+ (r,z) \cos(n_+ \phi) + w_- (r,z) \cos (n_- \phi)
\\ A^1_y &= - w_+ (r,z) \sin(n_+ \phi) - w_- (r,z) \sin (n_- \phi)
\\ A^2_x &= w_+ (r,z) \sin(n_+ \phi) - w_- (r,z) \sin(n_- \phi)
\\ A^2_y &= w_+ (r,z) \cos(n_+ \phi) - w_- (r,z) \cos (n_- \phi).
\end{split}
\ee
Furthermore, we assume that the electromagnetic scalar and vector potentials are rotationally symmetric and given by $\Phi(r,z)$ and $A^3_\phi (r,z)$, respectively.
The next step is to determine and numerically solve the equations of motion for this ansatz. The Yang-Mills equations are
\be
\label{ym}
0 = \frac{1}{\sqrt{-g}} \partial_\mu (\sqrt{-g}F^{a \mu \nu}) + \epsilon^{abc} A^b_\mu F^{c \mu \nu}.
\ee
To get the equation of motion for $w_+(r,z)$, we add the Yang-Mills equations \eqref{ym} for $(a,\nu) = (1,x)$ and $(a,\nu) = (2,y)$. We find that to obtain consistent equations, with $w_+$ independent of $\phi$, requires that the winding numbers are related by $n_- = n_+ + 2$. This relation between the winding numbers is also seen in the ordinary Ginzburg-Landau theory of $p$-wave superconductors \cite{heeb99}, and is ultimately due to the presence of mixed gradient terms such as $(D_x + iD_y)^2$ in the full equations of motion. Choosing $n_\pm = \mp 1$, which we expect corresponds to the lowest energy solution, yields
\be
\begin{split}
\label{eom1}
0 =& \partial_z (g \partial_z w_+) + \frac{1}{2r} \partial_r (r \partial_r w_+) - \frac{1}{2} \partial_r \left[ \frac{1}{r} \partial_r (r w_-) \right]
\\ & - \frac{1}{2r} w_- \partial_r A^3_\phi - \frac{1}{r} A^3_\phi \partial_r w_- -\frac{1}{2r^2}(A^3_\phi)^2 w_-
\\ & + \left[ \frac{\Phi^2}{g} + w_-^2 - w_+^2 - \frac{(A^3_\phi - 1)^2}{2r^2} \right] w_+.
\end{split}
\ee
Similarly, the equations of motion for $w_-$ and the gauge fields are
\begin{align}
\begin{split}
\label{eom2}
0 =& \partial_z (g \partial_z w_-) + \frac{1}{2r} \partial_r (r \partial_r w_-) - \frac{1}{2} \partial_r \left[ \frac{1}{r} \partial_r (r w_+) \right]
\\ & + \frac{1}{2r} w_+ \partial_r A^3_\phi + \frac{1}{r} A^3_\phi \partial_r w_+ -\frac{1}{2r^2}(A^3_\phi)^2 w_+
\\ & + \left[ \frac{\Phi^2}{g} + w_+^2 - w_-^2 - \frac{(A^3_\phi + 1)^2}{2r^2} \right] w_-
\end{split}
\\ 0 =& \partial^2_z \Phi + \frac{1}{r g} \partial_r (r \partial_r \Phi) - \frac{2}{g}(w_+^2 + w_-^2)\Phi \\
\begin{split}
0 =& \partial_z (g \partial_z A^3_\phi) + r \partial_r \left( \frac{1}{r} \partial_r A^3_\phi \right) + r \partial_r (w_-^2 - w_+^2)
\\ &+ (w_+ + w_-) \partial_r A^3_\phi + r (w_+ + w_-)(w_-^2 - w_+^2).
\end{split}
\end{align}
These equations can be solved numerically, subject to appropriate boundary conditions. At the boundary of AdS at $z=0$, the fields have the limiting forms
\be
\begin{split}
w_\pm =& \left< {\cal O}_\pm \right> z + \ldots
\\ \Phi =& \mu - \rho z + \ldots
\\ B (r) =& \frac{1}{r} \partial_r A^3_\phi,
\end{split}
\ee
where $\sqrt{ \left<{\cal O}_\pm \right> }$ are interpreted as the two components of the superconducting order parameter (the square root is necessary since ${\cal O}_\pm$ has mass dimension 2, whereas the superconducting order parameter should have mass dimension 1), $\mu$ is the chemical potential, $\rho$ is the charge density, and $B(r)$ is the magnetic field. In our numerical solution, we specify the value of $\mu$, as well as the conditions $w_\pm (z=0)=0$ and $\partial_z A^3_\phi = 0$. As discussed in Ref.\ \cite{domenech10}, this last condition is a Neumann boundary condition, which, unlike the more commonly used Dirichlet condition, allows for the presence of a dynamical gauge field in the boundary theory, and also, according to the AdS/CFT dictionary, leads to vanishing of the current operator in the boundary theory. At the horizon ($z=z_0$), we require $\Phi = 0$ and $A^3_\phi$ be regular. At the vortex core ($r=0$), the boundary conditions are $w_\pm = 0$, $A^3_\phi = 0$ and $\partial_r \Phi = 0$. Finally, far from the vortex core at the edge of our solution domain ($r=R$), we require $\partial_r w_\pm = 0$, $\partial_r \Phi = 0$ and $A^3_\phi = 1$. This last condition ensures that there is one quantum of magnetic flux passing through the vortex region \cite{domenech10}, with $\int d^2 r B(r) = 2 \pi$. To obtain our numerical solutions we have used the COMSOL 3.4 package \cite{comsol}.
\begin{figure}
\includegraphics[width=0.4\textwidth]{fig1.pdf}
\caption{(Top) Spatial profile of the $w_+$ (solid) and $w_-$ (dashed) components of the superconducting order parameter for an isolated vortex in a magnetic field at temperature $T/\mu=0.032$. (Bottom) Magnetic field profile for the same vortex configuration.
\label{vortex_profile}}
\end{figure}
Fig.\ \ref{vortex_profile} shows the spatial profile of the two components of the superconducting order parameter for an isolated vortex, through which a single quantum of magnetic flux penetrates the superconductor. It can be seen that, as the order parameter approaches its bulk value far from the vortex core, the $w_-$ component has a slightly smaller amplitude than the $w_+$ component. This is a consequence of the breaking of time-reversal symmetry, and the fact that the two components do not couple to the external field in the same way. The difference between the two components will grow as the field increases, however vortices will tend to proliferate at higher fields, so that the picture of an isolated vortex eventually ceases to be valid. The hump at $r \sim 0.2$ is an interesting feature that appears to be present at all temperatures. It may arise from the fact that the gradient terms in our theory are different from those in the usual Ginzburg-Landau theory, where such a hump is generally not present \cite{heeb99}.
The lower part of Fig.\ \ref{vortex_profile} shows the profile of the magnetic field near the vortex core. The exponential decay of the field with distance from the vortex core is a general property of superconductors and is also seen in the conventional Ginzburg-Landau theory. It is interesting to compare the size of the vortex core (the so-called ``coherence length") $\xi \sim 0.1$ to the penetration depth of the magnetic field $\lambda \sim 3$ \cite{lambda_note}. The ratio of these quantities defines the Ginzburg-Landau parameter $\kappa \equiv \lambda / \xi \sim 30$. The fact that $\kappa \gg 1$ means that the holographic $p$-wave superconductors are strongly type II, and are therefore expected to exhibit a vortex lattice solution near an upper critical magnetic field $H_{c2}$. Most of the superconductors that attract widespread theoretical interest, including the high-temperature cuprate superconductors, fall in this regime.
We begin our discussion of the superconducting properties near the upper critical field by considering the case of an ordinary, non-holographic $p$-wave superconductor in a magnetic field. In this case the ground state that is realized is known to depend on the coefficients of the kinetic terms in the free energy. For a two-dimensional superconductor with two complex components, the kinetic terms allowed by symmetry are $K_1 (D_i \eta_j)^* D_i \eta_j$, $K_2 (D_i \eta_i)^* D_j \eta_j$ and $K_3 (D_i \eta_j)^* D_j \eta_i$, and the ground state that is realized depends on the values of these coefficients \cite{sundaram89}. For $(K_2+K_3)/K_1 < 0$ the free energy is not bounded from below and the theory is unstable. (This is true at least for the linearized version of the theory. The theory may still be stable when higher order terms are included.) In the stable region, one of two possible ground states is realized. For $(K_2+K_3)/K_1 > 0$ and $K_2-K_3 > (K_2+K_3)^2/(2K_1+K_2+K_3)$ only one order parameter component is nonzero (e.g.\ $\eta_+ \sim \eta_1 + i \eta_2$ is nonzero if the field is in the $+ \hat z$ direction), and this component is in the lowest ($n=0$) Landau level. Such a state shall be denoted as $|0\rangle_+$. On the other hand, for $K_2-K_3 < (K_2+K_3)^2/(2K_1+K_2+K_3)$, both order parameter components are nonzero, and there is a mixture of $n=0$ and $n=2$ Landau levels (e.g. the state is of the form \ $c_+ |2\rangle_+ + c_- |0\rangle_-$ if the field is in the $+ \hat z$ direction, where the coefficients will generally depend on temperature and field). Following Ref.\ \cite{sundaram89}, we call these, respectively, the $A$ and $U$ phases. Writing out the Yang-Mills Lagrangian from Eq.\ \eqref{action} in terms of the fields $w_\pm$, it can be shown that the coefficients in our theory are $K_1 = 1$, $K_2 = 0$ and $K_3 = -1$, so it appears that we are on the boundary between the stable and unstable regions, and also--if we assume that the phase boundary remains unchanged for the holographic superconductor--on the boundary between the $A$ and $U$ phases described above. One can imagine changing the coefficients of these gradient terms by hand, thereby moving away from this critical point, but in that case we would no longer be dealing with pure Yang-Mills theory, which has the attractive feature of having no adjustable parameters.
Of course, the extent to which these results are relevant for holographic superconductors ought to be questioned. Since our ansatz leads to a Lagrangian in which there are no terms coupling gradients in the $z$ direction to those in the $xy$ plane, the criterion for stability should remain the same as in the non-holographic case. Also, by assuming that one order parameter component vanishes and employing separation of variables, it is shown below that the $A$ phase is a solution in the stable region of the phase diagram for the holographic case. Separation of variables, however, cannot be used to obtain the more complicated $U$ phase described above. It is possible that there is an analogue of this phase in the holographic superconductor, but because the $U$ phase in the non-holographic case is a complicated function involving multiple Landau levels, we expect that the same would be true in the holographic case. Characterizing such a state would most likely involve minimizing numerically the free energy of variational wavefunctions, and the problem of determining the exact form of such a state and comparing its energy with that of the $A$ phase is an interesting problem left for future study. It is important to note, however, that the criterion given above for distinguishing the $A$ and $U$ phases will not necessarily hold in the holographic case, so it is possible that the $A$ phase is in fact the unambiguous ground state for the particular Lagrangian that defines our theory. We therefore proceed pragmatically, assuming stability and the existence of a state with $w_-=0$ and following the original approach of Abrikosov \cite{abrikosov57}, which was also used in Refs.\ \cite{maeda10, zeng10}. (If $B<0$, the following discussion holds if we replace $w_+ \to w_-$.)
Since there will not be rotational symmetry for the vortex lattice as there was for the isolated vortex, we once again take all fields to be functions of all three spatial variables, with $A^a_\mu = A^a_\mu (x,y,z)$. In order to make notation more transparent and to allow for easier comparison with the existing literature on superconductivity, we now switch to more conventional notation in which the superconducting order parameters are represented by two complex scalar fields. Letting
\be
\begin{split}
\eta_{1,2} = A^1_{x,y} + iA^2_{x,y}
\\ \eta_\pm = \frac{1}{\sqrt 2} (\eta_1 \pm i \eta_2),
\end{split}
\ee
the Yang-Mills part of the action in Eq.\ \eqref{action} can be expressed as
\be
\begin{split}
\label{gl_action}
S^{YM} & = \frac{1}{2 q^2 \kappa^2} \int d^4x \bigg[ (\partial_z \Phi)^2 + \frac{1}{g} (\nabla \Phi )^2 + \frac{\Phi^2}{g} |\eta_+|^2
\\ & - g \left[ |\partial_z \eta_+|^2 + (\partial_z {\bf A})^2 \right] - \frac{1}{2} |D_x \eta_+ |^2
\\ & - \frac{1}{2} |D_y \eta_+ |^2 - \left( \partial_x A_y - \partial_y A_x + \frac{1}{2} |\eta_+|^2 \right)^2
\\ & + \frac{1}{2i} \left[ D_x \eta_+ ( D_y \eta_+)^* - ( D_x \eta_+)^* D_y \eta_+ \right] \bigg] ,
\end{split}
\ee
where we have assumed that we are in the $A$ phase described above and sufficiently close to the upper critical field that we can set the second order parameter component $\eta_- = 0$. Here and in what follows, the gradient operator is defined as $\nabla \equiv (\partial_x , \partial_y )$, and to simplify notation we have let $A^3_i \to A_i$.
As was done in Refs.\ \cite{maeda10, zeng10}, near the upper critical field, we can define $\epsilon \equiv (H_{c2} - H)/H_{c2}$ and expand the fields:
\be
\begin{split}
\Phi(x,y,z) &= \Phi^{(0)} + \epsilon \Phi^{(1)} + {\cal O} (\epsilon^2)
\\ A_{x,y} (x,y,z) &= A_{x,y}^{(0)} + \epsilon A_{x,y}^{(1)} + {\cal O} (\epsilon^2)
\\ \eta_+ (x,y,z) &= \epsilon^{1/2} \eta_+^{(1)} + \epsilon^{3/2} \eta_+^{(2)} + {\cal O} (\epsilon^{5/2}).
\end{split}
\ee
To leading order near $H_{c2}$, the electromagnetic fields are $\Phi = \mu (1-z/z_0)$, $A^3_y = x H_{c2}$ and $A^3_x = 0$. The higher order terms take into account the backreaction of the bosonic field on the electromagnetic fields. The magnetic field in the boundary theory is $H \equiv (\partial_x A_y - \partial_y A_x) |_{z=0}$. Taking the equation of motion for $\eta_+$ from Eq.\eqref{gl_action} and letting $\eta_+ (x,y,z) = m_+ (x,z;p) e^{ipy}$ gives
\be
\begin{split}
\label{eom_m}
0 =& \partial_z (g \partial_z m_+) + \frac{1}{2} \partial_x^2 m_+
\\ & + \left[ \frac{\Phi^2}{g} - \frac{1}{2} (H_{c2} x + p )^2 - \frac{3}{2} H_{c2} \right] m_+ ,
\end{split}
\ee
where we have neglected the term $\sim |m_+|^2 m_+$, since $m_+$ is small near $H_{c2}$. We note here that the vortex lattice solution for the $p$-wave superconductor was not obtained in Ref.\ \cite{zeng10} because the {\it gauge field} ansatz in that paper contained a complex phase factor, so that $A_\mu \sim e^{i p y}$. In our work, it is the complex bosonic field $\eta_+$ which is given the phase factor, which, in the language of the original gauge fields, corresponds to $A_\mu \sim e^{2 p y \tau^3}$. This is necessary to obtain the term $\sim (H_{c2}x+p)^2$ in Eq.\ \eqref{eom_m}. The distinction is important because, as we described above, the gauge field in this theory is real, with the role of the real and imaginary parts of the usual Ginzburg-Landau order parameter being played here by the $\tau^1$ and $\tau^2$ directions in $SU(2)$ space.
Taking advantage of the linearity of Eq.\ \eqref{eom_m}, we can again use separation of variables, letting $m_+ (x,z;p) = \rho(z)\gamma(x;p)$. We then obtain the following eigenvalue equations:
\begin{align}
0 =& -\partial_X^2 \gamma_n + X^2 \gamma_n - \lambda_n \gamma_n
\\ 0 =& \partial_z (g \partial_z \rho_n) + \left[ \frac{\mu^2}{g} \left( 1-\frac{z}{z_0} \right)^2 - \frac{H_{c2}}{2} (\lambda_n + 3) \right] \rho_n,
\label{rho}
\end{align}
where $X \equiv \sqrt{H_{c2}} (x+p/H_{c2})$. The first of these is just a harmonic oscillator equation, which is solved by the Hermite polynomials:
\be
\gamma_n (x;p) = e^{-X^2/2}H_n(X).
\ee
Here $n = 0,1,2,\ldots$ denotes the Landau energy level, and the corresponding eigenvalues are given by $\lambda_n = 2n+1$. The Abrikosov vortex lattice is given by a superposition of the lowest energy ($n=0$) solutions:
\be
m_+ (x,y,z) = \rho_0 (z) \sum_j c_j e^{i p_j y} \gamma_0 (x;p_j),
\ee
where the $c_j$ are coefficients that determine the structure of the vortex lattice. As shown in Refs.\ \cite{maeda10,zeng10,ge10}, the upper critical field $H_{c2}$ can be calculated at a given temperature by finding the highest field at which Eq.\ \eqref{rho} has a non-vanishing solution, indicating the presence of a superconducting condensate. The resulting phase diagram is shown in Fig.\ \ref{bc2}. We alert the reader to the characteristic upward curvature of $H_{c2}$. This curvature, stemming from Eq.\ (\ref{rho}),
is intrinsic to our theory and is thus reflective of the effects of field-induced correlations captured within the holographic approach.
\begin{figure}
\includegraphics[width=0.4\textwidth]{fig2.pdf}
\caption{Upper critical magnetic field for the $w_+$ component of the superconducting order parameter. The lower left part of the phase diagram is the superconducting region, and the upper right is the normal state. Note the upward curvature, which is
a direct consequence of the $z$-dependence in Eq.\ \eqref{rho}.
\label{bc2}}
\end{figure}
We next investigate the nature of the vortex lattice solution near $H_{c2}$. It is well known that the free energy in an ordinary ($s$-wave) Ginzburg-Landau theory is minimized when the vortex cores form a triangular lattice, and this has also been shown to be the case for a holographic $s$-wave superconductor \cite{maeda10}. Here we follow a similar approach to find the configuration that minimizes the free energy of the holographic $p$-wave superconductor. While our analysis is complicated somewhat compared to the $s$-wave case by the quartic term and the many gradient terms that appear in Eq.\ \eqref{gl_action}, we shall find that--just as in the $s$-wave case--the triangular vortex lattice minimizes the free energy.
Since all quantities are time-independent, the free energy is given by $\Omega = - S^{YM}_{\mathrm{OS}}/\int dt$, where $S^{YM}_{\mathrm{OS}}$ is the action evaluated with the fields $\eta_+$, $A_{x,y}$ and $\Phi$ on shell. The equation of motion for $\eta_+$ following from Eq.\ \eqref{gl_action} is
\be
\begin{split}
\label{eta_eom}
0 = & \partial_z ( g \partial_z \eta_+) + \frac{1}{2} (D_x + iD_y)(D_x - iD_y) \eta_+
\\ & + \frac{\Phi^2}{g} \eta_+ + (\partial_x A_y - \partial_y A_x) \eta_+ - \frac{1}{2} |\eta_+|^2 \eta_+ .
\end{split}
\ee
Multiplying this equation by $\eta_+^*$ and integrating over space, combining this result with Eq.\ \eqref{gl_action} gives the action with $\eta_+$ evaluated on shell:
\be
\label{omega}
\begin{split}
S^{YM} (\bar \eta_+) = & \frac{-1}{ 2 q^2 \kappa^2} \int d^4x \bigg[ - (\partial_z \Phi)^2 - \frac{1}{g} (\nabla \Phi)^2
\\ & + g (\partial_z {\bf A})^2 + (\partial_x A_y - \partial_y A_x )^2 - \frac{1}{4} |\bar \eta_+|^4 \bigg],
\end{split}
\ee
where integration by parts has been used, and a bar denotes fields evaluated on shell. We assume that the scalar field $\eta_+$ has compact support in the $(x,y)$ coordinates, so that the contributions from the boundaries $x = const.$ and $y=const.$ vanish when we integrate by parts. Furthermore, there is no boundary contribution from the horizon due to the regularity condition, and none from the AdS boundary at $z=0$ due to the boundary condition $\eta_+(z=0)=0$.
Similarly, the on-shell conditions for the fields $\Phi$ and $A_i$ can be calculated from Eq.\ \eqref{omega} and the result substituted back into the action, yielding the action with all fields evaluated on shell
\be
\begin{split}
S^{YM}_{\mathrm{OS}} = \frac{1}{ 2 q^2 \kappa^2} \bigg[ & \frac{1}{4} \int d^4x g |\bar \eta_+|^4
\\ & + \int d^3x \bar A_i \partial_z \bar A_i \bigg|_{z=0} \bigg].
\end{split}
\ee
Here there is no boundary contribution from the $\Phi$ terms upon integration by parts due to the compact support in the $(x,y)$ coordinates and the boundary condition $\Phi (z=0) = 0$. As discussed in Ref.\ \cite{maeda10}, the boundary contributions from the gauge fields at the horizon can also be ignored since they are independent of the field $\eta_+$, and our interest here is in finding the configuration of $\eta_+$ that minimizes the free energy. The boundary term at $z=0$ for the field $A_i$ does not in general vanish, however; and, according the AdS/CFT dictionary, the expectation value of the current operator in the boundary theory is proportional to $\partial_z A_i$.
The leading nonzero correction to the action is at ${\cal O}(\epsilon^2)$, so the free energy can be expressed as $\Omega \approx \Omega^{(0)} + \epsilon^2 \Omega^{(2)}$, with
\be
\label{omega2}
\begin{split}
\Omega^{(2)} = \frac{-1}{ 2 q^2 \kappa^2} \int d{\bf x} \bigg[ & \frac{1}{4} \int dz g | \eta^{(1)}_+|^4
\\ & + A^{(1)}_i \partial_z A^{(1)}_i \bigg|_{z=0} \bigg].
\end{split}
\ee
For simplicity we no longer use the bar to denote fields that are on shell. In order to evaluate Eq.\ \eqref{omega2}, we need to obtain an expression for $A_i^{(1)}$. At ${\cal O} (\epsilon)$, the equation of motion following from Eq.\ \eqref{gl_action} is
\be
\label{a_eom}
[ \partial_z (g \partial_z) + 2 \nabla^2 ] A^{(1)}_i = -j_i^{(1)} + \frac{1}{2} \epsilon_{ij} \partial_j |\eta_+^{(1)}|^2,
\ee
where we have defined
\be
\label{current}
j_i^{(1)} \equiv \frac{1}{2i} \left[ \eta_+^{(1)*} D_i^{(0)} \eta_+^{(1)} - (D_i^{(0)} \eta_+^{(1)})^* \eta_+^{(1)} \right] .
\ee
We have also chosen the gauge condition $\partial_i A_i = 0$ and applied it in deriving Eq.\ \eqref{a_eom}. It is important to note that Eq.\ \eqref{current} describes currents in the bulk theory, and is distinct from the boundary current operator alluded to above. Furthermore, since $\eta_+ (z,{\bf x}) \sim \gamma({\bf x})$, which describes the ground state of a simple harmonic oscillator, it is a simple matter to show \cite{degennes89} that $j_i^{(1)} = - \frac{1}{2} \epsilon_{ij} \partial_j |\eta_+^{(1)}|^2$. The antisymmetric symbol satisfies $\epsilon_{ji} = -\epsilon_{ij}$ and $\epsilon_{12}=1$.
The solution to Eq.\ \eqref{a_eom} can be expressed as
\be
\begin{split}
\label{a1}
A_i^{(1)}(z,{\bf x}) & = a_i ({\bf x})
\\ & - \int dz' d{\bf x}' G_B(z,z'; {\bf x} - {\bf x'}) j_i^{(1)}(z',{\bf x'}),
\end{split}
\ee
where $a_i ({\bf x})$ is the homogeneous part of the solution satisfying $\epsilon_{ij}\partial_i a_j = - H_{c2}$, and the Greens function $G_B$ satisfies
\be
\begin{split}
\label{gb}
[ \frac{1}{2} \partial_z (g \partial_z) + \nabla^2 ] G_B & (z,z'; {\bf x} - {\bf x'})
\\ & = - \delta(z-z') \delta({\bf x}-{\bf x'})
\\ G_B (z=0, {\bf x}) = \lim_{z \to z_0} & g(z) G_B (z,z'; {\bf x}) = 0.
\end{split}
\ee
We can now use Eq.\ \eqref{a1} to evaluate the second term in Eq.\ \eqref{omega2}:
\be
\begin{split}
\label{bdy_int}
& \int d{\bf x} A_i^{(1)} \partial_z A_i
\\ & = \int dz' d{\bf x'} d{\bf x} a_i ({\bf x}) \partial_z G_B (z,z'; {\bf x}-{\bf x'}) j_i^{(1)}(z',{\bf x'})
\\ & = \frac{ H_{c2}}{2} \int dz' d{\bf x'} d{\bf x} \partial_z G_B (z,z'; {\bf x}-{\bf x'}) |\eta_+^{(1)}(z',{\bf x'})|^2,
\end{split}
\ee
where in the second equality integration by parts and $\nabla G_B (z,z'; {\bf x}-{\bf x'}) = -\nabla' G_B (z,z'; {\bf x}-{\bf x'})$ were used. From the boundary condition in Eq.\ \eqref{gb}, the integral of $G_B$ can be performed:
\be
\int d{\bf x}G_B (z,z'; {\bf x}) = \int_0^{\mathrm{min}(z,z')} \frac{dz''}{g(z'')}.
\ee
Using this result along with Eq.\ \eqref{bdy_int}, we obtain the following expression for the free energy:
\be
\label{omega_2'}
\Omega^{(2)} = \frac{-1}{ 4 q^2 \kappa^2} \int dz d{\bf x} \bigg[ & \frac{1}{2} g | \eta^{(1)}_+|^4 + H_{c2} | \eta^{(1)}_+|^2 \bigg].
\ee
As pointed out in Ref.\ \cite{maeda10}, however, Eq.\ \eqref{omega_2'} is not our final result, since it depends on the normalization of $\eta_+^{(1)}$. This ambiguity in normalization is resolved by considering nonlinearity. From Eq.\ \eqref{eta_eom}, the equation of motion for $\eta_+^{(1)}$ is $\hat {\cal L} \eta_+^{(1)} = 0$, where we have defined the differential operator
\be
\begin{split}
\hat {\cal L} \equiv & \partial_z (g \partial_z) + \frac{1}{2}(D_x + iD_y)^{(0)}(D_x - iD_y)^{(0)}
\\ & + \frac{(\Phi^{(0)})^2}{g} + H_{c2}.
\end{split}
\ee
Also following from Eq.\ \eqref{eta_eom} is the equation of motion for $\eta_+^{(2)}$, which can be expressed as $\hat {\cal L} \eta_+^{(2)} = J$, where
\be
\begin{split}
J \equiv & \frac{i}{2}(A_x + iA_y)^{(1)}(D_x -iD_y)^{(0)} \eta_+^{(1)}
\\ & + \frac{i}{2}(D_x + iD_y)^{(0)} \left[ (A_x - iA_y)^{(1)} \eta_+^{(1)} \right]
\\ & +\left[ (\partial_x A_y - \partial_y A_x)^{(1)} - \frac{2}{g} \Phi^{(0)} \Phi^{(1)} + \frac{1}{2} |\eta_+^{(1)}|^2 \right] \eta_+^{(1)}.
\end{split}
\ee
With this equation of motion, we obtain the following identity:
\be
\begin{split}
\label{oc}
0 &= \int d^3 x \left[ \eta_+^{(1)*} J - \eta_+^{(1)*} \hat {\cal L} \eta_+^{(2)} \right]
\\ &= \int d^3 x \left[ \eta_+^{(1)*} J - \left( \hat {\cal L} \eta_+^{(1)} \right)^* \eta_+^{(2)} \right]
\\ &= \int d^3 x \eta_+^{(1)*} J,
\end{split}
\ee
where integration by parts along with the boundary condition $\eta_+ (z=0) = 0$ was used to obtain the second line. Again using integration by parts and $j_i^{(1)} = - \frac{1}{2} \epsilon_{ij} \partial_j |\eta_+^{(1)}|^2$ in Eq.\ \eqref{oc}, we obtain the condition
\be
\begin{split}
\label{oc2}
0 = & \int d^3x \bigg[ \frac{1}{4} |\eta_+^{(1)}|^4
\\ & - \left( \partial_x A^{(1)}_y - \partial_y A^{(1)}_x + \frac{2}{g} \Phi^{(0)} \Phi^{(1)} \right) |\eta_+^{(1)}|^2 \bigg].
\end{split}
\ee
In order to evaluate Eq.\ \eqref{oc2} we must determine the form of $\Phi^{(1)}$. At ${\cal O}(\epsilon)$, the equation of motion for $\Phi$ is
\be
( g \partial_z^2 + \nabla^2 ) \Phi^{(1)} = |\eta_+^{(1)}|^2 \Phi^{(0)}.
\ee
This equation has solution
\be
\label{phi1}
\begin{split}
\Phi^{(1)} (z,{\bf x}) = \int d{\bf x'} \int dz' & \frac{ \Phi^{(0)}(z') }{g(z')} G_t (z,z';{\bf x} - {\bf x'})
\\ & \times |\eta_+^{(1)} (z',{\bf x'})|^2,
\end{split}
\ee
where $G_t (z,z';{\bf x} - {\bf x'})$ is the Green function satisfying
\be
\begin{split}
\label{gt}
(g \partial_z^2 + \nabla^2) G_t (z,z';{\bf x} - {\bf x'}) =
\\ -g(z) \delta (z-z') \delta^{(2)} ({\bf x}-{\bf x'}).
\end{split}
\ee
In order to make further progress, we expand the Greens functions from Eqs.\ \eqref{gb} and \eqref{gt} in a basis of eigenfunctions:
\be
\begin{split}
\label{gt2}
G_t (z,z';{\bf x}) = \sum_\lambda \xi_\lambda (z) \xi^\dagger_\lambda (z') G_2 ({\bf x}, \lambda)
\\ -g(z) \partial_z^2 \xi_\lambda (z) = \lambda \xi_\lambda (z)
\\ \xi_\lambda (0) = 0 = \xi_\lambda (z_0),
\end{split}
\ee
and
\be
\begin{split}
\label{gb2}
G_B (z,z';{\bf x}) = \sum_\lambda \chi_\lambda (z) \chi^\dagger_\lambda (z') G_2 ({\bf x}, \lambda)
\\ - \frac{1}{2} \partial_z \left[ g(z) \partial_z \chi_\lambda (z) \right] = \lambda \chi_\lambda (z)
\\ \chi_\lambda (z=0) = \lim_{z\to z_0}g(z) \chi ' (z) = 0.
\end{split}
\ee
where the Greens function $G_2$ satisfies
\be
\label{g2}
(\nabla^2 - \lambda) G_2 ({\bf x},\lambda) = -\delta ({\bf x}).
\ee
Due to Eqs.\ \eqref{a1} and \eqref{phi1}, the free energy in this theory takes a nonlocal form, as opposed to the usual, non-holographic Ginzburg-Landau theory, which is completely local. This is due to the fact that the Ginzburg-Landau theory is a low energy effective expansion, whereas the AdS theory presented here retains the physics from all energy scales \cite{maeda10}. To get a local effective theory, we recognize that, in the long wavelength limit, $G_2 ({\bf x},\lambda)$ decays much more quickly than $|\eta_+^{(1)} ({\bf x})|^2$, so we can approximate
\be
\label{local}
\int d {\bf x'} G_2 ({\bf x} - {\bf x'}, \lambda) |\gamma ({\bf x'})|^2 \approx \frac{|\gamma ({\bf x})|^2}{\lambda}.
\ee
Using Eqs.\ \eqref{a1} and \eqref{phi1}, we can now give an explicit form of the condition Eq.\ \eqref{oc2} in the long wavelength limit:
\be
\begin{split}
\label{oc3}
0 & = \int d^3 x \bigg[ \left( \frac{1}{4} - \frac{2}{g}\Phi^{(0)} \Phi^{(1)} \right) |\eta_+^{(1)}|^4 - H_{c2} |\eta_+^{(1)}|^2
\\ &- |\eta_+^{(1)}|^2 \nabla^2 \int d^3x' G_B(z,z'; {\bf x} - {\bf x'}) |\eta_+^{(1)}(z',{\bf x'})|^2 \bigg]
\\ & \approx \int d^3 x \bigg[ \left( \frac{\rho^4}{4} - \frac{\alpha(z) \rho^2}{2} \right) |\gamma|^4 - H_{c2} \rho^2 |\gamma|^2 \bigg]
\end{split}
\ee
To obtain the first equality we have again used integration by parts and $\nabla' G_B(z,z'; {\bf x} - {\bf x'}) = -\nabla G_B(z,z'; {\bf x} - {\bf x'}) $, and the second equality gives the approximate form in the long wavelength limit, using Eqs.\ \eqref{gt2}-\eqref{g2}. We have also defined
\be
\alpha (z) \equiv \frac{4 \Phi^{(0)}}{g} \sum_\lambda \frac{ \xi_\lambda}{\lambda} \int dz' \frac{\Phi^{(0)}(z')}{g(z')} \rho^2 (z') \xi^\dagger_\lambda (z').
\ee
By combining Eqs.\ \eqref{oc3} and \eqref{omega_2'}, we can now give an approximate, local expression for the free energy density that is independent of the normalization of the order parameter:
\be
\begin{split}
\label{free_energy}
\frac{\Omega}{V} &= \frac{1}{V} \left[ \Omega^{(0)} + \epsilon^2 \Omega^{(2)} + \ldots \right]
\\ & \approx \frac{\Omega^{(0)}}{V} - \frac{2 (H_{c2}-H)^2 \left< \rho^2 \right> ^2 \left< 2 \rho^4 - \alpha \rho^2 \right>}{q^2 \kappa^2 \beta \left< \rho^4 - \alpha \rho^2 \right> ^2},
\end{split}
\ee
where $\left< \ldots \right>$ denotes spatial average, and the Abrikosov parameter is given by
\be
\beta \equiv \frac{\left< |\gamma |^4 \right>}{\left< |\gamma |^2 \right> ^2}.
\ee
Since the free energy density in Eq.\ \eqref{free_energy} is negative, minimizing the free energy corresponds to minimizing $\beta$. It is well known that the vortex lattice distribution that minimizes $\beta$ is the triangular vortex lattice, for which $\beta = 1.159$. This was also the result found for the holographic $s$-wave superconductor \cite{maeda10}.
In conclusion, we have shown the existence of a vortex solution in a holographic $p$-wave superconductor at low magnetic fields, as well as a vortex lattice solution near the upper critical magnetic field, $H_{c2}$. $H_{c2}$ exhibits a characteristic upward curvature, intrinsic to our theory, which reflects the effects of field-induced correlations captured by the holographic approach. The free energy was found to be minimized by the triangular vortex lattice. In the future it would be interesting to extend this theory to a BCS-like theory of fermions in AdS \cite{hartman10}, which would give insight into the possible types of $p$-wave pairing in holographic superconductors, as well as the tantalizing possibility of Majorana fermions, which are known to exist as bound states in the vortex cores of chiral $p$-wave superconductors.
\begin{acknowledgments}
We thank D.\ E.\ Kaplan, A.\ Salvio, I.\ Tolfree and Y.\ Wan for useful discussions. This work was supported by the Johns Hopkins-Princeton Institute for Quantum Matter, under Award No.\ DE-FG02-08ER46544 by the U.S. Department of Energy, Office of Basic Energy Sciences, Division of Materials Sciences and Engineering.
\end{acknowledgments}
|
\section{Introduction}
\label{sec:intro}
In the last few years proposals for new experimental realizations of
strongly-correlated many body systems emerged. Among them are
ultracold gases of atoms trapped in optical lattices
\cite{jaksch_cold_1998,greiner_quantum_2002,bloch_many-body_2008}
and light-matter systems \cite{hartmann_strongly_2006,
greentree_quantum_2006, angelakis_photon-blockade-induced_2007,
hartmann_quantum_2008,tomadin_many-body_2010}. The latter consist
of light modes which are confined in coupled cavity arrays. Due to
the finite overlap of their quantum mechanical wave functions,
photons are able to tunnel between adjacent cavities. Free photons
are non-interacting, however,
an effective
repulsive interaction
can be achieved by coupling photons to atoms or atomic-like structures
leading to interesting physical phenomena, which are subject to strong
correlations.
Different theoretical models have been proposed
to describe these effects.
In one scheme the atomic like structures are modelled by two level-systems, which leads to an interaction of the Jaynes-Cummings (JC) type \cite{jaynes_comparison_1963, greentree_quantum_2006}. Another scheme is based on four-level systems \cite{hartmann_strongly_2006} and takes advantage of electromagnetically induced transparency \cite{boyd_photonics:_2006}. Here, we focus on the first mentioned type of interaction. The physics of a single cavity at site $i$ is described by the JC Hamiltonian \cite{jaynes_comparison_1963}
\begin{equation}
\hat{H}^{JC}_i = \omega_c \, a_i^\dagger \, a_i^{\phantom{\dag}} + \epsilon \, \sigma_i^+ \, \sigma_i^-+g( a_i^{\phantom{\dag}}\,\sigma_i^+ + a_i^{\dagger}\,\sigma_i^- ) \;\mbox{,}
\label{eq:jc}
\end{equation}
where $\omega_c$ is the resonance frequency of the cavity, i.\thinspace{}e.\@\xspace, the frequency of the confined photons, $\epsilon$ is the energy spacing of
the two-level system, and $g$ is the atom-field coupling constant. The coupling between the atom and the photons is achieved by dipole interactions. The operator $a_i^\dagger$ ($a_i^{\phantom{\dag}}$) creates (annihilates) a photon at cavity $i$ and $\sigma_i^+$ ($\sigma_i^-$) is the raising (lowering) operator of the two-level system. The JC Hamiltonian conserves the particle number $\hat{n}_{i} = a_i^\dagger \, a_i^{\phantom{\dag}} + \sigma_i^+ \, \sigma_i^- $, which is a result of the rotating wave approximation \cite{haroche_exploringquantum:_2006}. The full system of coupled cavities is modelled by the Jaynes-Cummings lattice (JCL) Hamiltonian
\begin{equation}
\hat{H}^{JCL} = -t \sum_{\left\langle i,\,j \right\rangle} a_i^\dagger \, a_j^{\phantom{\dag}} + \sum_i \hat{H}_i^{JC} - \mu\,\hat{N}_p \;\mbox{,}
\label{eq:jcl}
\end{equation}
\begin{figure}
\centering
\includegraphics[width=0.35\textwidth]{jclfig}
\caption{ Schematic illustration of the JCL model in two dimensions. Red wavy arrows indicate photons, which are confined in optical cavities and interact with two-level systems (blue bubbles). }
\label{fig:jclfig}
\end{figure}
where the first term allows photons to tunnel between neighboring cavities $i$ and $j$. The restriction to nearest neighbors is indicated by the angle brackets $\langle \cdots \rangle$ around the summation indices. The hopping strength $t$ of the photons is given by the overlap integral of the photonic wave functions. The second term in \eq{eq:jcl} describes the JC physics of each cavity and the last term with the chemical potential $\mu$ controls the total particle number $\hat{N}_p = \sum_i \hat{n}_i$, which is a conserved quantity of the JCL Hamiltonian. An illustration of the two-dimensional JCL model is shown in \fig{fig:jclfig}.
Throughout this paper we use the dipole coupling $g$ as unit of energy. Under these considerations, the physics of the JCL model depends on three distinct parameters: the hopping strength $t$, the detuning $\Delta\equiv\omega_c-\epsilon$ and the modified chemical potential $\mu-\omega_c$.
The fundamental excitations present in the JCL system are termed polaritons. Polaritons are superpositions of both photons as well as two-level excitations. At zero hopping strength $t$ the JCL Hamiltonian reduces to the JC model with energies shifted by $-\mu \hat{n}_i$ and can thus be solved analytically \cite{jaynes_comparison_1963,hussin_ladder_2005,haroche_exploringquantum:_2006}. From this solution it follows that the energy which is necessary to add two polaritons to the cavity is larger than twice the energy which is necessary to add one polariton to the cavity, leading to a repulsive interaction between the particles. For small, non-zero hopping strength, the JCL system is in the so-called Mott phase, provided the polariton density is integer. In the case of light-matter systems the Mott phase can be regarded as a state of frozen light. For $t>0$ two energies are competing: The kinetic energy which is gained by the hopping process of the photons and the repulsive interaction between the particles. For $t$ larger than a critical hopping strength $t_c$, it is energetically favorable for the particles to delocalize on the whole lattice and to Bose condense in a state of zero momentum. In this parameter regime the JCL system is in the superfluid phase. The critical hopping strength $t_c$ defines the boundary of the quantum phase transition from the Mott to the superfluid phase.
Lately, there has been a great deal of research interest in understanding light-matter systems and in particular the JCL model. Most of the work has been devoted to study its ground state properties. The quantum phase transition from the Mott to the superfluid phase has been investigated by means of density matrix renormalization group \cite{rossini_mott-insulating_2007, rossini_photon_2008}, the variational cluster approach \cite{aichhorn_quantum_2008, knap_jcl_2010}, quantum Monte Carlo (QMC) \cite{zhao_insulator_2008}, and analytically by means of strong coupling perturbation theory \cite{schmidt_strong_2009, Schmidt_2010}. Results are also available on mean field level \cite{greentree_quantum_2006, koch_superfluid--mott-insulator_2009}; some signatures of the quantum phase transition have also been determined from exact diagonalization of small systems consisting of a few cavities \cite{angelakis_photon-blockade-induced_2007, makin_quantum_2008, irish_polaritonic_2008}. The spectral properties of the JCL model have been evaluated in Refs.~\cite{aichhorn_quantum_2008, pippan_excitation_2009, schmidt_strong_2009, knap_jcl_2010}.
In the present paper, we study in detail the polaritonic properties of the JCL model. This work extends our recent results published in Ref.~\cite{knap_jcl_2010} to two-dimensional systems. To be able to characterize polaritons, we first investigate the spectral properties of both particle species present in the JCL model, namely the photons as well as the two-level excitations. Based on this information we are able to introduce polariton quasiparticles as appropriate, wave vector, filling and band index dependent linear combinations of photons and two-level excitations. Furthermore we analyze the contributions of the individual constituents to the polariton quasiparticles. For completeness we present the boundary of the quantum phase transition from the Mott to the superfluid phase as well and compare our results to QMC results obtained in Ref.~\cite{zhao_insulator_2008}.
The remainder of this paper is organized as follows: in Sec.\@\xspace~\ref{sec:method} we present the numerical method we are using. Section\@\xspace~\ref{sec:qpt} contains the results for the quantum phase transition and Sec.\@\xspace~\ref{sec:sp} provides a detailed investigation of the spectral properties of both particle species. Polaritonic properties are analyzed in Sec.\@\xspace~\ref{sec:pol}. Finally, we conclude and summarize our findings in Sec.\@\xspace~\ref{sec:conclusions}.
\section{Variational cluster approach}
\label{sec:method}
We employ the variational cluster approach \cite{potthoff_variational_2003, koller_variational_2006} (VCA) as a numerical tool to study the quantum phase transition, spectral properties and polariton quasiparticles of the two-dimensional JCL model. In particular, VCA provides the single-particle Green's function $G(\ve k,\,\omega)$ of the physical system $\hat{H}^{JCL}$ and is based on the self-energy functional approach \cite{potthoff_self-energy-functional_2003-1, potthoff_self-energy-functional_2003} and the cluster perturbation theory \cite{snchal_spectral_2000, snchal_cluster_2002}. VCA has been previously applied to light-matter systems in Refs.~\cite{aichhorn_quantum_2008, knap_jcl_2010, knap_tcl_2010}.
The main idea of VCA is that the
self-energy $\mat \Sigma$ of the
physical system $\hat{H}^{JCL}$ is
approximated by the one of
a so-called reference system $\hat{H}^\prime$, which
shares its interaction part with $\hat{H}^{JCL}$ and is exactly
solvable. Typically, $\hat{H}^\prime $ is a
cluster decomposition of $\hat{H}^{JCL}$.
The ``optimal'' reference system is determined by requiring that
an appropriate functional $\Omega[\mat\Sigma]$ of the self-energy is stationary
(see \cite{potthoff_self-energy-functional_2003-1} for details).
$\Omega[\mat \Sigma]$ is the grand potential of the physical
system at the stationary point.
Additionally,
at the stationary point of $\Omega[\mat \Sigma]$, Dyson's equation is
recovered and thus the Green's function of the physical system can be
extracted there \cite{potthoff_self-energy-functional_2003-1}.
To evaluate $\Omega[\mat \Sigma]$ numerically, the self-energy is
parametrized by the single-particle parameters $\mat x$ of the
reference system $\hat{H}^\prime$. Due to this parametrization
the functional $\Omega[\mat \Sigma]$ becomes a function $\Omega(\mat
x)$, and is restricted to a smaller subspace of self-energies.
Physical quantities are evaluated at the stationary point of $\Omega(\mat x)$ with respect to the single-particle parameters $\mat x$. The accuracy of the results depends on the cluster size of the reference system, as correlations are taken into account exactly on the cluster level. Thus convergence on physical quantities is achieved by increasing the cluster size of the reference system.
The reference system $\hat{H}^\prime$ is solved by exact diagonalization, which we carry out by means of the band Lanczos method \cite{freund_roland_band_2000, aichhorn_variational_2006}. To evaluate $\Omega(\mat x)$ and the Green's functions $G(\ve k,\,\omega)$ we apply the bosonic $\mat Q$-matrix formalism \cite{knap_spectral_2010}. In our previous work on the one-dimensional JCL model we adapted the VCA procedure such that it provides the single-particle Green's function of both particle species, which involved as a subtlety the mapping of the two-level systems onto hard-core bosons. To summarize, we are able to extract the Green's function of photons $G^{ph}(\ve k,\,\omega)$ and the Green's function of two-level excitations $G^{ex}(\ve k,\,\omega)$. From the Green's function $G^{x}(\ve k,\,\omega)$, where $x$ stands either for photons ($ph$) or for two-level excitations ($ex$), we obtain the single-particle spectral function
\begin{equation}
A^x(\ve{k},\,\omega)\equiv-\frac{1}{\pi} \mbox{Im} \, G^x(\ve{k},\,\omega)
\label{eq:spectralfunction}
\end{equation}
and the density of states
\begin{equation}
N^x(\omega)\equiv \int A^x(\ve{k},\,\omega) \, d\ve{k} = \frac{1}{N} \sum_{\ve{k}} A^x(\ve{k},\,\omega)\;\mbox{.}
\label{eq:spe:dos}
\end{equation}
\section{Quantum phase transition}
\label{sec:qpt}
The two-dimensional JCL model exhibits a quantum phase transition from the Mott phase, where polaritons are localized within the cavities, to the superfluid phase, where polaritons delocalize on the whole lattice \cite{greentree_quantum_2006, aichhorn_quantum_2008, zhao_insulator_2008, schmidt_strong_2009}. We evaluate the phase boundary for zero detuning $\Delta=0$ by means of VCA from the minimal excitation energies of the system, see \fig{fig:pd}.
\begin{figure}
\centering
\includegraphics[width=0.35\textwidth]{pdc}
\caption{ Phase boundary of the two-dimensional JCL model for zero detuning $\Delta=0$ delimiting the first and the second Mott lobe from the superfluid phase. The phase boundary is evaluated for reference systems of size $L$. The inset shows the geometry of the 8 site cluster. The marks refer to parameters where spectral functions are evaluated. }
\label{fig:pd}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=0.48\textwidth]{sf}
\caption{ Photon spectral function $A^{ph}(\ve{k},\,\omega)$, first row [\fc{a}, \fc{e}], and density of states $N^{ph}(\omega)$, second row [\fc{b}, \fc{f}]. Two-level excitation spectral function $A^{ex}(\ve{k},\,\omega)$, third row [\fc{c}, \fc{g}], and density of states $N^{ex}(\omega)$, fourth row [\fc{d}, \fc{h}]. The spectral functions are evaluated for the parameters \fc{a}--\fc{d} $t=0.015$, $\mu-\omega_c=-0.75$, $\Delta=0$, marked with $\mat x$ and with the Roman numeral I in \fig{fig:pd}, and \fc {e}--\fc{h} $t=0.05$, $\mu-\omega_c=-0.81$, $\Delta=0$, marked with $\mat x$ and with the Roman numeral II in \fig{fig:pd}. Both parameter sets belong to the first Mott lobe.}
\label{fig:sf}
\end{figure}
Zero detuning means that the resonance frequency of the cavities is equal to the energy spacing of the two-level systems. We use the resonance frequency $\omega_c$ and the energy spacing $\epsilon$ of the two-level systems as variational parameters, i.\thinspace{}e.\@\xspace, $\mat x = \lbrace \omega_c,\,\epsilon \rbrace $, see Ref.~\cite{knap_jcl_2010} for details concerning the choice of the variational parameters. Results converge quickly with increasing cluster size $L$. We evaluated the phase boundary for clusters consisting of $L=2\times1$, $L=2\times2$ and $L=8$ sites, where the latter are arranged according to the inset of \fig{fig:pd}. In the case of $L=8$ site clusters, the numerical determination of the stationary point of $\Omega(\mat x)$ turns out to be highly non-trivial. The reason for this behavior is subject to further research, which is out of the scope of this work. The width of the Mott lobes decreases with increasing particle density, which results from the fact that the repulsive interaction itself depends on the particle density. The critical hopping strength $t_c^n$, which specifies the tip of the $n$th Mott lobe, is $t_c^1 \approx 0.055$ for the first lobe and $t_c^2 \approx 0.0047$ for the second lobe. The VCA result for $t_c^1$ is in good agreement with the QMC data from Ref.~\cite{zhao_insulator_2008}, where the authors obtained $t_c^1\approx0.052$ for the critical hopping strength of the first Mott lobe.
\section{Spectral properties}
\label{sec:sp}
Spectral functions $A^x(\ve{k},\,\omega)$ and densities of states $N^{x}(\omega)$ evaluated for both particle species are shown in \fig{fig:sf}.
The spectral properties have been evaluated for the identical variational parameter set $\mat x = \lbrace \omega_c,\,\epsilon \rbrace$ as in the case of the quantum phase transition, for reference systems of size $L = 2 \times 2$ and an artificial broadening $0^+ = 0.03$.
Spectral functions evaluated in the first Mott lobe consist of three bands. This can be understood from the zero hopping limit, where it is sufficient to consider a single cavity. The Hamiltonian then reduces to a block diagonal form, see for instance Refs.~\cite{jaynes_comparison_1963, haroche_exploringquantum:_2006} for a full solution of the single cavity problem. Each block corresponds to a specific polariton number, i.\thinspace{}e.\@\xspace, the states contributing to one block are $\ket{n,\,\downarrow}$ and $\ket{n-1,\,\uparrow}$, where the first quantum number denotes the number of photons and the second one states whether the two-level system is excited ($\uparrow$) or in its ground state ($\downarrow$). Diagonalizing the block yields the dressed states $\ket{n,\,\pm}$, where $\ket{n,\,-}$ is the ground state of the block and $\ket{n,\,+}$ its excited state. From these considerations it can be deduced that the single-particle excitations present in the spectral functions result from excitations from the ground state $\ket{n,\,-}$ to $\ket{n+1,\,-}$ or to $\ket{n+1,\,+}$. This yields two distinct particle bands, which we denote as $\omega_p^-$ and $\omega_p^+$, respectively. The same holds for the hole excitations, which corresponds to excitations from $\ket{n,\,-}$ to $\ket{n-1,\,-}$ or $\ket{n-1,\,+}$. We denote these bands as $\omega_h^-$ and $\omega_h^+$. However, for spectral properties evaluated in the first Mott lobe only one hole band exists, as the state $\ket{0,\,+}$ does not exist. This follows from the fact that in the zero hopping limit the block of zero particles consists only of one state $\ket{0,\,\downarrow}$.
The location of the bands is identical for both the photons and the two-level excitations, as they are connected by the dipole coupling. However, the weights differ significantly.
In particular, the particle band $\omega_p^-$ is much more pronounced in the photon spectral properties than in the two-level excitation spectral properties, whereas the opposite is true for the band $\omega_p^+$, which is almost not visible in $A^{ph}(\ve{k},\,\omega)$ but carries significant weight in $A^{ex}(\ve{k},\,\omega)$. The intensity of the hole band $\omega_h^-$ is similar in both cases.
\section{Polaritonic properties}
\label{sec:pol}
In this section, we investigate the polaritonic properties of the two-dimensional JCL model. To this end we follow Ref.~\cite{knap_jcl_2010}, and introduce polariton quasiparticle and quasihole creation operators as appropriate superpositions of photons and two-level excitations
\begin{align*}
p_{\alpha,\ve{k}}^\dagger &= \beta_p^\alpha(\ve{k})\,a_\ve{k}^\dagger + \gamma_p^\alpha(\ve{k})\,\sigma^+_\ve{k} \;\mbox{,} \\
h_{\alpha,\ve{k}}^\dagger &= \beta_h^\alpha(\ve{k})\,a_\ve{k} + \gamma_h^\alpha(\ve{k})\,\sigma^-_\ve{k} \;\mbox{.}
\end{align*}
The polariton quasiparticle states are defined as
\begin{align*}
\ket{\tilde{\psi}_{p,\ve{k}}^\alpha} &= \frac{p_{\alpha,\ve{k}}^\dagger \ket{\psi_0} }{ \sqrt{\bra{\psi_0} p_{\alpha,\ve{k}} \, p_{\alpha,\ve{k}}^\dagger \ket{\psi_0} }} \;\mbox{and} \\
\ket{\tilde{\psi}_{h,\ve{k}}^\alpha} &= \frac{h_{\alpha,\ve{k}}^\dagger \ket{\psi_0} }{ \sqrt{\bra{\psi_0} h_{\alpha,\ve{k}} \, h^\dagger_{\alpha,\ve{k}} \ket{\psi_0}}} \;\mbox{,}
\end{align*}
\begin{figure}
\centering
\includegraphics[width=0.48\textwidth]{qwBands}
\caption{ Photon amplitudes $\beta$ and two-level excitation amplitudes $\gamma$ of the polariton quasiparticle creation operators \fc{a}, \fc{c} $p_{-,\ve k}^\dagger$ corresponding to the band $\omega_p^-$ and \fc{b}, \fc{d} $p_{+,\ve k}^\dagger$ corresponding to the band $\omega_p^+$. In addition, the overlap $\lambda$ is shown in all panels (dashed line). The left column, \fc{a}--\fc{b}, corresponds to the parameters $t=0.015$, $\mu-\omega_c=-0.75$, $\Delta=0$ [see mark I in \fig{fig:pd}] and the right column, \fc{c}--\fc{d}, to $t=0.05$, $\mu-\omega_c=-0.81$, $\Delta=0$ [see mark II in \fig{fig:pd}]. }
\label{fig:qw}
\end{figure}
where polariton quasiparticle/quasihole operators are applied on the $N_p$-particle ground state $\ket{\psi_0}$. In the following arguments we neglect for simplicity the band index $\alpha$ and wave vector $\ve k$. The goal of this formulation is, to determine the coefficients of the linear combination $\beta$ and $\gamma$ such that the polariton quasiparticle states $\ket{\tilde{\psi}_{p/h}}$ describe best the true $N_p \pm 1$ particle system. Therefore the overlap of $\ket{\tilde{\psi}_{p/h}}$ with the true $(N_p\pm1)$-particle eigenstates has to be maximized with respect to the coefficients
\[\max_{\beta,\gamma} |\braket{\psi^{N\pm1}|\tilde\psi_{p/h}}|^2\;\mbox{.}\]
This condition yields a generalized eigenvalue problem \cite{knap_jcl_2010} of the form
\[\mat A \, \ve z = \lambda \, \mat S \, \ve z\;\mbox{,}\]
where $\mat A$ is a $2\times 2$ matrix containing the spectral weight of photons and two level excitations contributing to the specific wave vector and band index, $\mat S$ is the overlap matrix and $\ve z$ determines the coefficients $\beta$ and $\gamma$ up to a constant, which is specified by the conservation of the total spectral weight. The eigenvalue $\lambda$ determines the quality of the polariton picture as it corresponds to the normalized overlap specified above. Thus $\lambda$ is bound by $[0,\,1]$. The polariton amplitudes $\beta$ and $\gamma$ of the linear combination are shown in \fig{fig:qw} for the particle bands $\omega_p^{\pm}$.
For the band $\omega_p^-$ the polaritons can be almost described by photonic excitations as $\beta$ dominates over $\gamma$. In addition the polariton picture is well fullfilled as $\lambda \approx 1$. For the band $\omega_p^+$ the situation is reversed as $\gamma$ dominates over $\beta$. In this case the polariton picture is solely modestly fullfilled as $\lambda \approx 0.3$. Interestingly, for $\omega_p^-$ the amplitudes $\beta$ and $\gamma$ are of opposite sign, whereas for $\omega_p^+$ they are of the same sign. This behavior is already observed in the limit of a single JC cavity and is a result of the special structure of the eigenvectors \cite{knap_jcl_2010}. However, as compared to the single-cavity limit, in the coupled-cavity system the modulus of the amplitudes is significantly altered. Finally, in both cases $\omega_p^+$ and $\omega_p^-$, the polariton amplitudes $\beta$ and $\gamma$ depend slightly on the wave vector $\ve k$, which is also a pure lattice effect.
\section{Conclusions}
\label{sec:conclusions}
In the present paper, we investigated the polaritonic properties of the two-dimensional Jaynes-Cummings lattice model by means of the variational cluster approach. First, we evaluated the phase boundary delimiting the Mott from the superfluid phase and compared the critical hopping strength which determines the position of the Mott lobe tip with quantum Monte Carlo calculations. We found good agreement between the two approaches. In addition, we evaluated spectral functions and densities of states for both particle species present in the Jaynes-Cummings lattice model, namely photons as well as two-level excitations. Spectral functions generally consist of four bands, two particle bands and two hole bands. An exception are spectral functions evaluated for parameters of the first Mott lobe, which contain two particle bands but only one hole band. Furthermore, we investigated in detail the difference in the spectral weight of photons and two-level excitations. Based on the spectral properties of both particle species, we were able to introduce polariton quasiparticles as wave vector, band index and filling dependent superpositions of photons and two-level excitations. The amplitudes of the linear combinations determine the character of the polaritonic quasiparticles. Moreover, the amplitudes depend on the wave-vector, which is a pure lattice effect.
\section{Acknowledgments}
We made use of parts of the ALPS library \cite{albuquerque_alps_2007} for the implementation of lattice geometries and for parameter parsing.
We acknowledge financial support from the Austrian Science Fund (FWF) under the doctoral program ``Numerical Simulations in Technical Sciences'' Grant No. W1208-N18 (M.K.) and under Project No. P18551-N16 (E.A.).
|
\section{Introduction}
Ram pressure (and related processes) by the intracluster medium (ICM) can remove a galaxy's gas (Gunn \& Gott 1972). This process has been observed in various stages; for example, Vollmer (2009) separates Virgo galaxies into pre-peak, peak, and post-peak ram pressure groups. The amount of time that a galaxy has been stripped can be estimated using the length of the observable tail and the velocity of the galaxy (e.g. Oosterloo \& van Gorkom 2005; Sun et al. 2006). This calculation is uncertain due to difficulties in determining the three dimensional galaxy velocity. Another assumption implicit in this calculation is that the observed tail provides the true length of the stripped gas. In fact, tails have been observed in \ion{H}{I}, H$\alpha$, and X-ray emission, although never all from the same tail (e.g. Oosterloo \& van Gorkom 2005; Koopmann et al. 2008; Kenney et al. 2008; Yoshida et al. 2002; Yoshida et al. 2004a,b; Sun et al. 2007; Sun et al. 2006; Sun et al. 2010; Machacek et al. 2006; Sun \& Vikhlinin 2005). The lengths of tails observed in different wavelengths can be quite different; for example the \ion{H}{I} tail of NGC 4388 is nearly three times as long as the observed H$\alpha$ tail (Oosterloo \& van Gorkom; Yoshida et al. 2002). Another method used to calculate the age of a tail is to use the estimated survival time of H$\alpha$, as in Gavazzi et al. (2001). However, it is still not clear what dictates cloud survival or even what conditions are necessary to produce the various types of emission (H$\alpha$, X-ray, and \ion{H}{I}). Can all three types of emission coexist? What physical processes dominate the heating and mixing of stripped gas into the ICM? These processes include: turbulent mixing, which can generate intermediate temperature and density gas at constant pressure; shock heating, which heats the ISM; radiative cooling, which can lead to recompression of heated gas, and heat conduction, which can evaporate small clouds.
In this work we focus on answering these questions by simulating gas stripping and comparing our simulated tail to a single observed stripped galaxy, ESO 137-001, which has been studied observationally in some detail. ESO 137-001 is in A3627, which is the closest massive cluster (z=0.0163, $\sigma_{radial}$ = 925 km s$^{-1}$ and kT = 6 keV), similar to Coma and Perseus in mass and galaxy content (Sun et al. 2009 and references therein). ESO 137-001 is a small (0.2L$_*$; Sun et al. 2006), blue emission-line galaxy (Woudt et al. 2004), that is $\sim$200 kpc from the center of the cluster in projection. Because its radial velocity is close to the average velocity of A3627 (Woudt et al. 2004; Woudt et al. 2008), most of its motion is likely in the plane of the sky, and therefore the stripping process is seen edge-on. Sun et al. (2006) found a $\sim$70 kpc X-ray tail pointing away from the cluster center using \textit{Chandra} and XMM-\textit{Newton} data. Sun et al. (2007) then discovered a 40 kpc H$\alpha$ tail with over 30 emission-line regions extending through the length of the H$\alpha$ tail, and concluded that the emission-line regions are giant \ion{H}{II} regions. In a recent follow-up paper, Sun et al. (2009) used deep \textit{Chandra} data and \textit{Gemini} spectra to characterize the X-ray tail and \ion{H}{II} regions in detail. They found a narrower secondary X-ray tail with a similar length. They also confirmed that 33 emission-line regions are \ion{H}{II} regions, with the furthest seven regions beyond the tidally-truncated halo of 15 kpc that is calculated in Sun et al. (2007) using simulations by Gnedin (2003). In addition to these distinct \ion{H}{II} regions, they find diffuse H$\alpha$ emission.
Vollmer et al. (2001) searched for \ion{H}{I} in A3627, and did not detect any \ion{H}{I} in or around ESO 137-001 with a limiting column density of 2 $\times$ 10$^{20}$ cm$^{-2}$ and a resolution of 15". In fact, of the $\sim$80 galaxies identified by Woudt et al. (1998) in their search region, Vollmer et al. (2001) detected only 2 in \ion{H}{I}, finding that the \ion{H}{I} detection rate in A3627 is similar to that in Coma.
Sivanandam et al. (2009) observed ESO 137-001 with IRAC and IRS on \textit{Spitzer}. The IRS data extended to 20 kpc from the galaxy along the X-ray tail, and warm ($\sim$160 K) molecular Hydrogen was detected throughout the length of the observed region. The observed region contains $\sim$2.5 $\times$ 10$^7$ M$_\odot$ warm H$_2$ gas. They also identify star-forming regions using 8 $\mu$m data, which coincide with H$\alpha$ emitting regions.
There has been a substantial amount of theoretical work investigating ram pressure stripping in general (e.g. Schulz \& Struck 2001; Quilis, Bower \& Moore 2000; Roediger \& Br\"uggen 2008, Kronberger et al. 2008; Kapferer et al. 2009) -- see Tonnesen \& Bryan (2009, 2010; hereafter TB09 and TB10) for a more detailed discussion. There have also been simulations designed to predict or interpret observational characteristics of ram pressure stripped tails and the remaining disks (e.g. Vollmer et al. 2005, 2006, 2008), but detailed, quantitative predictions of all three observational probes have been missing to date (\ion{H}{I}, diffuse H$\alpha$, and X-ray emission).
In our previous work (TB10), we ran a set of high resolution simulations (about 38 pc resolution, which is small enough to marginally resolve giant molecular clouds) to understand how a multiphase ISM could affect the survival and structure of ram pressure stripped gas. We focused on how density fluctuations that are observed in the multiphase ISM of galaxies can affect gas tails. Including radiative cooling allowed us to estimate the density of and emission from \ion{H}{I}, H$\alpha$, and X-ray gas separately. We found that both the morphology and velocity structure of our tails agreed with observations of long gas tails (e.g. Oosterloo \& van Gorkom 2005). Our simulations also resulted in observable amounts of \ion{H}{I} and H$\alpha$ emission. However, the X-ray tail had a low surface brightness, which we attributed to the low pressure of the surrounding ICM. In this paper we will use the same method as in TB10, but have chosen ICM parameters comparable to the ICM around ESO 137-001. By focusing on the level of agreement between our simulations and the observations of ESO 137-001 we will be able to discuss the importance of physical mechanisms such as heat conduction, as well as predict the conditions under which \ion{H}{I}, H$\alpha$, and X-ray emission are produced in stripped tails.
The paper is structured as follows. After a brief introduction to our methodology, we provide the characteristics of our simulations and our method of producing simulated observations (\S 2). We then (\S 3) present our results, specifically focusing on the comparison with ESO 137-001. In \S 4 we discuss the broader implications of our simulation, and discuss our choice of radiative cooling floor and resolution in \S 5.1-2. Finally, we conclude in \S 6 with a summary of our results and predictions for observers.
\begin{table}
\begin{center}
\caption{Galaxy Stellar and Dark Matter Constants\label{tbl-const4}}
\begin{tabular}{c | c}
\tableline
Variable & Value\\
\tableline
M$_*$ & $1 \times 10^{11}$ M$_{\odot}$ \\
a$_*$ & 4 kpc \\
b$_*$ & 0.25 kpc\\
M$_{bulge}$ & $1 \times 10^{10}$ M$_{\odot}$ \\
r$_{bulge}$ & 0.4 kpc \\
r$_{DM}$ & 23 kpc \\
$\rho_{DM}$ & $3.8 \times 10^{-25}$ g cm$^{-3}$ \\
\tableline
\end{tabular}
\end{center}
\end{table}
\section{Methodology}
\label{sec-methodology}
We use the adaptive mesh refinement (AMR) code {\it Enzo} (Bryan 1999; Norman \& Bryan 1999; O'Shea et al. 2004). Our simulated region is 311 kpc on a side with a root grid resolution of $128^3$ cells. We allow an additional 6 levels of refinement, for a smallest cell size of 38 pc. We refine our simulation based on the local gas mass, such that a cell was flagged for refinement whenever it contained more than about $2 \times 10^4$ $M_{\odot}$. We found that this refined most of the galactic disk to 38 pc resolution by the time the wind hit; dense clumps in the wake were also refined to 38 pc resolution, while more diffuse components had lower resolution.
The simulation includes radiative cooling using the Sarazin \& White (1987) cooling curve extended to low temperatures as described in Tasker \& Bryan (2006). To mimic effects that we do not model directly (such as turbulence on scales below the grid scale, UV heating, magnetic field support, or cosmic rays), we cut off the cooling curve at a minimum temperature $T_{\rm min}$ so that the cooling rate is zero below this temperature. In the simulations described here we use either ${T_{\rm min}} = 8000$ K, or ${T_{\rm min}} = 300$ K. Both allow gas to cool below the threshold for neutral Hydrogen formation. In TB09, we found that the minimum temperature affected the range of masses and sizes of clouds forming in the disk, and the distribution of clouds throughout the disk, although the range of gas densities was very similar. The ${T_{\rm min}} = 300$ K case resulted in a more fragmented disk whose small clouds took longer to strip. Therefore the timescales for gas stripping differed, although the total amount of gas lost was similar. In the simulations presented in this paper, we find that the cooling floor affects the amount of fragmentation in the disks, but not the mass-distribution of gas densities. The larger radii of the remaining gas disks in the two ${T_{\rm min}} = 300$ K runs are due to the survival of dense clouds in the outer disk (Figures \ref{fig:xrayproj}, \ref{fig:haproj}, and \ref{fig:hiproj}).
In TB10, we found that the structure of the wake also depends somewhat on ${T_{\rm min}}$. Since the publication of TB10, we found and corrected an error in our implementation of the cooling rate that resulted in a slight shift in the peak of the cooling curve around $10^4$ K, but did not significantly affect cooling at low and high temperatures (this affected only the ${T_{\rm min}} = 300$ K simulation). Tests showed that this had only a small impact on the dynamics of the flow, and on the predicted \ion{H}{I} and X-ray measures, but did strongly affect the predicted H$\alpha$ emission, which is extremely sensitive to the gas temperature around $10^4$ K. This explains the enhanced H$\alpha$ emission in this paper for the ${T_{\rm min}} = 300$ K run as compared to TB10.
\begin{table}
\begin{center}
\caption{Gas Disk Constants\label{tbl-gconst4}}
\begin{tabular}{c | c }
\tableline
Variable & Value \\
\tableline
M$_{gas}$ & $1 \times 10^{10}$ M$_{\odot}$ \\
a$_{gas}$ & 7 kpc \\
b$_{gas}$ & 0.4 kpc \\
\tableline
\end{tabular}
\end{center}
\end{table}
\begin{table*}
\begin{center}
\caption{Runs summary\label{tbl-runs}}
\begin{tabular}{c | c | c | c | c | c}
\tableline
Run & v$_{\rm ICM}$ (km/s) & P$_{\rm ICM}$ (dyne/cm$^{2}$) & T$_{\rm ICM}$ (K)& P$_{\rm ram}$ (dyne/cm$^{2}$) & t$_{\rm proj}$ (Myr)\\
\tableline
T80vh & 1900 & 4.2 $\times$ 10$^{-11}$ & 8.3 $\times$ 10$^{7}$& 11.6 $\times$ 10$^{-11}$ & 85\\
T3vh & 1900 & 4.2 $\times$ 10$^{-11}$ & 8.3 $\times$ 10$^{7}$& 11.6 $\times$ 10$^{-11}$ & 75\\
T3vl & 1413 & 2.66 $\times$ 10$^{-11}$ & 7.3 $\times$ 10$^{7}$& 5.29 $\times$ 10$^{-11}$ & 110\\
Sun et al. 2010 & & 1.8 $\times$ 10$^{-11}$ & $\sim$7 $\times$ 10$^{7}$& & \\
\tableline
\end{tabular}
\end{center}
\end{table*}
Our galaxy model is the same as in TB10 and TB09, which used the spiral galaxy model described in Roediger \& Br\"uggen (2006). We list the model parameters for our galaxy in Tables~\ref{tbl-const4} and \ref{tbl-gconst4}. The stellar and dark matter components of the galaxy are static potentials. As in TB09 and TB10, our galaxy position and box size allows us to follow gas 200 kpc above (in the wind, or z, direction) the galaxy. To identify gas that has been stripped from the galaxy, we also follow a passive tracer which is initially set to 1.0 inside the galaxy (defined as gas that is above the ICM density) and $10^{-10}$ outside. In the following analysis, we will use a minimum tracer fraction of 0.25 to find gas stripped from the galaxy (as in TB10).
\subsection{Introduction to the Three Runs}\label{sec:ICM}
The galaxy initially evolves in a static, high-pressure ICM with $\rho=$ 8.7 $\times$ 10$^{-28}$ g cm$^{-3}$ and $T = $ 9.069 $\times$ 10$^6$ K ($\rho=$7.57 $\times$ 10$^{-28}$ g cm$^{-3}$ and $T = $ 1.04294 $\times$ 10$^7$ K for the slower wind case), to allow cool, dense gas to form in the galaxy (each of our three runs has about 3 $\times$ 10$^9$ M$_\odot$ of gas with densities at or above 10$^{-22}$ g cm$^{-3}$ when the wind hits the disk). This naturally generates a multiphase ISM (see Tasker \& Bryan 2006 and TB09 for more discussion of the ISM properties).
After 155 Myrs, we reset the boundary conditions to generate a constant ICM inflow along the inner z-axis, which is always face-on to the galaxy. In Table \ref{tbl-runs} we show the details of each of the three runs. The table includes the ICM parameters and the time after the wind has hit the galaxy at which the X-ray tail is 80 kpc long (this is how we choose the outputs to compare to the observations of ESO 137-001). `T80' indicates cooling to 8,000 K and `T3' indicates cooling to 300 K, while `vh' and `vl' indicate high and low velocity wind, respectively. See TB09 for other details regarding the general numerical setup, and TB10 for a discussion of the general impact of the cooling floor on the tail structure. In order to compare with both observations and our previous work, we use a face-on wind direction.
\subsection{Projections}\label{sec:projection}
\textit{Enzo} outputs the density and temperature of the gas in each cell. To transform these values into \ion{H}{I} column density and H$\alpha$ intensity, we used Cloudy, version 08.00 of the code last described by Ferland et al. (1998). Using a table of temperatures and densities, we calculated the hydrogen neutral fraction and H$\alpha$ emissivity. This table is then used to calculate the observational quantities for each cell, which are then summed to generate an image. This is described in detail in TB10, but briefly, we included CMB radiation, the cosmic ray background, bremsstrahlung radiation from the ICM and the 2005 version of the Haardt \& Madau (2001) $z=0$ metagalactic continuum, as implemented by Cloudy.
We chose to calculate the neutral fraction and H$\alpha$ emissivity for a plane-parallel gas cloud of width 100 pc. We selected this width because it loosely corresponds to the cell size of most of the gas in the tails, and accounts approximately for attenuation of the ionizing background radiation. If we assumed the radiative thin limit (by using a very small cloud size in Cloudy), it would somewhat decrease the amount of \ion{H}{I} we predict, and increase the H$\alpha$ emission for dense, low-temperature gas. We discuss the use of different cloud sizes in detail in TB10, however, in the simulations in this paper we find that using a 10 pc cloud does not significantly change our H$\alpha$ flux nor our \ion{H}{I} column densities because of the higher densities in our tail gas. This indicates that radiative transfer effects are not that important for this work.
To create X-ray surface brightness projections, we use a spectral lookup table that depends on temperature and density, assuming a constant metallicity of 0.3 solar, as computed using a Raymond-Smith code (Raymond \& Smith 1977), as updated in XSPEC (Arnaud 1996). The X-ray band we use is 0.5 keV to 2.0 keV, following Sun et al. (2006).
\section{Comparison to Observations}
In this section, we compare our simulated stripped tail to observations of ESO 137-001. We choose an output time at which the X-ray tail length is about 80 kpc in order to match the observations of Sun et al. (2010); this time is shown in Table \ref{tbl-runs}. We are using this comparison with the observations of ESO 137-001 to better understand the physics at work in producing X-ray, H$\alpha$, and \ion{H}{I} tails, so we have not tuned our simulation specifically to find an exact match to the galaxy and tail of ESO 137-001. While our ICM parameters are similar to those of the ICM near ESO 137-001, they are not the same. The final row in Table \ref{tbl-runs} displays the ICM parameters from the observations of Sun et al. (2010). We have not attempted to model the exact angle between the galaxy's disk and orbital motion (Woudt et al. (2008) find a position angle of 125$^\circ$), and instead use a face-on wind, and we are using a larger galaxy (Initially our galaxy has a radius of 26 kpc, and in the comparison projections the radius is about 15 kpc, while the 2MASS isophotal radius for ESO 137-001 is 6.1 kpc (Skrutskie et al. 2006)). In the following comparisons we will take note of the differences between our simulations and the observations of Sun et al. (2006; 2007; 2010).
\begin{figure*}
\includegraphics[scale=1.55,trim= 7mm 6mm 15mm 50mm,clip]{fig1a}
\includegraphics[scale=1.55,trim= 7mm 6mm 15mm 50mm,clip]{fig1b}
\includegraphics[scale=1.55,,trim= 7mm 6mm 15mm 50mm,clip]{fig1c}
\includegraphics[scale=0.55]{fig1d}
\caption{X-ray surface brightness maps of our three simulated galaxies, seen side-on. From left to right, we show T3vl, T80vh, T3vh. We have chosen the outputs with tail lengths of 80 kpc for our comparison to the observations of Sun et al. (2006; 2007; 2010). See discussion in Section \ref{xraycomp}}
\label{fig:xrayproj}
\end{figure*}
We have chosen the output time at which each simulation has an X-ray tail whose length is about 80 kpc. This means that the length of time that each galaxy has been stripped in the different simulations is not the same. In Table \ref{tbl-tempmass}, we list the amount of gas in the tail in three different temperature ranges. These masses include all of the gas above 10 kpc from the disk with tracer fractions of at least 25\%, so some of the gas included is at low density, as shown in Figure \ref{fig:rhot} (too low to be observable). Nevertheless, it can be very loosely stated that longer stripping times may result in more mixing of hot gas into the ICM, and in more dense gas being stripped and condensing into clouds in the tail. This can be seen by comparing T80vh or T3vl with T3vh (of course there are also other differences in the simulations that influence the tail gas, such as cooling floor and ICM pressure). However, we stress that in this paper we mainly focus on what causes a tail to be bright in X-ray and H$\alpha$ emission, so focus more on ICM pressure than the amount of time a galaxy has been stripped (although we do consider this timescale in our discussion of heat conduction).
\begin{table}
\begin{center}
\caption{Amount of Stripped gas at Different Temperatures\label{tbl-tempmass}}
\begin{tabular}{c | c | c | c}
\tableline
Run & T $<$ 10$^4$ K & 10$^4$ $<$ T $<$ 10$^5$ K & 7 $\times$ 10$^5$ $<$ T $<$ 4 $\times $10$^7$ K\\
& 10$^9$ M$_\odot$ & 10$^9$ M$_\odot$ & 10$^9$ M$_\odot$ \\
\tableline
T80vh & 1.4 & 1.6 & 6.0 \\
T3vh & 1.1 & 1.2 & 9.1 \\
T3vl & 4.2 & 2.9 & 5.9 \\
\tableline
\end{tabular}
\end{center}
\end{table}
\subsection{X-ray}
\subsubsection{Comparison to ESO 137-001}\label{xraycomp}
We first compare the X-ray characteristics of our simulations to observations. The X-ray surface brightness projections of our runs are shown in Figure \ref{fig:xrayproj}. As in Sun et al. (2006; 2010), we measure the X-ray emission between 0.5-2.0 keV. The surface brightness profiles along the tails are in rough agreement with that of ESO 137-001, which is bright near the disk and has a second bright region about 40 kpc from the disk before becoming less luminous to the end of the tail. Our model does not reproduce the exact X-ray surface brightness profiles: we do not have a model that includes both the bright emission near the disk and a surface brightness decrease across the entire tail before the second brightness peak.
\begin{table*}
\begin{center}
\caption{X-ray tail attributes\label{tbl-xtail}
\begin{tabular}{c | c | c | c | c}
\tableline
Run or & \textit{l $\times$ w} & L$_{0.5-2 \rm{keV}}$ & L$_{0.5-2}$ {\rm corrected}$^{\tablenotemark{a}}$ & T \\
Observation & (kpc $\times$ kpc)& (10$^{40}$ erg s$^{-1}$) & (10$^{40}$ erg s$^{-1}$) & (10$^7$ K)\\
\tableline
T80vh & 80 $\times$ 26 & 66.3 & 4.8 & 2.5\\
T3vh & 80 $\times$ 30 & 93.0 & 5.1 & 2.1 \\
T3vl & 80 $\times$ 30 & 41.9 & 3.6 & 1.4 \\
Sun et al. 2010 & 80 $\times$ 8, 80 $\times$ 7 & (8.3 $\pm$ 0.4) & (8.3 $\pm$ 0.4) & 0.93 $\pm$ 0.05 \\
\tableline
\end{tabular}
\tablenotetext{a}{The corrected luminosity multiplies the simulated X-ray luminosity by the ratio between the simulated and observed tail volumes and the ratio between the simulated and observed ICM thermal pressures (see text).}
\end{center}
\end{table*}
To make the X-ray projections, we adopted a minimum observable surface brightness of 7.1 $\times$ 10$^{-6}$ erg cm$^{-2}$ s$^{-1}$, which we estimated from the total luminosity of the observed tail and brightness profiles in Sun et al. (2010) (our cluster background is about 3.6 $\times$ 10$^{-6}$ erg cm$^{-2}$ s$^{-1}$). This results in a luminosity difference of an order of magnitude between the lowest and highest surface brightness features in all of our simulated tails, similar to that in the observed tail of ESO 137-001.
Our tails are narrow and show a nearly constant width along the entire tail, in agreement with the X-ray morphology reported in Sun et al. (2006; 2010). This is in contrast to simulations that do not include radiative cooling and produce flared tails (e.g. Roediger, Br\"uggen \& Hoeft 2006) However, it is clear that we do not have separated X-ray tails as in the observations. This is not surprising given the explanation of Sun et al. (2010) that the two tails likely result from the stripping of two spiral arms. As we discuss in detail in TB09, our disks fragment but do not form spiral arms.
In Table \ref{tbl-xtail}, we list some of the characteristics of our simulated X-ray tails to compare with the observed tail. To calculate the luminosity of the tail, we find the total energy emitted and subtract the background emission from the ICM. As the table demonstrates, the simulated tails are wider, more luminous, and have a higher average (emission-weighted) temperature than the (spectroscopic) temperature measured by Sun et al. (2010). We now address these differences.
The tail is most likely wider because we are modeling a large spiral galaxy, while ESO 137-001 is thought to be a $\sim$ 0.2L$_*$ galaxy, with a smaller galactic radius (Sun et al. 2006 and references therein). In fact, the entire volume of the tail in any of our three runs is nearly an order of magnitude larger than the tail of ESO 137-001 (we assume that the tails are cylindrical with the heights and diameters denoted in Table \ref{tbl-xtail}). The other difference is that our ICM pressure is also somewhat larger than calculated in Sun et al. (2010) by either a factor of 1.47 (T3vl) or 2.33 (T3vh and T80vh).
These differences in tail volume and ICM pressure impact the observables in two ways. First, as we discuss in TB10 and below, the compression of stripped gas by the ICM determines the temperature and density distribution of gas in the tail. Therefore, a higher ICM pressure results in higher-density hot gas ($T > 10^6$ K), which will have a higher X-ray emissivity. As we argue in the next section, this makes the X-ray luminosity proportional to the ICM pressure. Second, if we assume that the filling factor of X-ray emitting gas is the same in our simulations as in the tail of ESO 137-001, then the total luminosity is also directly proportional to the volume of the entire tail. The X-ray luminosity after applying these corrections is also shown in Table \ref{tbl-xtail} (assuming X-ray luminosity is directly proportional to both the ICM pressure and the volume of the tail). When we account for these differences, our measured luminosities are within about a factor of two of the X-ray luminosity of ESO 137-001.
As shown in Table~\ref{tbl-xtail}, there is also a factor of two difference between the temperatures of the simulated and observed X-ray tails; however, this is most likely due to the different ways in which this quantity is determined in simulations as compared to observations. We use luminosity-weighted temperatures, which tend to be higher than spectroscopic temperatures when there is a range of gas temperatures (see Mazzotta et al. 2004). Sun et al. (2010) discuss how the spectral fitting of the iron-L hump biases their temperatures low if there are significant emission components at \textit{k}T = 0.4 - 2 keV, which is certainly the case in our simulation and in the tail of ESO 137-001 (Sun et al. 2010; Sivanandam et al. 2009). Therefore, the agreement is probably considerably better than indicated in Table~\ref{tbl-xtail}.
\begin{figure*}
\begin{center}
\includegraphics[scale=0.8,trim= 1mm 90mm 10mm 75mm,clip]{fig2}
\end{center}
\caption{Contour plots showing the mass in gas at different densities and temperatures for the T3vl run on the left, ${T_{\rm min}} = 300$ K (from TB10) on the right. The contours are spaced by a factor of 10 in mass. For a gas cell to be included in the contour plot, it must have at least 25\% of its mass originating from the galaxy (based on the tracer fluid). The two curves denote lines of constant X-ray luminosity per mass (or emissivity per density). This figure illustrates that high ICM pressure produces an X-ray bright tail. See discussion in Section \ref{xrayexplain}}
\label{fig:rhot}
\end{figure*}
\subsubsection{What Makes a Tail X-ray Bright?}\label{xrayexplain}
In this section, we examine what lights up an X-ray bright tail, and explain why some observed tails, like ESO 137-001, exhibit X-ray emission, while others do not. In Figure \ref{fig:rhot} we show the mass-weighted distribution of density and temperature of gas in the wake that originated from the galaxy from our T3vl run (left) and from the ${T_{\rm min}} = 300$ K run from TB10 using the corrected cooling curve (right). The plots include all of the gas located between 10 kpc and 240 kpc above the disk that has at least 25\% of it's mass originating in the galaxy (as determined by the tracer fraction). We choose to highlight these two cases because the ICM pressure is the only difference between the two simulations (the thermal pressures differ by a factor of 15). As discussed in more detail in TB10, the ($T > 10^5$ K) gas in the wake is largely in pressure equilibrium and so falls roughly along a line of constant pressure. The impact of the higher ICM pressure can be seen as a shift to higher density and temperature in the left panel.
\begin{figure*}
\begin{center}
\includegraphics[scale=10.5,trim= 21mm 8.75mm 23.1mm 84.9mm,clip]{fig3a}
\includegraphics[scale=10.5,trim= 21mm 8.75mm 23.1mm 84.5mm,clip]{fig3b}\\
\includegraphics[scale=10.5,trim= 21mm 8.75mm 23.1mm 84.5mm,clip]{fig3c}
\includegraphics[scale=10.5,trim= 21mm 8.75mm 23.1mm 84.5mm,clip]{fig3d}
\end{center}
\caption{Projections of a very thin (0.3 kpc) slice from the T80vh run (images are 16 kpc $\times$ 19 kpc). The upper left panel shows X-ray surface brightness (red brightest), which demonstrates that X-ray emitting gas is associated with diffuse, mixed gas, but not necessarily material recently stripped from dense clouds in the tail. The lower left panel (temperature) demonstrates that the brightest X-ray emitting gas is associated with intermediate temperatures. Similarly, the upper right panel shows that H$\alpha$ emission is produced primarily at the edges of dense clouds. The lower right panel (gas surface density) demonstrates that the brightest H$\alpha$ is produced in high surface density clouds.}
\label{fig:slices}
\end{figure*}
In red we also plot two contours of constant luminosity per mass (or emissivity per density). The lower contour is at 10$^{-2}$ erg s$^{-1}$ g$^{-1}$. If the minimum observable X-ray surface brightness is 10$^{-5}$ erg s$^{-1}$ cm$^{-2}$, then in order for X-ray emission to reach this level, there must be a sufficient amount of gas at or above the red contour. More precisely, we need a column density of hot gas of 10$^{21}$ cm$^{-2}$ in order to produce observable emission (this corresponds to about 10$^4$ M$_\odot$ along a single line of sight through our 38 pc $\times$ 38 pc cells). In general, only the colder gas (T $\le$ 10$^4$ K) in the tail has these high surface densities (10$^{21}$ cm$^{-2}$), so we also plot an emissivity line which only requires a surface density of 10$^{20}$ cm$^{-2}$ in order to be observable (the upper 10$^{-1}$ erg s$^{-1}$ g$^{-1}$ contour). Gas that is hot enough to emit X-rays (hotter than $\sim$7 $\times$ 10$^5$ K) has this lower surface density in the tail (10$^{20}$ cm$^{-2}$).
This figure shows that gas at higher densities and lower temperatures than the ICM (but above $\sim$ 10$^6$ K) will be emitting the most strongly in X-rays. As galactic gas is stripped it either cools into clouds ($T < 10^5$ K) or is compressed to the ICM pressure and begins to mix with ICM gas. While the tail gas is at the high ICM pressure -- but before it is completely mixed with the ICM -- it will be X-ray bright. The mixing occurs along the line of constant pressure seen in Figure~\ref{fig:rhot}.
The separation between cold clouds and X-ray emitting gas is seen clearly in the left panel of Figure \ref{fig:slices}, which shows a thin slice from a detailed section above the disk. The cold clouds are black in this figure, indicating no X-ray emission, while the X-ray emitting gas is not confined to regions close to the dense clouds. Most of the X-ray bright cells in our simulation have between 70\% and 90\% of their gas originating from the galaxy. In the lower left panel of Figure \ref{fig:slices} we show the gas temperature, which when compared with the X-ray slice again highlights that while some mixing and radiative cooling can enhance the X-ray emission, but too much lowers the density of the hot gas to the point where the X-ray emissivity is low. The X-ray bright gas may have either been stripped as hot gas from the disk and slowly mixed, or stripped from the dense clouds in the tail and mixed into the ICM.
\begin{figure*}
\includegraphics[scale=1.55,trim= 7mm 6mm 15mm 50mm,clip]{fig4a}
\includegraphics[scale=1.55,trim= 7mm 6mm 15mm 50mm,clip]{fig4b}
\includegraphics[scale=1.55,,trim= 7mm 6mm 15mm 50mm,clip]{fig4c}
\includegraphics[scale=0.55]{fig4d}
\caption{H$\alpha$ intensity projections. These can be compared with the diffuse emission observed by Sun et al. (2007). From left to right: T3vl, T80vh, T3vh. See Section \ref{sec:ha} for discussion.}
\label{fig:haproj}
\end{figure*}
\subsection{H$\alpha$}\label{sec:ha}
Next, we turn to H$\alpha$ emission, shown in Figure~\ref{fig:haproj}. The minimum observable surface brightness that we adopt for these maps is 2 $\times$ 10$^{-18}$ erg s$^{-1}$ cm$^{-2}$ arcsec$^{-2}$ (see Sun et al. (2007) and references therein). As described earlier, we use Cloudy to determine the H$\alpha$ emissivity given a gas temperature and density.
We find, as shown in Figure \ref{fig:haproj}, highly structured, long tails of H$\alpha$ emission. Note that we do not include UV radiation from star formation or AGN (except from the metagalactic background, as described in section \ref{sec:projection}). Figure \ref{fig:slices} shows that H$\alpha$ emission peaks around the edges of cold clouds, which can be seen by comparing the upper right panel of H$\alpha$ emission with the lower right panel showing gas surface density.
\begin{figure*}
\includegraphics[scale=1.55,trim= 7mm 6mm 15mm 50mm,clip]{fig5a}
\includegraphics[scale=1.55,trim= 7mm 6mm 15mm 50mm,clip]{fig5b}
\includegraphics[scale=1.55,,trim= 7mm 6mm 15mm 50mm,clip]{fig5c}
\includegraphics[scale=0.55]{fig5d}
\caption{\ion{H}{I} column density projections at the maximum resolution of our simulation (38 pc). From left to right: T3vl, T80vh, T3vh. See Section \ref{sec:h1} for discussion.}
\label{fig:hiproj}
\end{figure*}
The ${T_{\rm min}}=8000$ K and ${T_{\rm min}} = 300$ K runs show a similar amount of H$\alpha$ emission. This is because, while the minimum temperature affects the temperature in the central regions of the clouds, it does not strongly affect the cloud edges in the simulation, whose characteristics are more determined by the interaction between the cloud and the ICM. We find that changing the cloud size parameter in our Cloudy run has only a small effect on the total H$\alpha$ flux (less than 10\% in all three runs). This is because most of our emission is produced by collisional processes rather than photoionization, and so is mostly dependent on the gas temperature, not the optical depth to ionizing photons (recall that we do include star formation in the simulation and so do not model HII regions within the tail).
\begin{table}
\begin{center}
\caption{H$\alpha$ tail attributes\label{tbl-hatail}}
\begin{tabular}{c | c | c | c}
\tableline
Run or & \textit{l $\times$ w} & f$_{\rm{H}\alpha}/10^{-14}$ & f$_{\rm{H}\alpha}/10^{-14}$ {\rm corrected}$^{\tablenotemark{a}}$ \\
Observation & (kpc $\times$ kpc) & (erg s$^{-1}$ cm$^{-2}$) & (erg s$^{-1}$ cm$^{-2}$)\\
\tableline
T80vh & 67 $\times$ 27 & 75.8 & 2.23\\
T3vh & 65 $\times$ 29 & 75.9 & 2.00 \\
T3vl & 80 $\times$ 30 & 97.9 & 1.96 \\
Sun et al. (2007) & $\sim$40 $\times$ 6 & 4.4 & 4.4\\
\tableline
\end{tabular}
\tablenotetext{a}{The corrected flux multiplies the simulated H$\alpha$ flux by the ratio between the simulated and observed tail volumes.}
\end{center}
\end{table}
Because we do not include star formation, we cannot compare our H$\alpha$ emission to the 33 \ion{H}{II} regions seen by Sun et al. (2007). We do compare the total flux from our tail to the observed diffuse H$\alpha$ emission in Table \ref{tbl-hatail}. Again, the volume of our diffuse tail is much larger than that observed by Sun et al. (2007), and we show the corrected H$\alpha$ flux by dividing by the volume ratio in the table. The difference in tail widths is from the difference in galaxy sizes, while the length of the observed tail is the minimum length of the diffuse emission because there is a bright star in the field (Sun et al. 2007). When we take the different volumes of the tail into account, our simulated H$\alpha$ flux is less than that observed from the tail of ESO 137-001, but in all three cases, the simulated H$\alpha$ flux is within a factor of 2.5 of the observed flux.
There are a number of possible explanations for our slightly lower predicted H$\alpha$ flux, including the possibility that some unresolved \ion{H}{II} regions are counted as diffuse flux in the observations. Other heating sources such as thermal conduction --- an effect we do not include in the simulations --- could give rise to more H$\alpha$ flux than we see in our simulations. We also find that numerical resolution plays a role in the total H$\alpha$ emission, as we will discuss in Section~\ref{sec-resolution}.
\begin{figure}
\begin{center}
\includegraphics[scale=1.55,trim= 7mm 6mm 15mm 50mm,clip]{fig6a}
\includegraphics[scale=0.55,trim= 0mm 5mm 0mm 0mm,clip]{fig6b}
\end{center}
\caption{\ion{H}{I} column density projection smoothed to the resolution of the observations by Vollmer et al. (2001) using ATCA. The only run with observable \ion{H}{I} in the tail using 30" resolution and a minimum column density of 2 $\times$ 10$^{20}$ is T3vl. See Section \ref{sec:h1} for discussion.}
\label{fig:ATCAV01}
\end{figure}
\subsection{\ion{H}{I}}\label{sec:h1}
In this section we consider \ion{H}{I} column density, comparing projections of our simulations to observations. In Figure \ref{fig:hiproj} we show the \ion{H}{I} column density at the best resolution of our simulation. As discussed in Section \ref{sec:projection}, we use Cloudy to determine the neutral fraction given a temperature and density (with the assumed metagalactic UV flux) and then apply that value in our projection routine.
In Figure \ref{fig:ATCAV01} we only show a projection comparable to the observations performed by Vollmer et al. (2001), using a resolution of 30" and a minimum column density of 2 $\times$ 10$^{20}$ cm$^{-2}$. We also correct for the difference in disk sizes by dividing our column densities by a factor of 3 (the ratio of X-ray tail diameters). Only the tail of T3vl would have been observed by Vollmer et al. (2001). This is because in the slower velocity case there is a smaller velocity difference between the clouds and the ICM wind and the ICM is slighty less dense (83\% of the ICM density in the higher velocity cases), which results in less cloud ablation and more high density gas (both a slightly higher maximum density of gas and slightly more gas at any particular density above $n \sim 1$ cm$^{-3}$). This difference between the non-detection by Vollmer et al. (2001c) and our prediction that T3vl would be detected implies that ESO 137-001 is most likely moving more quickly through the ICM than 1413 km s$^{-1}$. The \ion{H}{I} column densities in T3vh and T80vh are within a factor of 3 and 1.3 respectively below the maximum column density as determined by the non-detection of Vollmer et al. (2001).
Figure \ref{fig:slices} clearly shows that bright H$\alpha$ emission is produced only at the edges of dense neutral clouds. This spatial correspondence between neutral and H$\alpha$ emitting gas agrees with the observations of Sivanandam et al. (2009), who found molecular hydrogen in the tail to the farthest distance they could search--20 kpc. We predict that with a deeper observation \ion{H}{I} will be observed to at least 40 kpc (the length of the observed H$\alpha$ tail), and likely even farther (our shortest \ion{H}{I} tail is 65 kpc in T3vh), unless heat conduction is quite efficient.
\section{Discussion}
\subsection{Which Stripped Galaxies will have X-ray Tails?}
In this section, we comment on the likelihood of finding more X-ray tails. In order for a tail to have observable X-ray emission, it must be in a high pressure ICM. We can estimate the pressure necessary to produce bright X-ray emission simply by shifting our mass-weighted distribution of gas until the lowest mass contour lies along the upper 10$^{-1}$ erg s$^{-1}$ g$^{-1}$ contour. We choose this contour because in every simulation we have run, projections of surface density have shown that most of the tail has a column density of at least 10$^{20}$ cm$^{-2}$. Gas in the tail with higher surface densities tends to be in cold clouds (T $<$ 10$^4$ K) (see TB10), while X-rays are emitted by hot (T $>$ 10$^6$ K), diffuse gas with lower column densities (Figure \ref{fig:rhot}). This method results in a minimum ICM pressure of 9 $\times$ 10$^{-12}$ erg cm$^{-3}$.
In Table \ref{tbl-clusterrad} we show the cluster radius at which the ICM pressure falls below our minimum value. In general, we calculated the cluster radius using a $\beta$-model and constant temperature from the papers cited in column 3. There are three exceptions to this. First, the X-ray intensity contours in Figure 17 of Wang et al. (2004) denote the highest density ICM region (and are not spherically symmetric), which is the only region of A2125 with a high enough density to produce our minimum ICM pressure for X-ray bright tails. Second, we find the radius of Perseus with our minimum ICM pressure directly from Figure 9 in Ettori et al. (1998). Finally, we use the entropy measurements of Cavagnolo et al. (2009) to find that as close as 100 kpc from M87 the ICM pressure is below 9 $\times$ 10$^{-12}$ erg cm$^{-3}$. Note that these are first estimates generally using spherically symmetric cluster profiles, which oversimplify the structure of the ICM.
We can now comment on whether observed X-ray tails are in high pressure ICMs (defined here as at or above 9 $\times$ 10$^{-12}$ dyn cm$^{-2}$). The two X-ray tails in A3627, ESO 137-001 and ESO 137-002, are both within the high pressure ICM (Sun et al. 2007; Sun et al. 2010). However, Sun et al. (2010) find that using their spectroscopic temperatures, the tail gas is over-pressured with respect to the ICM. This might indicate that the spectroscopic temperature does not correctly model the tail temperature. We find that turbulent pressure is not strong enough to greatly increase the gas pressure in the tail (Figure \ref{fig:rhot} and TB10).
C153 is near the X-ray peak of A2125 (Wang et al. 2004). UGC 6697 in A1367 is at least 450 kpc from the X-ray peak of the cluster, outside of our calculated high-pressure radius. However, it is near an infalling subcluster and in a higher-density ICM than predicted using the $\beta$-model centered on the X-ray peak of A1367. Using the fit to the subcluster by Donnelly et al. (1998), Sun \& Vikhlinin (2005) calculated the surrounding pressure to be 7.9 $\times$ 10$^{-12}$ dyn cm$^{-2}$, very close to our minimum pressure for observable X-ray emission.
\subsection{Efficiency of Heat Conduction}
\begin{table}
\begin{center}
\caption{Where in clusters will tails be X-ray bright?\label{tbl-clusterrad}}
\begin{tabular}{c | c | c }
\tableline
Cluster & Outer radius (kpc) & Reference\\
\tableline
A3627 & 250 \textit{h$_{50}^{-1}$} & Bohringer et al. (1996); \\
& & Sun et al. (2010)\\
A2125 & 28.5 - 89 \textit{h$_{71}^{-1}$} & Wang et al. (2004)\\
A1367 & 274 \textit{h$_{50}^{-1}$} & Mohr et al. (1999) \\
CL 0024+0016 & 250 \textit{h$_{70}^{-1}$} & Zhang et al. (2005)\\
Coma & 613 \textit{h$_{50}^{-1}$} & Briel et al. (1992) \\
Perseus & 2000 \textit{h$_{50}^{-1}$} & Ettori et al. (1998) \\
Virgo (M87) & $<$ 100 \textit{h$_{70}^{-1}$} & Cavagnolo et al. (2009)\\
\tableline
\end{tabular}
\end{center}
\end{table}
Heat conduction, which we do not include in these simulations, could be important for the survival of cool clouds in the ICM and H$\alpha$ emission; however, it can be suppressed by magnetic fields (e.g. Vollmer et al. 2001). If it is an efficient way to transport heat from the ICM to cold, stripped gas, then the survival time of \ion{H}{I} clouds would be less than predicted in this paper, and the length of the tails would be shorter. We can estimate the efficiency of heat conduction by comparing an analytic calculation of the evaporation time of the most dense clouds in our simulations to the length of time we expect clouds have survived in order to produce the observations of ESO 137-001.
First, we estimate the evaporation time for a typical cloud if heat conduction is not suppressed, and we will define this calculated evaporation time as the time for cloud evaporation if heat conduction is 100\% efficient. We follow Cowie \& McKee (1977), as in Vollmer et al. (2001). In our clouds, we find that the mean free path for ions is comparable to or greater than the temperature scale length, so we need to use the saturated heat flux equations.
Evaporation time is proportional to $f^{-1} r_{\rm cloud}^{11/8} n_{\rm cloud}$ $T_{\rm ICM}^{-5/4} n_{\rm ICM}^{-11/8}$, where $f$ is the conduction efficiency relative to Spitzer. Solving for the evaporation time for each of our three runs, we find that using a cloud radius of 100 pc, and the maximum gas density in our tails at the time of our comparison, the evaporation times of our three runs are $\sim 4$ Myr for T80vh, $\sim 8$ Myr for T3vh, and $\sim 10$ Myr for T3vl. We choose to use the highest density found in our tails because we want to calculate the longest plausible evaporation time in order to find the most conservative estimates for the maximum efficiency of heat conduction. The ${T_{\rm min}}=300$ K runs have a maximum density in their tails of 2 $\times$ 10$^{-23}$ g cm$^{-3}$ and T80vh has a maximum density of 10$^{-23}$ g cm$^{-3}$.
We have shown that diffuse H$\alpha$ emission directly traces neutral clouds (Figure \ref{fig:slices}), which means that the dense clouds must not be entirely evaporated before they reach 40 kpc above the disk. We are also able to measure the velocity of gas in our tails (see TB10 for details), so can estimate the time it would take for a cloud to reach 40 kpc above the disk. We use generous estimates of 1200 km s$^{-1}$ for the high velocity cases and 900 km s$^{-1}$ for the lower velocity run in order to calculate the shortest time it would take these clouds to reach 40 kpc above the disk (which will maximize the value we find for the efficiency of heat conduction). Therefore, the efficiency of heat conduction must be less than 24\% (T3vl and T3vh), or 12\%(T80vh).
Once there is a deep observation of the \ion{H}{I} tail, we will be able to compare the length of ESO 137-001's tail to the lengths of our simulated tails to find a {\it minimum} efficiency of heat conduction. To do this we will use a similar argument to the one above; namely, compare the amount of time a cloud takes to reach its height above the disk to the evaporation time, assuming that heat conduction is responsible for destroying the clouds. It is important to consider our simulations in this calculation, because they show that even without heat conduction, \ion{H}{I} may not extend along the entire length of the X-ray tail.
The simulations also show the limitations of using a simple velocity and distance argument to calculate the length of time a galaxy has been stripped--using 900 km s$^{-1}$ and the length of the \ion{H}{I} tail we find that the galaxy in T3vl has been stripped for 87 Myr, while the actual time after the wind has hit the galaxy is 110 Myr. The high velocity cases have an even larger discrepancy between the actual stripping time and the time calculated using the \ion{H}{I} tail length.
Despite these caveats, we can make a very rough first estimate using Figure \ref{fig:ATCAV01} (T3vl) and the non-detection result of Vollmer et al. (2001). The \ion{H}{I} in our smoothed projection begins about 50 kpc above the disk. If the non-detection of Vollmer et al. (2001) means that there are no neutral clouds at this height above the disk, then we can calculate a minimum efficiency for heat conduction, which is 18.5\%. However, as we discussed above, we expect that a deeper observation will find \ion{H}{I} beyond this height above the disk.
\section{Numerical Considerations}\label{sec-numerical}
\subsection{Radiative Cooling Floor}\label{sec-cool}
As noted earlier, we used two different values for our radiative cooling floor (8000 K and 300 K) in order to explore in a simple way the potential impact of processes which we do not include in the simulation. We discuss the physical motivations for the two cooling floors in TB10. We find that the change in the minimum temperature makes only a relatively minor change in the morphology of the flow, and doesn't significantly change the resulting length of the tail. This also translates into only a relatively small difference in the predicted observables, with changes of less than 30\% in both X-ray and H$\alpha$ luminosities.\footnote{Note that this differs from the conclusion reached in TB10, where we found that the ${T_{\rm min}}=300$ K run predicted lower H$\alpha$ emission by more than an order of magnitude. This incorrect conclusion was reached because of the cooling curve error for that run discussed in section~\ref{sec-methodology}.} The biggest difference is in the survival of \ion{H}{I} clouds, with ${T_{\rm min}} = 300$ K predicting longer lived clouds. Still, the impact on \ion{H}{I} observations is quite small. The ratios of the three observables is also quite constant, generally to within 30\%, as can be see from an inspection of the right two panels of Figures~\ref{fig:xrayproj}, \ref{fig:haproj}, and \ref{fig:hiproj}, and from Tables~\ref{tbl-xtail} and \ref{tbl-hatail}.
\subsection{Resolution}\label{sec-resolution}
Resolution is most likely to affect the survival and structure of our dense clouds in the tail. The most direct results would be different amounts of neutral gas and H$\alpha$ emission. A number of examinations of the survival of clouds with a variety of physics included have been performed. Mellema et al. (2002) and Yirak et al. (2009) include radiative cooling. Fragile et al. (2004) also discuss how including self-gravity increases the lifetime of clouds. Nakamura et al. (2006) discuss the impact of using smooth cloud boundaries on the growth of instabilities. They (and Yirak et al. 2009) find that a low density gradient results in slower growth of instabilities, which can retard cloud destruction. Our most dense clouds have an analytically calculated destruction timescale due to turbulent viscous stripping (using eq. (22) in Nulsen 1982) of more than 1 Gyr, so resolution should not have a large impact over the timescales we consider in these simulations. We use the same cloud parameters as in the evaporation time calculation, although even with an order of magnitude lower density clouds ($\sim$ 10$^{-24}$ g cm$^{-3}$), the destruction timescale is still about 200 Myr, which is much longer than the time at which we make our projections. We use a relative velocity difference of 400 km s$^{-1}$ (although we do not show velocity plots in this paper, see TB10 for a detailed discussion of the tail velocity). However, lower density clouds in our simulated tails cool and are compressed by the ICM, which may instead be destroyed by the ICM wind if we had better resolution that resulted in a steeper density gradient at the cloud edge.
Although we do not perform a detailed resolution study of the runs in this paper, we do compare T3vl to a run with only 5 allowed levels of refinement, to a minimum cell size of 76 pc, a factor of two worse in resolution. We compare outputs with the same amount of gas stripped from the galaxy. The total X-ray luminosity of the lower resolution tail is $\sim$90\% of the luminosity of T3vl (i.e. only a 10\% change), while the total H$\alpha$ flux is only 28\% of T3vl. The total HI flux is also decreased by a similar amount (to 33\% of the value in T3vl). For both resolutions, there is a good correspondence between the relative HI and H$\alpha$ level and morphology; therefore, we argue that the ratio of HI to H$\alpha$ is more robust than the absolute level of either.
Both the small discrepancy in the X-ray luminosity and larger discrepancy in the H$\alpha$ flux can be explained by considering what gas is producing the emission. X-ray emission is produced by gas that is mixed with the ICM and is not localized near the dense clouds in the tail, and so is not very dependent on high resolution. One explanation for the slightly lower X-ray luminosity is that mixing happens more quickly and the gas is heated out of the X-ray bright regime.
Unlike the diffuse nature of the X-ray emitting gas, H$\alpha$ emission occurs at the edges of clouds. We find that in projection the range of H$\alpha$ intensities is the same as T3vl, meaning that the different resolution element size does not strongly affect the density and temperature at the edges of the dense clouds (and therefore does not strongly affect the H$\alpha$ emission). The main difference is that there are fewer dense clouds in the less-resolved wind. In fact, the total \ion{H}{I} column density in the low resolution tail is about 30\% of the total \ion{H}{I} column density in T3vl--very similar to the H$\alpha$ flux fraction. In projection, there are about half as many pixels with \ion{H}{I} column densities over 10$^{19}$ cm$^{-2}$. Since we expect the total \ion{H}{I} column density to also depend on the number of clouds along the line of sight of the projection, we find that there should be roughly 26\% as many clouds (by simply squaring the 51\% we find in the projection plane). Of course, clouds are more than a single pixel, so this is a rough approximation. However, this indicates that the amount of H$\alpha$ emission closely follows the amount of \ion{H}{I} column density, and likely the number of \ion{H}{I} clouds. There are fewer \ion{H}{I} clouds in the low resolution run because with lower resolution the maximum density is lower and therefore the clouds are destroyed and mixed more quickly into the ICM.
Deep, high resolution observations in \ion{H}{I} will allow us to determine how many and for how long dense clouds survive, which will point to one of these mixing scenarios.
\section{Conclusions}
We have run detailed galaxy simulations including radiative cooling and made comparisons to the observed tail of ESO 137-001 to understand the physical mechanisms at work in the ICM. We compare three cases in which we vary our cooling floor between 8,000 K or 300 K and the ICM parameters as shown in Table \ref{tbl-runs}. Our main conclusions are as follows:
1) The X-ray luminosity of our simulated tails are an excellent match to the X-ray luminosity of the tail of ESO 137-001. This suggests that we are correctly modeling the phase distribution of gas in the tail, and that the mixing of the hotter stripped ISM (T $>$ 10$^5$ K) is being accurately modeled in the simulations.
2) We find that bright X-ray emission depends upon a high surrounding ICM pressure, and find a minimum necessary pressure of 9 $\times$ 10$^{-12}$ erg cm$^{-3}$. This conclusion agrees well with the local environment in clusters where bright X-ray tails are observed.
3) We compare our H$\alpha$ fluxes to the total diffuse H$\alpha$ flux measure by Sun et al. (2007). As in our X-ray emission, we find an excellent match between our simulations and observations.
4) We predict that deeper observations will find \ion{H}{I} gas to at least 40 kpc above the disk, because diffuse H$\alpha$ emission directly traces neutral clouds (and has been observed to 40 kpc above the disk). The observations of molecular hydrogen by Sivanandam et al. (2009) strengthen this prediction. We also conclude that the mismatch between T3vl and Vollmer et al. (2001) indicates that the higher velocity cases better match the observations.
5) Using the fact that H$\alpha$ traces neutral gas and calculating the evaporation times of our simulated clouds (based on their sizes and densities), we calculated a maximum efficiency for heat conduction of 24\% (T3vh and T3vl) or 12\% (T80vh).
By modeling a simulation using a high ICM pressure, we have shown that X-ray emission can coexist with \ion{H}{I} and H$\alpha$ emission. We have also found that \ion{H}{I} and H$\alpha$ emission spatially coincide because H$\alpha$ is mostly produced at the edges of neutral clouds (Figure \ref{fig:slices}), while X-ray emission is generated in hot, diffuse gas that is mixing with the ICM. This is seen by the more even distribution in the tail (Figures \ref{fig:xrayproj} and \ref{fig:slices}), and by the fact that the X-ray tail can be longer than the \ion{H}{I} and H$\alpha$ tails.
Our excellent agreement with X-ray observations gives us confidence that we are correctly modeling the mixing, cooling and heating rate of X-ray emitting gas. This in turn means that heat conduction is not acting strongly to heat the diffuse gas in the tail, and that small scale turbulence (below our resolution scale) is not quickly mixing stripped gas. Both of these mechanisms would act to heat the gas in the tail out of the X-ray bright range and therefore lower the total luminosity, which would make our agreement to the observations worse. We cannot rule out the possibility that we have less gas mass at higher emissivities than in ESO 137-001.
We also find excellent agreement between our simulated H$\alpha$ flux and observations of diffuse H$\alpha$ emission. We robustly conclude that diffuse H$\alpha$ emission coincides with \ion{H}{I} gas, and we find that H$\alpha$ emission outlines \ion{H}{I} clouds in all of our simulations. The agreement between the total H$\alpha$ emission in our cases with different cooling floors underscores the fact that the edges of clouds, which are interacting with the ICM, emit the most strongly in H$\alpha$, not the central regions of the clouds, which are more likely to have radiatively cooled to the minimum allowed temperature. Although our agreement with observations indicates that we may also be correctly modeling the edges of dense clouds, we cannot dismiss the possibility that we are not fully resolving the cloud edges. In Section \ref{sec-resolution} we discuss how lower resolution results in fewer \ion{H}{I} clouds, and less H$\alpha$ emission. Therefore, we cannot rule out that properties of our simulations that have not converged at our current resolution--the number of clouds, the rate of the decline of density at the edge of the cloud and the smallest scale of turbulent heating--have combined with the lack of heat conduction in our simulations to result in an accidental, and incorrect, agreement with observations. Observations of \ion{H}{I} in the tail will provide an important check to these results.
This work highlights the importance of comparing simulations to detailed, multi-wavelength observations of individual systems. We are able to make predictions about this particular galaxy, such as the existence of \ion{H}{I} gas in the tail and that its three-dimensional velocity relative to the ICM is probably larger than 1413 km s$^{-1}$. However, our simulation is not able to make any prediction about why there is a separated tail in ESO 137-001. We also do not reproduce the exact surface brightness distribution along the X-ray tail. As we have discussed, our goal was not to reproduce this specific tail. In order to do this, we would recommend modeling a smaller galaxy and matching the inclination angle between the galaxy and the ICM wind.
Using our comparison with the observations of ESO 137-001, we also draw more general conclusions about the importance of turbulence and the efficiency of heat conduction in the ICM. We conclude that the mixing rate of the hot stripped ISM (T $>$ 10$^5$ K) is well-modeled in our simulations using only adiabatic compression and turbulent mixing down to the resolution of our simulations (38 pc). We call upon observers to test our predictions of where in clusters tails will be X-ray bright, and to use deep observations to verify the connection between H$\alpha$ and \ion{H}{I} gas. Observations of the \ion{H}{I} tail of ESO 137-001 can be used to test mixing of cold clouds into the ICM in our simulations.
\acknowledgements
We acknowledge support from NSF grants AST-0547823, AST-0908390, and AST-1008134, as well as computational resources from the National Center for Supercomputing Applications, NSF Teragrid, and Columbia University's Hotfoot cluster. We thank Jacqueline van Gorkom, Jeffrey Kenney and the referee Bernd Vollmer for useful discussions, as well as Elizabeth Tasker for invaluable help setting up the initial conditions.
|
\section{Introduction}
Despite our poor knowledge of the sulfur chemistry in the interstellar medium, the abundance of S-bearing
molecules observed in protostars has been used to put some constraints on the age of the sources being studied
\citep{1997ApJ...481..396C,1998A&A...338..713H,2003A&A...399..567B,2003A&A...412..133V,2004A&A...422..159W,2009A&A...504..853H}.
Most of the rate coefficients of the reactions involving sulfur-bearing molecules have not been studied in
the laboratory because of the difficulty of conducting experiments with them. Rate coefficients are then uncertain and
many must be taken with caution. Another problem related to sulfur chemistry is the depletion of the element.
Observations of the atoms in the diffuse medium have shown a relationship between the amount of depletion of the
elements (in a more or less refractory form) with the density of the cloud \citep[see][and references
therein]{2009ApJ...700.1299J}. Sulfur is one of the exceptions that do not show this depletion.
As far as
observations can go, the atomic abundance seems to be the same as the cosmic sulfur abundance of about
$1.5\times 10^{-5}$ compared to the proton density \citep{2001ApJ...554L.221S}. If using such an elemental abundance
in chemical models for cold clouds, we would overpredict gas phase abundances of SO, SO$_2$, and
CS by several orders of magnitude \citep{1999MNRAS.306..691R}. For this reason, a depletion of
this element is usually assumed to proceed in dense clouds where the amount of S available for the chemistry
is taken to be two orders of magnitude less than the cosmic abundance. Recently, by reanalyzing hundreds of
observations of atomic sulfur line, \citet{2009ApJ...700.1299J} has argued that not observing the sulfur depletion in
diffuse medium is an observational bias. If confirmed, this result would solve one of the interstellar sulfur
mysteries.
The form of sulfur on interstellar grains is not known. OCS is the only molecule actually observed on
interstellar ices \citep{1997ApJ...479..839P}. Since hydrogenation is expected to be the most efficient
process on grains, H$_2$S is also probably formed, although with a smaller abundance than the observational
limit of $10^{-7}$ \citep{1998ARA&A..36..317V}. H$_2$SO$_4$ has been proposed and looked for without any
success \citep{2003MNRAS.341..657S}. Finally, \citet{2004A&A...422..159W} propose that sulfur is in the form
of polymers or aggregates of sulfur that would be quickly converted into atomic sulfur once evaporated in
the gas phase. In protostellar envelopes, when the temperature increases, the sulfured molecules evaporated from the
grains are transformed into SO first and then into SO$_2$
\citep{1997ApJ...481..396C,2004A&A...422..159W}. Thus abundances of these species have been used on several
occasions to trace the chemical evolution of such objects.
\citet{2003A&A...412..133V} have studied the sulfur chemistry in nine high-mass protostars and find that high-energy transitions probing the inner parts of the protostars where the temperature exceeds 100~K were
mandatory for using S-bearing molecules as chemical clocks. They were only able to obtain the abundance of SO$_2$
in these regions. In addition, they propose that OCS is the main carrier of sulfur on grains, based on the
high excitation temperature of the molecule and its high abundance in the protostars.
In the present work, we made a
1D modeling of the sulfur chemistry in four high-mass compact objects, IRAS 18264, IRAS 18162, IRAS 05358, and W43-MM1, presumably representative of the youngest stages of massive star formation. We used the density and temperature profiles determined from a detailed modeling of the SED
observed in these sources by \citet{2009A&A...504..853H}.
\citet{2009A&A...504..853H} also observed the main S-bearing molecules in these sources in order to get an
estimation of their abundance and possible evolutionary stage. The observed line profiles have shown a non-thermal line broadening that they attribute to supersonic diffusion. We then included a treatment of the diffusion in
our chemical modeling in order to study its impact on the abundance profiles computed by the model.
This paper is organized as follows. Our 1D chemical model and the model parameters are described in
section~\ref{model}. Section~\ref{results} shows the abundance (compared to H) and abundance ratio profiles
computed by the model in the studied sources, as well as the effect of turbulent mixing. We discuss the
particular problem of CS in section~\ref{CS}. We compare our model predictions with the observed abundances in
section~\ref{comp_obs} and conclude in the last section.
\section{Model description}\label{model}
The model used for this study is the Nahoon chemical model developed by \citet{2005A&A...444..883W,2010A&A...517A..21W}. This code originally computes the chemical evolution of a list of species as a function of time for a fixed temperature and density for a single point. In this new version of Nahoon, the chemical evolution is computed in a 1D structure and accounts for diffusive mixing induced by turbulent transport.
Starting from the chemical composition of a molecular cloud (see \S \ref{init_cond}), we then compute the chemical abundances as a function of the radius to the center of the protostars and as a function of time using the physical structures (density and temperature profiles) observed in the four high-mass protostars described in section~\ref{phys_sec}. We compute the evolution of chemical abundances $x_i$ following
\begin{equation}
\frac{\partial x_i}{\partial t} = P_i - L_i +\frac{1}{\rho r^2}
\frac{\partial}{\partial r} \kappa
\rho r^2 \frac{\partial}{\partial r} x_i
\end{equation}
where $P_i$ and $L_i$ are the chemical production
and loss terms, $\rho$ is the density, $\kappa$ the turbulent
diffusivity, and $r$ the spherical radius. We compute turbulent
diffusion in spherical coordinates. \\
\subsection{The chemistry}
Gas-phase chemistry is computed and species are allowed to deplete on grains and evaporate by direct temperature effect and indirect heating by collisions between grains and cosmic ray particles. No reactions on grains are taken into account in this work. Only an approximation for the formation of H$_2$ on grains is considered following \citet{1984ApJS...56..231L}. The freezeout of gas-phase species onto grains is computed using equation 1 of \citet{1992ApJS...82..167H}. The evaporation of species from the grain mantles is also computed according to \citet{1992ApJS...82..167H} for direct thermal evaporation and from \citet{1993MNRAS.261...83H} for evaporation induced by collisions between grains and cosmic-ray particles.
The chemistry is described by the Srates network from \citet{2004A&A...422..159W}. Srates is a reduced network dedicated to sulfur chemistry in warm sources. It contains 929 chemical reactions involving 76 gas-phase and 32 surface species for the elements H, He, C, O and S. This network has been assembled from different sources (OSU, UDFA and the NIST databases) and updated using the osu latest network (osu-09-2008, http://www.physics.ohio-state.edu/$\sim$eric/research.html). To select the reduced network, several previous works were followed \citep{2002A&A...381L..13R,1979ApJS...41..555H,1980ApJ...236..182H,1993MNRAS.262..915P,1997ApJ...481..396C} and the results of simulations obtained with this reduced network were tested against the ones of larger networks valid at low temperature. Srates can be downloaded at the following address: http://kida.obs.u-bordeaux1.fr/models. The binding energies for OCS and H$_2$S are taken to be 2888 and 2743~K, respectively. These binding energies were computed using equation 1 described in \citet{2004MNRAS.354.1133C} and the evaporation temperatures of these species measured by TPD experiments from the same paper (Herma Cuppen, private communication). H$_2$ cosmic-ray ionization rate is $1.3\times 10^{-17}$~s$^{-1}$.
\subsection{Diffusion}\label{turb_model}
Nonthermal broadening of lines is observed in various prestellar cores. This behavior is usually assumed to
be the consequence of some sort of unresolved gas motions, whose exact origin and properties are unknown.
However, we assume that these motions produce some kind of mixing inside the source and are often referred to as turbulence.
Our purpose is to investigate the exact influence of a turbulent-like mixing on the distribution of S-bearing
species, without presuming the exact nature of this "turbulence". The simplest way to do so is to use a mixing
length approach \citep{Prandtl25}. We added mixing terms to all the differential equations in order to couple each spatial point in the protostar envelopes with its adjacent points. Turbulence induces an enhanced diffusion in the source, whose effective
diffusivity can be written as
\begin{equation}
\kappa = {\rm v}_T\ l_T
\end{equation}
where v$_T$ and $l_T$ are a characteristic turbulent velocity and length scale (the mixing length),
respectively. Expressing v$_T$ and $l_T$ would require a detailed model for
turbulence in protostellar cores. This is not, however, within the scope of the paper. Instead, we aim to use observed
quantities as upper limits for both parameters.
To estimate the maximum efficiency of the turbulent diffusion, we assume that the velocity v$_T$ cannot be
higher than the sound speed ($\sqrt{\frac{RT}{\mu}}$, with $R$ the gas constant, $T$ the gas temperature, and
$\mu$ the mean molecular weight), and the mixing length $l_T$ is expected to be smaller than the radius of the
envelope ($\sim 3\times 10^{17}$~cm). We obtain a maximum turbulent diffusivity $\kappa_{max}$ of about
$2\times 10^{22}$~cm$^2$~s$^{-1}$.
\subsection{Initial conditions for the protostellar envelope}\label{init_cond}
\begin{table}
\caption{Grain mantle composition. }
\begin{center}
\begin{tabular}{l|cccccc}
\hline
\hline
Species & Abundance (/H) & Ref. \\
\hline
CO & $4\times 10^{-6}$ & (1) \\
H$_2$O & $5\times 10^{-5}$ & (2) \\
H$_2$CO & $2\times 10^{-6}$ & (3) \\
CH$_4$ & $5\times 10^{-7}$ & (4) \\
CH$_3$OH & $2\times 10^{-6}$ & (5) \\
OCS & $5\times 10^{-8}$ & (6) \\
H$_2$S & $5\times 10^{-8}$ & \\
S (mod 1) & $1.45\times 10^{-5}$ & \\
S (mod 2) & 0 & \\
\hline
\end{tabular}
\end{center}
References: (1) \citet{2000ApJ...536..347G}, (2) \citet{1996A&A...315L.333S}, (3) \citet{2001A&A...376..254K}, (4) \citet{1996ApJ...472..665C}, (5) \citet{1998A&A...336..352B} and (6) \citet{1997ApJ...479..839P}.
\label{mantles}
\end{table}%
Protostars form from dense clouds. As a consequence, for the initial conditions of the protostars, we computed the chemical composition of a dense molecular
cloud with a temperature of 10~K, a total H density of $2\times 10^4$~cm$^{-3}$, a visual extinction of 30, and
a cosmic-ray ionization rate of $1.3\times 10^{-17}$~s$^{-1}$. For the dense cloud, species are assumed to be
initially in the atomic form (coming from the diffuse medium) except for H$_2$. Elemental abundances (compared
to total H) used for this calculation are 0.09 for He, $2.56\times 10^{-4}$ for O and $1.2\times 10^{-4}$ for
C$^+$ \citep{2008ApJ...680..371W}. For S$^+$, we chose an elemental abundance of $5\times 10^{-7}$ for
the dense cloud in order to approximately reproduce the observed abundances of gas-phase S-bearing molecules
\citep[see][]{2004A&A...422..159W}. Using our reduced network, we have to integrate over a long time
($10^7$~yr) to obtain abundances that are similar to the ones observed in dense clouds. We in fact obtain abundances similar to our previous study
\citep[][Composition A of Table 1]{2004A&A...422..159W}. This time is probably too long compared to the age of dense molecular clouds. Using larger networks of more than four thousand gas-phase reactions, observations in dense clouds are usually reproduced at a few $10^5$~yr \citep[see for instance][]{2006A&A...451..551W}. The reason we need more time is most likely that we have fewer chemical species and that it takes more time for the ionization fraction of the gas to decrease. At high temperatures however, we do not have this problem.
This calculation gives us our gas-phase
composition prior to the formation of the protostar. The CO abundance computed by the model on the grain surfaces (result of the depletion of gas-phase CO) is about ten times larger than what is observed. We decreased this value to $4\times 10^{-6}$ \citep[CO abundance observed in interstellar ices towards the massive protostar line of sight W33 by][]{2000ApJ...536..347G} and put the rest of the CO back in the gas phase. The total CO gas-phase abundance is then $6.08\times 10^{-5}$. This change will not affect our chemistry since most of the protostar envelopes are above the evaporation temperature of CO ($\sim$18~K). Like most dense cloud chemical models, our simulations fail to
reproduce the observed limits on the O$_2$ abundance in dense clouds, which is $5\times 10^{-8}$ (compared to H) in L134N
\citep{2003A&A...402L..77P}, 160 times smaller than the abundance computed by our model. We then
decreased the gas-phase abundance to its observed limit, and the rest of the oxygen we put in the abundance of atomic oxygen. The computed O$_2$ abundance on grains is $5\times 10^{-5}$. For H$_2$O, our model predicts a gas-phase abundance
of $4\times 10^{-8}$, below the observational limit in L134N \citep[$1.5\times
10^{-7}$][]{2000ApJ...539L.101S} so we did not change it.
Since we did not compute the grain surface chemistry occurring during the dense cloud phase, we modified the
grain surface composition obtained from our modeling. For the species
observed on interstellar grains, we adopted the abundances listed in Table~\ref{mantles}
\citep[see][]{2004A&A...422..159W}. Note that CO$_2$, one of the main constituent of grain mantles, is not present in our network. OCS is the only S-bearing species observed in the solid state in the
interstellar medium. H$_2$S is highly suspected of being present, although at a nondetectable level. We set
its abundance on grain surfaces to the upper limit calculated by \citet{1998ARA&A..36..317V}. All the rest of
the sulfur was assumed to be either in the atomic form on grains, our model 1, or in a refractory form
that cannot evaporate at the temperatures of hot cores, our model 2.
This gas-phase and grain mantle composition is used as initial conditions for the whole envelope of the four protostars we studied. Since we do not include any chemical reaction on the grain surfaces, the gradual depletion of the species during the evolution of the chemistry only affects the fraction of species in the gas and on the surface and does not change the chemistry itself.
\subsection{Selected sources: physical structure and observed abundances}\label{phys_sec}
\begin{figure}
\includegraphics[width=1\linewidth]{16164fg1.eps}
\caption{Temperature and density profiles in the four sources from \citet{2009A&A...504..853H}. \label{phys}}
\end{figure}
\begin{table}
\caption{S-bearing molecular abundances ($\times 10^{-10}$) derived in the four sources by \citet{2009A&A...504..853H}. }
\begin{center}
\begin{tabular}{l|l|llll}
\hline
\hline
Molecule & T (K) & \multicolumn{4}{c}{Sources} \\
& & w43-mm1 & IRAS 18264 & IRAS 05358 & IRAS 18162 \\
\hline
CS & 60 & 36 & 4.0 & 2.3 & - \\
& 100$^a$ & 160 & 1.1 & - & 4.6 \\
OCS & 60 & 92 & 2.5 & 2.4 & 1.4 \\
& 100 & 130 & 1.1 & 3.7 & 3.3 \\
H$_2$S & 60 & 3.0 & 1.3 & 1.2 & 3.1 \\
& 100 & 4.0 & 0.15 & 0.51 & 0.30 \\
SO & 60 & 0.90 & 0.48 & 1.1 & 2.4 \\
& 100 & 24 & 0.64 & - & 3.6 \\
SO$_2$ & 60 & 2.0 & 0.42 & 0.62 & 1.7 \\
& 100 & 160 & 2.0 & 6.4 & 7.6 \\
\hline
\end{tabular}
\end{center}
$^a$ The CS abundance at this excitation temperature was derived using the CS (7-6) intensity line blended with an H$_2$CO line \citep[see][]{2009A&A...504..853H}.\\
The symbol - means that the no abundance was derived because the molecular line has not been observed.
\label{obs}
\end{table}%
The source sample from Herpin et al. (2009) is made of two IR-quiet dense cores \citep[MSX Flux$_{21\mu m}<10$ Jy at 1.7 kpc, following the definition of][]{2007A&A...476.1243M} and two slightly brighter ones (IRAS05358+3543 and IRAS18162$-$2048), with bolometric luminosities (0.6--2.3)$\times 10^4$~L$_{\odot}$ at distances 1.8--5.5~kpc and sizes of $\sim$0.11--0.13~pc. These massive dense cores are meant to represent the earliest phases of the high-mass star formation.
From the fitted SEDs and the literature, Herpin et al (2009) propose a rough evolutionary classification of the four objects (W43-MM1 being the youngest object): W43-MM1~$\rightarrow$ IRAS18264$-$1152~$\rightarrow$ IRAS05358+3543~$\rightarrow$ IRAS18162$-$2048. Nevertheless and despite this youth, W43-MM1 appears to lie apart from the other massive dense cores as it has very likely already developed a hot core.
The temperature and density profiles for the four sources have been derived by \citet{2009A&A...504..853H} from the analysis of the spectral energy distribution (SED) with the radiative transfer code MC3D \citep{1999A&A...349..839W} \citep[see also][]{2008A&A...488..579M}. The profiles are shown in Fig.~\ref{phys}. Temperatures and densities increase towards the center of the sources. The three IRAS sources show very similar profiles, whereas W43-MM1 is denser and colder. For the chemical modeling, we assumed that the gas temperature and density are abruptly raised from the cold molecular phase to the ones currently observed following previous works \citep{1997ApJ...489..122D,2000A&A...358L..79V,2002A&A...389..446D}. This is of course a strong assumption but we prefer not to add another uncertain parameter. Considering the fast evolution of sulfur chemistry at high temperature (see section 3.1), we do not expect this assumption to affect our main conclusions. Only a careful coupling of the chemistry with a dynamical infalling envelope can, however, answer that question.
\citet{2009A&A...504..853H} observed several transitions of OCS, H$_2$S, SO, SO$_2$, and CS in the four sources,
as well as some of their $^{34}$S minor isotopologues. They constrained the species abundances at different radii depending on their excitation temperatures (60, 75, and 100~K) (see Table 5 of their paper). We report in Table~\ref{obs} their observed abundances for excitation temperatures of 60 and 100~K. We assume that we are at LTE so the excitation temperature of the gas reflects the gas and dust temperatures. The presence of a hot core in W43-MM1 might explain the observed higher abundance of H$_2$S and OCS in that source relative to the other ones. Moreover, strong shocks within this source probably influence the chemistry. One surprising result is the small abundance of the S-bearing molecules found in all four sources compared to the cosmic abundance (three to four orders of magnitude smaller). We compare our modeling with their observational results below.
\section{Modeling results}\label{results}
\subsection{Abundances through the protostellar envelopes}
\begin{figure}
\includegraphics[width=1\linewidth]{16164fg2.eps}
\caption{OCS, H$_2$S, CS, SO and SO$_2$ abundance (/H) profiles as a function of radius from the center in the envelope of four high mass protostars (IRAS 18264, IRAS 18162, IRAS 05358 and W43-MM1). Initial composition of Model 1 is used. Integrated time is $10^4$~yr. \label{ab_mod1}}
\end{figure}
\begin{figure}
\includegraphics[width=1\linewidth]{16164fg3.eps}
\caption{OCS, H$_2$S, CS, SO and SO$_2$ abundance (/H) profiles as a function of radius from the center in the envelope of four high mass protostars (IRAS 18264, IRAS 18162, IRAS 05358 and W43-MM1). Initial composition of Model 2 is used. Integrated time is $10^4$~yr. \label{ab_mod2}}
\end{figure}
\begin{figure}
\includegraphics[width=1\linewidth]{16164fg4.eps}
\caption{OCS, H$_2$S, CS, SO and SO$_2$ abundance (/H) profiles as a function of radius from the center in the envelope of IRAS 18264. Initial composition of Model 2 is used. Integrated times are $10^3$~yr for the figure on the top and $5\times 10^4$~yr for the figure on the bottom. \label{ab_time}}
\end{figure}
\begin{figure}
\includegraphics[width=1\linewidth]{16164fg5.eps}
\caption{Abundance (/H) profiles of a selection of species as a function of radius from the center in the envelope of W43-MM1. Initial composition of Model 2 is used. Integrated time is $10^4$~yr. \label{ab_autre_w43}}
\end{figure}
\begin{figure}
\includegraphics[width=1\linewidth]{16164fg6.eps}
\caption{Abundance (/H) profiles of a selection of species as a function of radius from the center in the envelope of IRAS 18264. Initial composition of Model 2 is used. Integrated time is $10^4$~yr. \label{ab_autre_i18264}}
\end{figure}
Figures~\ref{ab_mod1} and \ref{ab_mod2} show the abundances of the main S-bearing molecules computed by the chemical model as a function of the radius to the central object in the envelope of the four protostars at $10^4$~yr, which was approximately the age derived by \citet{2002A&A...389..446D} for the massive protostar AFGL 2591. This age was derived by comparing the predictions of chemical models with a large number of molecules observed in this source. Although AFGL2591 may be a bit more evolved than the four sources studied here (AFGL2591 is a mid-IR bright HMPO, while the 4 others are mid-IR quiet HMPOs), all these are still HMPOs, i.e. not hot molecular cores or ultra-compact HII regions. As far as we know, it is the only large chemical study of such sources so we will use it as a reference. The two figures have been obtained for a different amount of the atomic sulfur initially available on the grain surfaces (models 1 and 2, see \ref{init_cond}). IRAS 18264, IRAS 18162 and IRAS 05358 show very similar molecular profiles since their temperature and density
gradients are similar (Fig.~\ref{phys}), except for radii larger than $7\times 10^3$~AU where the depletion of the species is more pronounced in IRAS 18162. W43-mm1 shows different abundance profiles,
especially the outer envelope where the molecules are more depleted because of higher densities. In this
source, both models 1 and 2 lead to smaller abundances of SO, H$_2$S, and CS molecules compared to the other sources. All species show strong variations in their abundances with radius.
The radius at which the gas and dust temperatures are around 40-50~K represents a transition from an "outer envelope", where the S-bearing molecules are depleted on the grain mantles, and an "inner envelope" where the abundances of the S-bearing species are much larger. In the inner envelope, towards the center of
the protostar, all molecules except SO$_2$ again show a decrease in their abundances at radii depending on the source, the molecule and the model. In IRAS 18264 and model 1 for instance, OCS drops at 200~AU and SO abundance at 400~AU. The H$_2$S abundance decreases by more than one order of magnitude between 1000 and 400~AU before increasing again. In all sources, SO$_2$ is the only species whose abundance is rather constant in the inner envelope, whereas the CS abundance is quite low and shows a strong dip where the temperature is about 200~K (see discussion in section~\ref{CS}). In the very outer part of the envelopes (at radii larger than 10 000~AU), the species abundances increase again because the density of the gas is lower and species have not had time yet to deplete on the grains.
In model 1, a large quantity of atomic sulfur is evaporated in the
gas phase when the temperature is above $\sim$20~K. Atomic sulfur is then quickly converted into SO by reacting with
O$_2$. SO in turn produces SO$_2$ by reactions with OH and O. For temperatures higher than $\sim 40-50$~K, OCS and H$_2$S are evaporated in the gas phase, and their abundances are determined by their initial solid abundances. In model 2, it is the destruction of H$_2$S through
the following paths that releases the necessary sulfur to produce SO and SO$_2$ at high temperature: H$_2$S + H$_3$O$^+$
$\rightarrow$ H$_3$S$^+$ + H$_2$O, H$_3$S$^+$ + e$^-$ $\rightarrow$ HS + H$_2$ and HS + O $\rightarrow$ SO + H.
SO$_2$ is the dominant S-bearing molecule after a few $10^4$~yr in all models at radii smaller than $7\times 10^3$~AU. We anticipate the comparison with the observation given in section \ref{comp_obs} by saying that model 1 is far from reproducing our observations since very low abundances of S-bearing molecules have been observed. For this reason, we only discuss model 2 in the rest of the paper.
Some other species abundances (for model 2 and $10^4$~yr) are shown in Fig.~\ref{ab_autre_w43} for the source w43-mm1 and Fig.~\ref{ab_autre_i18264} for the source IRAS 18264 (representative of the IRAS sources of our sample). In IRAS 18264, the CO abundance is constant over the envelope ($\sim 8\times 10^{-5}$) since the temperature is above 20~K, CO depletion occurs whereas at radii larger than 10000 AU in w43-mm1. Inside the protostellar envelopes (radii smaller than 200~AU for w43-mm1 and 600~AU for IRAS 18264), the main reservoir of oxygen is water, except for CO. At larger radii, the oxygen is mainly in the atomic form. This transition between H$_2$O and O corresponds to the dip in the CS abundance (see also section 4). Similarly, the peak in HOCS$^+$, precursor of CS, around 2000~AU in w43-mm1 and 5000~AU in IRAS18264 is responsible for the peak in the CS abundance.
The abundances in Figs.~\ref{ab_mod1} and \ref{ab_mod2} are displayed for an age of
$10^4$~yr. By age, we mean the time during which the chemistry evolves after the temperature and
density of the envelope have reached their present state. As an example, the chemical profiles in the source IRAS 18264, for the elemental abundances
of model 2, are shown in
Fig.~\ref{ab_time} for two different chemical ages: $10^3$ and $5\times 10^4$~yr. The chemistry
evolves quickly inside the protostellar envelopes and the abundances are similar at $10^3$ and $10^4$~yr except for the
SO$_2$ abundance, which increases significantly with time. As time evolves after $10^4$~yr, OCS and CS at
radii 2000-6000~AU are destroyed by HCO$^+$. Abundances in the outer parts of the envelope also evolve toward
more depletion.
It is interesting to notice that the abundance of OCS does not change much at radii smaller than 2000 AU with
time. In fact, it takes more than $10^5$~yr to significantly destroy this molecule in the conditions of the
studied massive protostars. If OCS were the major carrier of S-bearing molecules on grains as suggested by
\citet{2003A&A...412..133V}, time scales for the sulfur chemistry would be longer than what we considered
here.
\subsection{Abundances ratios}
\begin{figure}
\includegraphics[width=1\linewidth]{16164fg7.eps}
\caption{SO/SO$_2$, OCS/SO$_2$, H$_2$S/SO$_2$ and CS/SO$_2$ abundance ratios as a function of time at radius where the temperature is 100~K in IRAS 18264 and W43-MM1. Initial composition of Model 2 is used. \label{abrap_mod2}}
\end{figure}
It is common when comparing with observations to use abundance ratios between two observed species rather than
abundances compared to H$_2$. We have observational constraints on the abundances at two different radii
corresponding to gas temperatures of 60~K and 100~K (see section \ref{phys_sec}). The abundance (and abundance
ratios) observed at 60~K are probably not good tracers for the evolution of the protostars since we probe the
regions of evaporation versus depletion processes of the sulfur-bearing species. To simplify the problem
since we have many parameters, we display in Fig.~ \ref{abrap_mod2} the abundance ratios of SO, OCS, H$_2$S,
and CS compared to SO$_2$ in IRAS 18264, and W43-MM1 as a function of time for the radii corresponding to the
temperature 100~K (2680~AU in IRAS 18264 and 640~AU in W43-MM1).
We did not show the abundance ratios for IRAS
18162 and IRAS 05358, because they are very similar to those in IRAS 18264. All four abundance ratios decrease with time. SO/SO$_2$ shows a monotone variation with time contrary to the three other ratios, which drastically decrease at times between $10^4$ and $10^5$~yr. Since we do not reproduce the observed abundance of CS, OCS/SO$_2$ and H$_2$S/SO$_2$
seem to be the best candidates to constrain some evolutionary time for these sources.
\subsection{Influence of diffusion}
\begin{center}
\begin{figure}
\includegraphics[width=1\linewidth]{16164fg8.eps}
\caption{OCS, H$_2$S, CS, SO and SO$_2$ abundance (/H) profiles as a function of radius from the center in the envelope of IRAS 18264. Initial composition of Model 2 is used. Integrated times is $10^4$~yr. Three different turbulent diffusivity $\kappa$ are used for each panel: $2\times 10^{14}$~cm$^{-2}$~s$^{-1}$ for the upper panel, $2\times 10^{16}$~cm$^{-2}$~s$^{-1}$ for the panel in the middle and $2\times 10^{18}$~cm$^{-2}$~s$^{-1}$ for the lower panel. \label{nu}}
\end{figure}
\end{center}
The general effect of diffusion is to smooth out the abundance profiles across the envelope and decrease the
depletion of the species abundances in the outer parts. Abundance profiles for the source IRAS 18264 and model 2 are shown
in Fig.~\ref{nu} for three different values of the turbulent diffusivity $\kappa$: $2\times
10^{14}$~cm$^{-2}$~s$^{-1}$, $2\times 10^{16}$~cm$^{-2}$~s$^{-1}$, and $2\times 10^{18}$~cm$^{-2}$~s$^{-1}$. These figures can be compared to Fig.~\ref{ab_mod2} where $\kappa$ is zero.
Our largest $\kappa$ is four orders of magnitude smaller than the maximum value derived in
section~\ref{turb_model}, but the effect on the chemistry is already strong. In that case, the abundances of OCS, H$_2$S, SO, and
SO$_2$ are then constant and equal up to radius $5\times 10^3$~AU. The CS abundance is also rather smooth, but
four orders of magnitude less than the other species. For smaller $\kappa$, the smoothing is less efficient
and becomes negligible for $\kappa$ lower than $2\times 10^{14}$~cm$^{-2}$~s$^{-1}$.
\section{The CS abundance}\label{CS}
As already noticed, we predict that CS is much less abundant than the other S-bearing species contrary to the observations, which show abundances of the same order for all species \citep[see also][]{2004A&A...422..159W}. In our model, CS is
produced at high temperature by the dissociative recombination of HCS$^+$ in the central part where the
temperature is above 300~K and the dissociative recombination of HOCS$^+$ at lower temperature. HCS$^+$ is
itself produced by C$^+$ + H$_2$S and HOCS$^+$ by HCO$^+$ + OCS and H$_3^+$ + OCS. For the destruction, we
also have two regimes: CS + H$^+$ at temperatures higher than 300~K and CS + O for lower temperatures.
\citet{1997ApJ...481..396C} and \citet{1998A&A...338..713H} have produced much more CS than in this work because
they did not include any atomic oxygen in their initial conditions. Except for helium and hydrogen, they
considered that all species were already in a molecular form. Oxygen was then in the form of CO, H$_2$O, and
O$_2$. Contrary to what was said in \citet{2004A&A...422..159W}, changing the rate coefficient of the
dissociation of CO by secondary photons produced by cosmic-rays does not change the abundance of CS predicted
at high temperature. Since this problem is not seen in cold dense sources \citep[see for instance][]{2006A&A...451..551W}, we may be missing some formation paths for CS at high temperature. Some rate coefficients may also not be accurate for the temperature range where we are using them. As an example, some new recommendations for the rate coefficient of the neutral-neutral reaction O + CS for temperatures between 150 and 300~K have been posted on the KIDA database (http://kida.obs.u-bordeaux1.fr) following this work. However, the small proposed changes in this rate coefficient does not change the gas-phase CS abundance in our simulations.
That we do not reproduce the CS abundance at 100~K may also come from our other assumptions in our modeling such as our initial conditions, the temperature and density profiles, the brutal rise in the temperature and density from the molecular cloud stage. This, however, should not influence the abundance of the other S-bearing molecules by more than a factor of a few since they all have similar observed abundances. The CS problem may also be related to the form of sulfur on grains. Recent laboratory experiments show that H$_2$S on ISM analog surfaces are easily destroyed by energetic particles, and surface S then recombines to form OCS, SO$_2$, C$_2$S, and a majority of sulfur-rich residuum \citep{2010A&A...509A..67G}. The surface C$_2$S may then be evaporated in the gas phase and contribute to the CS chemistry.
Last but not least, we stress that the discrepancy might also originate in the observed abundances. The observed abundance of CS representative of the 100~K layers in the protostellar envelopes was derived using the CS (7-6) line emission, which has an upper level energy of 65.8 K. First of all, these abundances are derived independently for each observed transition of each species, assuming that the emission for a given line mostly comes from one single place at a given temperature and density, whereas lines are excited over several layers. Considering the upper level energy of CS (7-6), it is likely that this transition is also excited in the layers of the protostellar envelopes at temperatures below 100~K where our chemical model predicts larger abundances of CS. The CS abundance is predicted by our chemical models to vary much more with the radius than for the other S-bearing abundances at these temperatures. As a result, an uncertainty about the location of the emission may lead to a larger uncertainty concerning the abundance. Furthermore, this line was blended with an H$_2$CO line, and the procedure to separate lines did not work well in IRAS 18264 and w43-mm1 \citep[see][]{2009A&A...504..853H}. The CS abundances at 100~K derived by \citep{2009A&A...504..853H} may then be overestimated, and one would need to observe higher energy transitions of CS (unfortunately not accessible from the ground) to confirm its abundance at high temperature.
\section{Comparison with observations}\label{comp_obs}
\begin{center}
\begin{figure}
\includegraphics[width=1\linewidth]{16164fg9.eps}
\caption{OCS/SO$_2$ (upper panel), H$_2$S/SO$_2$ (middle panel) and SO/SO$_2$ (lower panel) abundance ratios as a function of time at radius where the temperature is 100~K in the four studied sources. Initial composition of Model 2 is used. For each source, the observed abundance ratio is reported on the curves as horizontal thick line. Two values are drawn for IRAS 18162 as two SO$_2$ abundances have been determined using two different lines \citep[see][]{2009A&A...504..853H}. No observed value SO/SO$_2$ is reported for IRAS 05358. \label{abrap_obs}}
\end{figure}
\end{center}
In this section, we discuss the comparison between our modeling and observations of S-bearing molecules
in the four high-mass protostars from \citet{2009A&A...504..853H}. As a summary, very small abundances have
been found compared to the cosmic abundance of sulfur ($1.5\times 10^{-5}$ compared to H). The sum of all
observed S-bearing species is only $2\times 10^{-9}$ to $5\times 10^{-8}$. This result suggests that we are
closer to model 2 than model 1, i.e. a large part of the sulfur in this object is in a stable, not
observed form and does not participate in the gas phase chemistry. We discuss this point in the next section. We are far from reproducing all the observed abundances correctly. We overproduce them by more than one order
of magnitude even using model 2, except for CS which is strongly underestimated.
The OCS/SO$_2$, H$_2$S/SO$_2$, and SO/SO$_2$ abundance ratios
could in theory be used to trace the age of these sources. We report in Fig.~ \ref{abrap_obs} the abundance
ratios predicted by model 2 as a function of time for the radius corresponding to 100~K. As stated in section 3.2, the abundances derived at radii where the temperature is about 60~K trace regions of the envelopes where the depletion processes compete with the evaporation from the grains. For this reason we focus on the abundances obtained for shells of gas at higher temperature probably more representative of the evolution of the protostar itself.
In the same
figures we plot the ratios observed in each source.
The OCS/SO$_2$ and H$_2$S/SO$_2$ abundance ratios are above the observed ones until a few $10^4$~yr and then
decrease sharply to the observational values and below. As a consequence, in the observed range for these
ratios, the constraint on the age is very strong.
This time depends on the physical properties of the source. The three IRAS sources
being very similar, the time of decrease is almost the same. W43-MM1 is denser and the time at which the
abundance ratios drop is less. The constraint on the relative age of the sources provided by the observed
ratios OCS/SO$_2$ and H$_2$S/SO$_2$ is in good agreement with the evolution sequence W43-MM1 $\rightarrow$
IRAS 18264 $\rightarrow$ IRAS 05358 $\rightarrow$ IRAS 18162 proposed using other methods
\citep[see][]{2009A&A...504..853H}. If one considers the uncertainties in both the observed
\citep{2009A&A...504..853H} and modeled \citep{2005A&A...444..883W} abundance ratios, no definitive conclusion
can be drawn from these observations. The constraints on W43-MM1 are maybe more robust and indicate that
the source is younger than the three other ones. One important conclusion here is that more than the age of
the source, it is the physical conditions in the envelope that directly determine the variation in the OCS,
H$_2$S, and SO$_2$ abundance. There is of course a relation between the age and the physical conditions. OCS/SO$_2$ and H$_2$S/SO$_2$ can, however, not be used to compare protostars that are close in age. The SO/SO$_2$
abundance ratios are more sensitive to time. We do need a very young chemical model for W43-MM1 to reproduce
the observed abundance ratio in this source, which does not agree with the other constraints. The
SO/SO$_2$ abundance ratios observed in IRAS 18162 and IRAS 18264 are also too close to conclude anything about their age.
Observations also seem to indicate significant variations in the species abundances with the radius in a source.
H$_2$S, for instance, was observed to vary by one order of magnitude in IRAS 18264 for radii between $2.6\times
10^3$ and $5.4\times 10^3$~AU, which is approximately reproduced by our model. Including diffusion
smooths the predicted abundance profiles. If chemical variations are so strong across these objects, then we
would expect the turbulent diffusivity to be smaller than $2\times 10^{14}$~cm$^{2}$~s$^{-1}$. Supersonic
velocities due to diffusion have been observed in these sources. Assuming a turbulent velocity of about 1~km
s$^{-1}$ as observed, the maximum mixing length would be a few $10^{-4}$~AU, much smaller than what
we would expect. The very small mixing
efficiency associated with the observed nonthermal velocity raises questions about the nature of this
"turbulence". Either turbulence has a very peculiar nature, or more likely, the observed structure is
unresolved and composed of smaller structures of varying turbulence intensity. Of course, nonthermal motions
can be produced by a variety of processes (rotation, infall, outflow, etc), which will not always induce an
efficient mixing, but can still be mistaken for turbulence, at least partly.
\section{Conclusions}
We performed a 1D chemical modeling of the sulfur-bearing species in four young high-mass protostars (IRAS 18264, IRAS 18162, IRAS 05358, and W43-MM1) using temperature and density profiles determined from their SED by \citet{2009A&A...504..853H}. In addition to the time-dependent chemistry, we studied the effect of a turbulent-like mixing on the abundance profiles through the envelopes. Here are our main conclusions:
\begin{itemize}
\item [-] Sulfur chemistry depends strongly on the 1D physical conditions. Any observed set of abundances should be compared with a chemical model computed with the same temperature and density traced by the observations.
\item [-] To use sulfur chemistry as chemical clocks, observations tracing the gas at temperatures higher than 70~K should be used to avoid confusion with depletion mechanisms. At lower temperatures, S-bearing molecules stick on grains so that their abundance in the gas phase will also depend upon this process.
\item [-] In our case, no conclusion can be drawn on the relative age of IRAS 18264, IRAS 18162 and IRAS 05358 if one considers the uncertainties in the observed and modeled abundances because these sources are too close in age. W43-MM1 seems, in contrast, younger than the other sources.
\item [-] Turbulent mixing could occur in young high-mass protostars on a too small geometric scale to affect
the chemical abundance profiles through the envelope. Such a small scale suggests that either turbulence is
very unusual or the structure is unresolved. There is a possibility that at least some of the nonthermal
broadening of lines would be the consequence of a non turbulent process, inducing a very inefficient mixing
inside the protostellar envelopes.
\item [-] The CS molecule is predicted by our chemical models to be less abundant than the other S-bearing species by several orders of magnitude at temperatures above 100~K \citep[see also][]{2004A&A...422..159W}. This seems to disagree with the observations, although the observed abundances from \citet{2009A&A...504..853H} may be overestimated. If the disagreement were true, however, this suggests that some formation paths are missing in our networks at high temperature.
\item [-] Our comparisons between observations and chemical modeling seem to indicate that the majority of the sulfur is still in a refractory form on grains. It is interesting to note that \citet{2004A&A...422..159W} also need a depletion by a factor of ten of the elemental sulfur abundance to reproduce observations in Orion-KL, whereas no depletion is mandatory to reproduce the observations in the low-mass protostar IRAS 16293-2422. Based on laboratory experiments, \citet{2010A&A...509A..67G} have recently suggested that H$_2$S on interstellar grains are easily destroyed by energetic particles to form OCS, SO$_2$, and a majority of sulfur-rich residuum, which could be polymers of sulfur or amorphous aggregates of sulfur, as suggested by \citet{2004A&A...422..159W}. In the environments of high-mass star formation, interstellar ices are probably exposed to stronger particle fluxes so that a larger quantity of the atomic sulfur could be converted in refractory form. The identification of the sulfur-rich residuum found by \citet{2010A&A...509A..67G} could certainly bring new insight into the reservoir of sulfur in the interstellar medium.
\end{itemize}
\begin{acknowledgements}
We thank the referee for his careful reading of the paper and his suggestions. V.W. acknowledges the French CNRS program PCMI for partial support of this work.
\end{acknowledgements}
|
\section{Introduction}
\input{intro.tex}
\section{Basic notions}
\input{notions.tex}
\section{A dichotomy of monoids}
\input{dichotomy.tex}
\section{Capabilities of valence automata and transducers}
\input{automata.tex}
\section{Capabilities of valence grammars}
\input{grammars.tex}
\bibliographystyle{alpha}
|
\section{Introduction}
This submission is based on \cite{ABdMM}, where a complete discussion with references is presented. Due to space constraints we focus here on the last part of the Corfu talk - proceedings covering the first part of the talk (also presented as a shorter talk in ICHEP) can be found in \cite{Ivo_ICHEP}.
The small neutrino masses can be
understood within the see-saw mechanism where
the Standard Model (SM) is extended by adding new heavy states - right-handed (RH)
neutrinos in type I.
The see-saw mechanism also contains the ingredients for leptogenesis. We discuss
only ``unflavoured'' leptogenesis, as in
the framework of flavour symmetry models the
heavy singlet neutrinos typically have masses above $10^{13}$~GeV and
for $T\gtrsim 10^{12}$ GeV lepton flavours are indistinguishable.
In the standard thermal leptogenesis scenario singlet neutrinos $N_{\alpha}$ are
produced by scattering processes after inflation. Subsequent out-of-equilibrium
decays of these heavy states generate a CP-violating asymmetry given
by
\begin{equation}
\label{eq:cp-asymm}
\epsilon_{N_\alpha} = \frac{1}{4v^2 \pi (m_D^{R\,\dagger} \;m_D^R)_{\alpha\al}}
\sum_{\beta\neq \alpha} \mbox{Im}\, \left[\left((m_D^{R\,\dagger}
\;m_D^R)_{\beta\alpha}\right)^2\right]
f(z_\beta)\,,
\end{equation}
where $z_\beta=M_\beta^2/M_\alpha^2$, $f(z_\beta)$ is the loop function,
$m_D^R \equiv m_D V_R$ (i.e. in the basis where $M_R$ is diagonal).
We consider now a supersymmetric model based on $G_f=A_4\times Z_3\times Z_4$ \cite{Lin_Lepto} where relevant next to leading order (NLO) corrections appear only in the Dirac mass and can be parametrised with only 3 complex parameters. Neutrino masses arise only through type I see-saw so the results from \cite{ABdMM} apply.
$A_4$ is responsible for Tri-Bimaximal (TB) mixing, $Z_3\times Z_4$ generates the charged fermion hierarchy and avoids large mixing between the fields of the charged leptons and neutrino sectors.
The breaking sector includes scalar superfields $\varphi_T$, $\xi'$, $\varphi_S$, $\xi$ and $\zeta$. The assignments
are shown in table \ref{tab2}.
\begin{table}[h!]
\begin{center}
\begin{tabular}{|c|ccccccc|ccccc|}
\hline
&&&&&&&&&&&&\\[-3mm]
& $L$ & $e^c$ & $\mu^c$ & $\tau^c$ & $N^c$ & $H^u$ & $H^d$ & $\varphi_T$ & $\xi'$ & $\varphi_S$ & $\xi$ & $\zeta$ \\[0.3 mm]
\hline
&&&&&&&&&&&&\\[-3mm]
$A_4$ & $3$ & $1$ & $1$ & $1$ & $3$ & $1$ & $1$ & $3$ & $1'$ & $3$ & $1$ & $1$ \\[0.3 mm]
$Z_3$ & $1$ & $1$ & $1$ & $1$ & $\omega$ & $1$ & $1$ & $1$ & $1$ & $\omega$ & $\omega$ & $\omega^2$ \\[0.3 mm]
$Z_4$ & $1$ & $-i$ & $-1$ & $1$ & $1$ & $1$ & $-i$ & $i$ & $i$ & $1$ & $1$ & $1$ \\[0.3 mm]
\hline
\end{tabular}
\caption{Matter and scalar content of the model and their transformation properties under $G_f$ \cite{Lin_Lepto}.}
\label{tab2}
\end{center}
\end{table}
The LO Yukawa superpotential is expressed as an expansions in the cut-off of the theory $\Lambda$:
\begin{eqnarray}
&&\begin{array}{rcl}
\mathcal{W}_\ell&=&\dfrac{1}{\Lambda} y_\tau \left(L\varphi_T\right)\tau^cH^d+ \dfrac{1}{\Lambda^2} y_\mu^{(1)} \left(L\varphi_T\right)''\xi'\mu^cH^d+\dfrac{1}{\Lambda^2} y_\mu^{(2)} \left(L\varphi_T\varphi_T\right)\mu^cH^d+ \\ [3 mm]
&+&\dfrac{1}{\Lambda^3} y_e^{(1)} \left(L\varphi_T\right)'\left(\xi'\right)^2e^cH^d+\dfrac{1}{\Lambda^3} y_e^{(2)} \left(L\varphi_T\varphi_T\right)''\xi'e^cH^d+ \dfrac{1}{\Lambda^3} y_e^{(3)} \left(L\varphi_T\varphi_T\varphi_T\right)e^cH^d\;, \\
\end{array}\\
&&\hspace{3mm}\mathcal{W}_\nu=-\dfrac{1}{\Lambda} y \left(LN^c\right)\zeta H^u+ x_a\left(N^cN^c\right)\xi+ x_b\left(N^cN^c\varphi_S\right)\;,
\end{eqnarray}
where $(\ldots)$, $(\ldots)'$ and $(\ldots)''$ stand for the contraction in the representations $1$, $1'$ and $1''$ of $A_4$, respectively.
The flavon superfields acquire the following VEVs ($v_T$, $u'$, $v_S$, $u$ and $w$ are small parameters):
\begin{equation} \begin{array}{lllll}
\vev{\varphi_T}= \left(
\begin{array}{c}
0 \\
v_T \\
0 \\
\end{array}
\right)\;, &
\vev{\xi'}= u'\;, &
\vev{\varphi_S}=\left(
\begin{array}{c}
v_S \\
v_S \\
v_S \\
\end{array}
\right)\;,&
\vev{\xi}= u\;, &
\vev{\zeta}= w\;,
\end{array}
\end{equation}
and these VEVs can be aligned naturally from the scalar potential and lead to TB mixing \cite{Lin_Lepto}. The symmetry prevents deviations from these VEVs at NLO and allows the order of magnitude relations between parameters $v_T \sim u'$ and $v_S \sim u \sim w$, assuming at most a mild hierarchy among the two sets.
The neutrino mass matrix gets contributions from the type I see-saw. We have:
\begin{eqnarray}
\label{neutLO}
m_D=\left(
\begin{array}{ccc}
1 & 0 & 0 \\
0 & 0 & 1 \\
0 & 1 & 0
\end{array} \right)\dfrac{y\,w\,v^u}{\Lambda}\,,
&\quad&
M_R= \left(
\begin{array}{ccc}
b+2d & -d & -d \\
-d & 2d & b-d \\
-d & b-d & 2d
\end{array} \right)u\,,
\end{eqnarray}
with $v^u=\vev{H^u}$, $b\equiv2x_a$ and $d\equiv2x_bv_S/u$.
$M_R$ and $m_\nu$ are diagonalised by the TB mixing matrix $U_{TB}$, giving as eigenvalues $M_1=|b+3d|$, $M_2=|b|$, $M_3=|b-3d|$ and $m_i=(y\,w\,v^u)^2/(\Lambda^2M_i)$.
To estimate $\epsilon_{N_\alpha}$, we write the Dirac mass matrix in the basis of diagonal and real RH neutrinos:
\begin{equation}
m_D^R=m_DU_{TB}D'\,,
\end{equation}
where $D'=\text{diag}(e^{i\phi_1/2},\,e^{i\phi_2/2},\,e^{i\phi_3/2})$ and $\phi_\alpha$ are the phases of $b+3d$, $b$, $b-3d$ respectively (eigenvalues of $M_R$).
$m_D^{R\dag} m_D^R$ is diagonal and thus $\epsilon_{N_\alpha}=0$, in agreement with the model-independent proof in \cite{ABdMM}.
A non-vanishing asymmetry can be obtained at NLO. In this
model $Z_3\times Z_4$ only admits NLO corrections to the Dirac terms, and the ones that can not be reabsorbed in a redefinition of the LO parameters are:
\begin{equation}
-\mathcal{W}^{NLO}_\nu=\dfrac{1}{\Lambda} y_1 \left(LN^c\right)'\left(\varphi_S\varphi_S\right)''H^u+ \dfrac{1}{\Lambda} y_2 \left(LN^c\right)''\left(\varphi_S\varphi_S\right)'H^u+\dfrac{1}{\Lambda} y_3 \left(\left(LN^c\right)_A\varphi_S\right)\xi H^u\,,
\end{equation}
where $(\ldots)_A$ refers to the asymmetric contraction of triplets. The deviations to $m_D$ can be written as
\begin{equation}
\label{md1}
m_D^{(1)}=
\left(
\begin{array}{ccc}
0 & y_1+y_3 & y_2-y_3 \\
y_1-y_3 & y_2 & y_3 \\
y_2+y_3 & -y_3 & y_1 \\
\end{array}
\right)v^u\dfrac{v_S^2}{\Lambda^2}\,,
\end{equation}
where $y_3$ accounts for the ratio $u/v_S$.
Including \eq{md1}, the TB mixing receives small perturbations according to $U_\nu=U_{TB}\delta U$, where only the element $(\delta U)_{13}$ is relevant. Parametrising this term as:
\begin{equation}
(\delta U)_{13}=\sqrt{\dfrac{3}{2}}\sin\theta_{13}e^{i \delta }\sim\mathcal{O}\left(\dfrac{v_S}{\Lambda}\right)\,,
\end{equation}
where $\delta$ is the CP-violating Dirac phase in the standard parametrisation of the mixing matrix, we get
\begin{equation}
\label{expang}
\sin^2\theta_{23}=\dfrac{1}{2}(1+\sqrt2\cos \delta \sin\theta_{13}) \qquad\qquad\sin^2\theta_{12}=\dfrac{1}{3}(1+\sin^2\theta_{13})\,.
\end{equation}
We estimate NNLO perturbations of order $\sin^2\theta_{13}$ and thus $\sin^2\theta_{12}$ will receive non-negligible corrections.
We can impose an upper bound on $v_S/\Lambda$ by requiring that the correction to the TB value of $\sin^2\theta_{12}$ does not take it outside the experimental $3\sigma$ range: the maximal allowed deviation from the TB value is $0.05$ and from there we impose the bound $v_S/\Lambda<\mathcal{O}(0.23)$.
We consider now $m_D^{R\prime}$ (the NLO Dirac neutrino mass matrix in the basis of diagonal and real RH neutrinos). We can write:
\begin{equation}
\label{expmDRp}
m_D^{R\prime}=m_D^R+m_D^{(1)}U_{TB}D'\,,
\end{equation}
and calculate the relevant product for leptogenesis, $m_D^{R\prime\dag}m_D^{R\prime}$, keeping only the first terms in the expansion in the small parameter $v_S^2/\Lambda^2$:
\begin{equation}
\label{md2}
m_D^{R\prime\dag}m_D^{R\prime}=m_D^{R\dag}m_D^R+ \left(D^*U_{TB}^Tm_D^{(1)\dag}m_D^R+h.c.\right)\,,
\end{equation}
where in the second term the only off-diagonal entries are the $13$ and $31$ ones. In the case of inverted hierarchy for the effective neutrinos, the lightest RH neutrino is $N_2$. The summation in the numerator of eq. (\ref{eq:cp-asymm}) does not contain the term $13$ and therefore $\epsilon_{N_2}$ is vanishing also at NLO. This, however, does not mean leptogenesis can not be realized in this case. Since there is only a mild hierarchy between $N_2$, $N_1$ and $N_3$ and neither $\epsilon_{N_1}$ nor $\epsilon_{N_3}$ vanish, leptogenesis will proceed through $N_{1,3}$ dynamics. In the normal hierarchy (NH), the RH neutrino mass spectrum is $M_{N_3}<M_{N_2}<M_{N_1}$. There is a mild hierarchy between $N_3$ and $N_2$ while the hierarchy between $N_3$ and $N_1$ is large (around a factor 9). Consequently, the lepton asymmetry generated in $N_1$ decays will be, in general, erased by the lepton number violating interactions of $N_3$. Only $N_3$ dynamics becomes relevant for the generation of a lepton asymmetry in this case. Note that if the hierarchy between $N_1$ and $N_3$ decreases (as could be in the case of a quasi-degenerate spectrum), so it becomes mild, $N_1$ dynamics should be taken into account. Henceforth, for simplicity, we will consider only the NH case for which, according to eq. (\ref{eq:cp-asymm}), the CP-violating parameter $\epsilon_{N_3}$ can be written as
\begin{equation}
\label{expep}
\epsilon_{N_3}=\dfrac{1}{8\pi}\dfrac{1}{v_u^2\left(m_D^{R\,\dagger} \;m_D^R\right)_{11}}\mathbb{I}\mbox{m}\left[\left(\left(m_D^{R\,\dagger} \;m_D^R\right)_{13}\right)^2\right]f\left(\dfrac{M_1^2}{M_3^3}\right)\,.
\end{equation}
In the following figures we show a series of scatter plots related to the predictions of the model and the connections among low-energy observables and $\epsilon_{N_3}$. The points correspond only to the NH neutrino spectrum (in which we take $v_{S}/\Lambda\sim w/\Lambda=0.007\div0.23$, $\tan\beta=2\div50$ and we treat $y$, $y_1$, $y_2$ and $y_3$ as random numbers with modulus between $0.1$ and $2$).
\begin{figure}[ht!]
\centering
\includegraphics[width=3.8cm]{eN-delta13-num.jpg}
\includegraphics[width=3.8cm]{eN-delta23-num.jpg}
\includegraphics[width=3.8cm]{eN-delta12-num.jpg} \caption{Correlation between $\epsilon_{N_3}$ and $\sin^2\theta_{13}$, $\sin^2\theta_{23}$, $\sin^2\theta_{12}$. The horizontal line marks $\epsilon_{N_3}\sim 10^{-6}$, the vertical lines correspond to the central values and bounds at $1$ and $2~\sigma$ level of $\sin^2\theta_{13, 23, 12}$.}
\label{fig:eN-13-23}
\end{figure}
In figure~\ref{fig:eN-13-23} we plot $\epsilon_{N_3}$ and $\sin^2\theta_{13, 23, 12}$. As expected by comparing \eq{expep} with \eq{expang}, $\epsilon_{N_3}$ is correlated to all low-energy mixing angles.
\begin{figure}[ht!]
\centering
\includegraphics[width=3.8cm]{eN-deltalep-num.jpg}
\includegraphics[width=3.8cm]{eN-deltamaj-numnew.jpg}
\caption{The high energy $\epsilon_{N_3}$ versus low energy $\delta$ and $\Delta \phi_{12}$. The horizontal line marks to $\epsilon_{N_3}\sim 10^{-6}$.}
\label{fig:del-lep}
\end{figure}
In figure~\ref{fig:del-lep} we plot $\epsilon_{N_3}$ and the low-energy CP-Dirac and Majorana phases. It shows that the low energy CP-Dirac phase $\delta$ is not correlated to $\epsilon_{N_3}$. This result is not surprising: the phases that enter in $\epsilon_{N_3}$ are related to the phases present in $M_R$, while $\delta$ arises by the phases that appear in what we defined as $m_D^{(1)}$ in \eq{expmDRp}. In contrast the difference between the Majorana phases $\phi_1$ and $\phi_2$ ($\Delta \phi_{12}$) presents a correlation with
$\epsilon_{N_3}$: at LO the phenomenological analysis of this model shows that the NH spectrum can be reproduced only if $\Delta \phi^0_{12} $ is small. At LO, $\Delta\phi^0_{12}$ coincides with the corresponding Majorana phase difference $\Delta \phi_{12}^R$ of the right handed neutrinos. By perturbing the neutrino Dirac mass matrix we introduce new arbitrary phases that can vary in all the interval $(0,2 \pi)$. However the NLO contributions are responsible for deviating the lepton mixing angles away from TB values and also for slightly modifying the neutrino spectrum, in the range allowed by the data fit. This means that in general at NLO
\begin{equation}
m_i \sim m_i^0 +\delta m_i\,,
\end{equation}
where $\delta m_i$ are complex parameters and $m_i^0$ the neutrino mass eigenvalues at LO. Requiring now that $\Delta m^2_{12}$ is still in the range indicated for $\Delta m^2_{sol}$ we have that $\delta m\sim |\delta m_{1,2}| \sim 10^{-3}$ eV for $|m_1^0|,|m_2^0|\sim \mathcal{O}(\sqrt{\Delta m^2_{sol}})$. A straight computation shows then that the Majorana phase $\Delta \phi_{12}$ satisfies
\begin{equation}
\tan \Delta \phi_{12}\sim \tan \Delta \phi^0_{12}+ \alpha \frac{\delta m}{ \mathcal{O}(\sqrt{\Delta m^2_{sol}})}\,,
\end{equation}
where $\Delta \phi^0_{12}$ is the LO phase difference and $\alpha\in (0,1)$ a parameter that takes into account that the $\delta m_{1,2}$ phases run into the interval $(0, 2 \pi)$. We can estimate the maximal deviation of $\Delta\phi_{12}$ by its LO value getting
\begin{equation}
\Delta\phi_{12}- \Delta\phi^0_{12} \sim \frac{\pi}{10}\,.
\end{equation}
Notice that the left panel of fig.~\ref{fig:del-lep} shows that the majority of all the points are indeed inside the interval $(-\pi/8,\pi/8)$ in perfect agreement with our analytical results for a small LO $\Delta\phi^0_{12}\leq0.1$.
In conclusion, in the model considered it is possible to obtain correlations between low-energy observables and high-energy CP-violation, but it confirms that in general no correlation is present between high and low-energy CP-violating parameters.
There is correlation between deviations from TB angles (e.g. non-vanishing value of $\theta_{13}$) and the
value of the CP asymmetry parameter. We note that to have viable leptogenesis the model's allowed parameter space is rather constrained.
\begin{acknowledgement}
The work of IdMV was supported by FCT under the grant
SFRH/BPD/35919/2007 and
through the projects POCI/81919/2007, CERN/FP/83503/2008 and CFTP-FCT
UNIT 777 which are partially funded through POCTI (FEDER) and by the
Marie Curie RTN MRTN-CT-2006-035505.
\end{acknowledgement}
|
\section{Introduction}
Sensorless control of Permanent-Magnet Synchronous Motors (PMSM) at low velocity remains a challenging task. Most of the existing control algorithms rely on the motor saliency, both geometric and saturation-induced, for extracting the rotor position from the current measurements through high-frequency signal injection~\cite{Ji_Sul_2003,CorleL1998IToIA}. However some magnetic saturation effects such as cross-coupling and permanent magnet demagnetization can introduce large errors on the rotor position estimation~\cite{GugliPV2006IToIA,Bi_bo_2009}. These errors decrease the performance of the controller. In some cases they may cancel the rotor total saliency and lead to instability. It is thus important to correctly model the magnetic saturation effects, which is usually done through d-q magnetizing curves (flux versus current). These curves are usually found either by finite element analysis FEA or experimentally by integration of the voltage equation~\cite{StumSDHT2003IToIA, Stumb2005}. This provides a good way to characterize the saturation effects and can be used to improve the sensorless control of the PMSM~\cite{Zh_li_2007,Gugl_2004}. However the FEA or the integration of the voltage equation methods are not so easy to implement and do not provide an explicit model of the saturated PMSM.
In this paper a simple parametric model of the saturated PMSM is introduced (section~\ref{sec:model}); it is based on an energy function~\cite{BasicMR2010IToAC,Dur_2011} which simply encompasses the saturation and cross-magnetization effects.
In section~\ref{sec:Identi} a simple estimation method of the magnetic parameters is proposed and rigorously justified: fast-varying pulsating voltages are impressed to the motor with rotor locked; they create current ripples from which the magnetic parameters are estimated by linear least squares. In section~\ref{sec:experiment} experimental results on two kinds of motors (with surface-mounted and interior magnets) illustrate the relevance of the approach.
\section{An energy-based model for the saturated PMSM}\label{sec:model}
\subsection{Energy-based model}
The electrical subsystem of a two-axis PMSM expressed in the synchronous $d-q$ frame reads
\begin{align}
\label{eq:elecd}\frac{d\phi_d}{dt} &=u_d-Ri_d+\frac{d\theta}{dt}\phi_q\\
\label{eq:elecq}\frac{d\phi_q}{dt} &=u_q-Ri_q-\frac{d\theta}{dt}(\phi_d+\phi_m),
\end{align}
where $\phi_d,\phi_m$ are the direct-axis flux linkages due to the current excitation and to the permanent magnet, and $\phi_q$ is the quadrature-axis flux linkage; $u_d,u_q$ are the impressed voltages and $i_d,i_q$ are the currents; $\theta$ is the rotor (electrical) position and $R$ is the stator resistance.
The currents can be expressed in function of the flux linkages thanks to a suitable energy function~$\HH(\phi_d,\phi_q)$ by
\begin{align}
\label{eq:CurrentFluxd}i_d &=\partial_1\HH(\phi_d,\phi_q)\\
\label{eq:CurrentFluxq}i_q &=\partial_2\HH(\phi_d,\phi_q),
\end{align}
where $\partial_k\HH$ denotes the partial derivative w.r.t. the $k^{th}$ variable, see~\cite{BasicMR2010IToAC,Dur_2011}; without loss of generality $\HH(0,0)=0$.
For an unsaturated PMSM this energy function reads
\begin{align*}
\HH_l(\phi_d,\phi_q) &=\frac{1}{2L_d}\phi^2_d+\frac{1}{2L_q}\phi^2_q
\end{align*}
where $L_d$ and $L_q$ are the motor self-inductances, and we recover the usual linear relations
\begin{align*}
i_d &=\partial_1\HH(\phi_d,\phi_q) =\frac{\phi_d}{L_d}\\
i_q &=\partial_2\HH(\phi_d,\phi_q) =\frac{\phi_q}{L_q}.
\end{align*}
Notice the expression for $\HH$ should respect the symmetry of the PMSM w.r.t the direct axis, i.e.
\begin{equation}\label{eq:sym}
\HH(\phi_d,-\phi_q)=\HH(\phi_d,\phi_q),
\end{equation}
which is obviously the case for~$\HH_l$.
Indeed, \eqref{eq:elecd}-\eqref{eq:elecq} is left unchanged by the transformation
$$(\phi_d',u_d',i_d',\phi_q',u_q',i_q',\theta'):=(\phi_d,u_d,i_d,-\phi_q,-u_q,-i_q,-\theta);$$
this implies
\begin{align*}
\partial_1\HH(\phi_d',\phi_q') &=\partial_1\HH(\phi_d,\phi_q)\\
\partial_2\HH(\phi_d',\phi_q') &=-\partial_2\HH(\phi_d,\phi_q),
\end{align*}
i.e.
\begin{align*}
\partial_1\HH(\phi_d,-\phi_q) &=\partial_1\HH(\phi_d,\phi_q)\\
\partial_2\HH(\phi_d,-\phi_q) &=-\partial_2\HH(\phi_d,\phi_q).
\end{align*}
Therefore
\begin{align*}
\frac{d\HH}{d\phi_d}(\phi_d,-\phi_q) &=\partial_1\HH(\phi_d,-\phi_q)\\
&=\partial_1\HH(\phi_d,\phi_q)\\
&=\frac{d\HH}{d\phi_d}(\phi_d,\phi_q)\\
\frac{d\HH}{d\phi_q}(\phi_d,-\phi_q) &=-\partial_2\HH(\phi_d,-\phi_q)\\
&=\partial_2\HH(\phi_d,\phi_q)\\
&=\frac{d\HH}{d\phi_q}(\phi_d,\phi_q).
\end{align*}
Integrating these relations yields
\begin{align*}
\HH(\phi_d,-\phi_q) &=\HH(\phi_d,\phi_q)+c_d(\phi_q)\\
\HH(\phi_d,-\phi_q) &=\HH(\phi_d,\phi_q)+c_q(\phi_d),
\end{align*}
where $c_d,c_q$ are functions of only one variable. But this makes sense only if $c_d(\phi_q)=c_q(\phi_d)=c$ with $c$ constant. Since $H(0,0)=0$, $c=0$, which yields~\eqref{eq:sym}.
\subsection{Parametric description of magnetic saturation}\label{sec:paramsat}
Magnetic saturation can be accounted for by considering a more complicated magnetic energy function~$\HH$, having $\HH_l$ for quadratic part but including also higher-order terms. From experiments saturation effects are well captured by considering only third- and fourth-order terms, hence
\begin{multline*}
\HH(\phi_d,\phi_q)=\HH_l(\phi_d,\phi_q)\\
+\sum_{i=0}^3\alpha_{3-i,i}\phi_d^{3-i}\phi_q^i+\sum_{i=0}^4\alpha_{4-i,i}\phi_d^{4-i}\phi_q^i.
\end{multline*}
This is a perturbative model where the higher-order terms appear as corrections of the dominant term~$\HH_l$.
The $9$ coefficients $\alpha_{ij}$
together with $L_d$, $L_q$ are motor dependent. But \eqref{eq:sym} implies $\alpha_{2,1}=\alpha_{0,3}=\alpha_{3,1}=\alpha_{1,3}=0$, so that the energy function eventually reads
\begin{multline}\label{eq:EnerSat}
\HH(\phi_d,\phi_q) =\HH_l(\phi_d,\phi_q) +\alpha_{3,0}\phi_d^3+\alpha_{1,2}\phi_d\phi_q^2 \\
+\alpha_{4,0}\phi_d^4+\alpha_{2,2}\phi_d^2\phi_q^2+\alpha_{0,4}\phi_q^4.
\end{multline}
From~\eqref{eq:CurrentFluxd}-\eqref{eq:CurrentFluxq} and~\eqref{eq:EnerSat} the currents are then explicitly given by
\begin{align}
\notag i_d &=\partial_1\HH(\phi_d,\phi_q)\\
\label{eq:id}&=\frac{\phi_d}{L_d}+3\alpha_{3,0}\phi_d^2+\alpha_{1,2}\phi_q^2 +4\alpha_{4,0}\phi_d^3+2\alpha_{2,2}\phi_d\phi_q^2\\
\notag i_q &=\partial_2\HH(\phi_d,\phi_q)\\
\label{eq:iq}&=\frac{\phi_q}{L_q}+2\alpha_{1,2}\phi_d\phi_q+2\alpha_{2,2}\phi_d^2\phi_q+4\alpha_{0,4}\phi_q^3,
\end{align}
which are the flux-current magnetization curves. Fig.~\ref{fig:iphi} shows examples of these curves in the more familiar presentation of fluxes w.r.t currents obtained by numerically inverting~\eqref{eq:CurrentFluxd}-\eqref{eq:CurrentFluxq}; the motor is the IPM of section~\ref{sec:experiment}.
The model of the saturated PMSM is thus given by \eqref{eq:elecd}-\eqref{eq:elecq} and~\eqref{eq:id}-\eqref{eq:iq}. It is in state form with $\phi_d,\phi_q$ as state variables. The magnetic saturation effects are represented by the $5$ additional parameters $\alpha_{3,0}, \alpha_{1,2}, \alpha_{4,0}, \alpha_{2,2}, \alpha_{0,4}$.
\begin{figure}[t!]\centering
\subfigure[$\phi_d(i_d,i_q=Constant)$]{\includegraphics[width=\columnwidth]{idphid}\label{fig:idphid}}
\subfigure[$\phi_q(i_d=Constant,i_q)$]{\includegraphics[width=\columnwidth]{iqphiq}\label{fig:iqphiq}}
\caption{Flux-current magnetization curves (IPM)}\label{fig:iphi}
\end{figure}
\subsection{Model with $i_d,i_q$ as state variables}
The model of the saturated PMSM is often expressed with $i_d,i_q$ as state variables, e.g.~\cite{StumSDHT2003IToIA}. Starting with flux-current magnetization curves in the form
\begin{align}
\label{eq:FluxCurrentd}\phi_d &=\Phi_d(i_d,i_q)\\
\label{eq:FluxCurrentq}\phi_q &=\Phi_q(i_d,i_q)
\end{align}
and differentiating w.r.t time,
\eqref{eq:elecd}-\eqref{eq:elecq} then becomes
\begin{align*}
L_{dd}(i_d,i_q)\frac{di_d}{dt}+L_{dq}(i_d,i_q)\frac{di_q}{dt} &=u_d-Ri_d+\frac{d\theta}{dt}\phi_q\\
L_{qd}(i_d,i_q)\frac{di_d}{dt}+L_{qq}(i_d,i_q)\frac{di_q}{dt} &=u_q-Ri_q-\frac{d\theta}{dt}(\phi_d+\phi_m),
\end{align*}
where
\begin{align*}
\begin{pmatrix}L_{dd}(i_d,i_q) & L_{dq}(i_d,i_q)\\ L_{qd}(i_d,i_q)& L_{qq}(i_d,i_q)\end{pmatrix}
=&\begin{pmatrix}\partial_1\Phi_d(i_d,i_q) & \partial_2\Phi_d(i_d,i_q)\\
\partial_1\Phi_q(i_d,i_q) & \partial_2\Phi_q(i_d,i_q)\end{pmatrix}.
\end{align*}
Though not always acknowledged $L_{dq}$ and $L_{qd}$ should be equal. Indeed, plugging \eqref{eq:CurrentFluxd}-\eqref{eq:CurrentFluxq} into \eqref{eq:FluxCurrentd}-\eqref{eq:FluxCurrentq} gives
\begin{align*}
\phi_d &=\Phi_d\bigl(\partial_1\HH(\phi_d,\phi_q),\partial_2\HH(\phi_d,\phi_q)\bigr)\\
\phi_q &=\Phi_q\bigl(\partial_1\HH(\phi_d,\phi_q),\partial_2\HH(\phi_d,\phi_q)\bigr).
\end{align*}
Taking the total derivative of both sides of these equations w.r.t. $\phi_d$ and $\phi_q$ then yields
\begin{align*}
\begin{pmatrix}1&0\\0&1\end{pmatrix}
=&\begin{pmatrix}L_{dd}\partial_{11}\HH+L_{dq}\partial_{12}\HH & L_{dd}\partial_{21}\HH+L_{dq}\partial_{22}\HH\\
L_{qd}\partial_{11}\HH+L_{qq}\partial_{12}\HH & L_{qd}\partial_{21}\HH+L_{qq}\partial_{22}\HH\end{pmatrix}\\
=&\begin{pmatrix}L_{dd} & L_{dq}\\ L_{qd} & L_{qq}\end{pmatrix}
\begin{pmatrix}\partial_{11}\HH & \partial_{21}\HH\\ \partial_{12}\HH & \partial_{22}\HH\end{pmatrix}.
\end{align*}
Since $\partial_{12}\HH=\partial_{21}\HH$ the second matrix in the last line is symmetric, hence the first; in other words $L_{dq}=L_{qd}$.
To do that with the model of section~\ref{sec:paramsat} the nonlinear equations \eqref{eq:id}-\eqref{eq:iq} must be inverted. Rather than doing that exactly, we take advantage of the fact the coefficients $\alpha_{i,j}$ are experimentally small. At first order w.r.t. the~$\alpha_{i,j}$ we obviously have $\phi_d=L_di_d+\OO(\abs{\alpha_{i,j}})$ and $\phi_q=L_qi_q+\OO(\abs{\alpha_{i,j}})$. Plugging these expressions into~\eqref{eq:id}-\eqref{eq:iq} we easily find
\begin{align}
\notag\phi_d &=L_d\bigl(i_d-3\alpha_{3,0}L_d^2i_d^2 -\alpha_{1,2}L_q^2i_q^2-4\alpha_{4,0}L_d^3i_d^3\\
\label{eq:phid}&\quad-2\alpha_{2,2}L_dL_q^2i_di_q^2\bigr)+\OO(\abs{\alpha_{i,j}}^2)\\
\notag\phi_q &=L_q\bigl(i_q-2\alpha_{1,2}L_dL_qi_di_q-2\alpha_{2,2}L_d^2L_qi_d^2i_q\\
\label{eq:phiq}&\quad-4\alpha_{0,4}L_q^3i_q^3\bigr)+\OO(\abs{\alpha_{i,j}}^2).
\end{align}
Finally,
\begin{align*}
L_{dd}(i_d,i_q) &=L_d\bigl(1-6\alpha_{3,0}L_di_d-12\alpha_{4,0}L_d^2i_d^2-2\alpha_{2,2}L_q^2i_q^2\bigr)\\
L_{dq}(i_d,i_q) &=L_{qd}(i_d,i_q)=-2L_dL_q^2i_q(\alpha_{1,2}+2\alpha_{2,2}L_di_d)\\
L_{qq}(i_d,i_q) &=L_q\bigl(1-2\alpha_{1,2}L_di_d-2\alpha_{2,2}L_d^2i_d^2-12\alpha_{0,4}L_q^2i_q^2\bigr).
\end{align*}
\section{A procedure for estimating the magnetic parameters}\label{sec:Identi}
\subsection{Principle}\label{sec:principle}
\begin{figure}[t!]
\centerline{\includegraphics[width=1\columnwidth]{currentHFValue}}
\caption{Experimental illustration of equation~\eqref{eq:iprocd}: time response of~$i_d$}\label{fig:HFcurrent}
\end{figure}
To estimate the $7$ magnetic parameters in the model, we propose a procedure which is rather easy to implement and reliable. With the rotor locked in the position~$\theta=0$, we inject fast-varying pulsating voltages
\begin{align}
\label{eq:uprocd}u_d(t) &=\bar u_d+\widetilde u_df(\W t)\\
\label{eq:uprocq}u_q(t) &=\bar u_q+\widetilde u_qf(\W t),
\end{align}
where $\bar u_d,\bar u_q,\widetilde u_d,\widetilde u_q,\W$ are constant and $f$ is a periodic function with zero mean. The pulsation $\W$ is chosen large enough w.r.t. the motor electric time constant. It can then be shown, see section~\ref{sec:justification}, that after an initial transient
\begin{align}
\label{eq:iprocd}i_d(t) &=\bar i_d+\widetilde i_dF(\W t)+\OO(\tfrac{1}{\W^2})\\
\label{eq:iprocq}i_q(t) &=\bar i_q+\widetilde i_qF(\W t)+\OO(\tfrac{1}{\W^2}),
\end{align}
where $\bar i_d=\frac{\bar u_d}{R},\bar i_q=\frac{\bar u_q}{R},\widetilde i_d,\widetilde i_q$ are constant and $F$ is the primitive of~$f$ with zero mean ($F$ has clearly the same period as~$f$); fig.~\ref{fig:HFcurrent} shows for instance the current $i_d$ obtained for the SPM of section~\ref{sec:experiment} when starting from $\id(0)=0$ and applying a square signal $u_d$ with $\Omega=500Hz$, $\bar u_d=23V$ and~$\widetilde u_d=30V$.
On the other hand using the saturation model the amplitudes $\widetilde i_d,\widetilde i_q$ of the current ripples turn out to be
\begin{align}
\notag\widetilde i_d &=\frac{1}{\W}\Bigl(\frac{\widetilde u_d}{L_d}
+2\alpha_{2,2}L_q\bar i_q(2L_d\bar i_d\widetilde u_q+L_q\bar i_q\widetilde u_d)\\
\label{eq:tid}&\quad+12\alpha_{4,0}L_d^2\bar i_d^2\widetilde u_d
+6\alpha_{3,0}L_d\bar i_d\widetilde u_d+2\alpha_{1,2}L_q\bar i_q\widetilde u_q\Bigr)\\
\notag\widetilde i_q &=\frac{1}{\W}\Bigl(\frac{\widetilde u_q}{L_q}
+2\alpha_{2,2}L_d\bar i_d(2L_q\bar i_q\widetilde u_d+L_d\bar i_d\widetilde u_q)\\
\label{eq:tiq}&\quad+12\alpha_{0,4}L_q^2\bar i_q^2\widetilde u_q
+2\alpha_{1,2}(L_d\bar i_d\widetilde u_q+L_q\bar i_q\widetilde u_d)\Bigr).
\end{align}
As $\widetilde i_d,\widetilde i_q$ can easily be measured experimentally, these expressions provide a means to identify the magnetic parameters from experimental data obtained with various values of~$\bar u_d,\bar u_q,\widetilde u_d,\widetilde u_q$.
\subsection{Estimation of the parameters}
Since combinations of the magnetic parameters always enter~\eqref{eq:tid}-\eqref{eq:tiq} linearly, they can be estimated by simple linear least squares; moreover by suitably choosing $\bar u_d,\bar u_q,\widetilde u_d,\widetilde u_q$, the whole least squares problem for the $7$ parameters can be split into several problems involving fewer parameters:
\begin{itemize}
\item with $\bar u_d=\bar u_q=0$, hence $\bar i_d=\bar i_q=0$, and $\widetilde u_d=0$ (resp.~$\widetilde u_q=0$) equation~\eqref{eq:tid} (resp. equation~\eqref{eq:tiq}) reads
\begin{equation}\label{eq:LdLq}
L_d=\frac{1}{\W}\frac{\widetilde u_d}{\widetilde i_d}
\qquad\Bigl(\text{resp.~~}L_q=\frac{1}{\W}\frac{\widetilde u_q}{\widetilde i_q}\Bigr)
\end{equation}
\item with $\bar u_q=0$, hence $\bar i_q=0$, and $\widetilde u_q=0$, \eqref{eq:tid} reads
\begin{equation}\label{eq:saturation1}
\widetilde i_d=\frac{\widetilde u_d}{\W}\left(\frac{1}{L_d}+6\alpha_{3,0}L_d\bar i_d+12\alpha_{4,0}L_d^2\bar i_d^2\right).
\end{equation}
Notice \eqref{eq:tiq} reads $\widetilde i_q=0$ hence provides no information
\item with $\bar u_d=0$, hence $\bar i_d:=0$, and $\widetilde u_q=0$, \eqref{eq:tid}-\eqref{eq:tiq} read
\begin{align}
\label{eq:saturation3d}\widetilde i_d &=\frac{\widetilde u_d}{\W}\Bigl(\frac{1}{L_d}+2\alpha_{2,2}L_q^2\bar i_q^2\Bigr)\\
\label{eq:saturation3q}\widetilde i_q & =\frac{2\widetilde u_d}{\W}\alpha_{1,2}L_q\bar i_q
\end{align}
\item with $\bar u_d=0$, hence $\bar i_d:=0$, and $\widetilde u_d=0$, \eqref{eq:tid}-\eqref{eq:tiq} read
\begin{align}
\label{eq:saturation4d}\widetilde i_d &=\frac{2\widetilde u_q}{\W}\alpha_{1,2}L_q\bar i_q\\
\label{eq:saturation4q}\widetilde i_q &=\frac{\widetilde u_q}{\W}\Bigl(\frac{1}{L_q}+12\alpha_{0,4}L_q^2\bar i_q^2\Bigr).
\end{align}
\end{itemize}
$L_d$ (resp.~$L_q$) is then immediately determined from~\eqref{eq:LdLq}; $\alpha_{3,0}$ and $\alpha_{4,0}$ are jointly estimated by least squares from~\eqref{eq:saturation1}; $\alpha_{2,2}$, $\alpha_{1,2}$ and $\alpha_{0,4}$ are separately estimated by least squares from respectively \eqref{eq:saturation3d}, \eqref{eq:saturation3q}-\eqref{eq:saturation4d} and~\eqref{eq:saturation4q}.
\subsection{Justification of section~\ref{sec:principle}}\label{sec:justification}
The assertions of section~\ref{sec:principle} can be rigorously justified by a straightforward application of second-order averaging of differential equations~\cite[p.~40]{SandeVM2007book}. Indeed the electrical subsystem \eqref{eq:elecd}-\eqref{eq:elecq} with locked rotor (i.e. $\frac{d\theta}{dt}=0$) and input voltages~\eqref{eq:uprocd}-\eqref{eq:uprocd} reads when setting $\tau=\W t$
\begin{align}
\label{eq:dynaved}\frac{d\phi_d}{d\tau} & =\frac{1}{\W}\bigl(\bar u_d+\widetilde u_df(\tau)-Ri_d(\phi_d,\phi_q)\bigr)\\
\label{eq:dynaveq}\frac{d\phi_q}{d\tau} & =\frac{1}{\W}\bigl(\bar u_q+\widetilde u_qf(\tau)-Ri_q(\phi_d,\phi_q)\bigr).
\end{align}
This system is in the so-called standard form for averaging, with a right hand-side periodic in~$\tau$ and $\frac{1}{\W}$ as a small parameter. Therefore its solution is given by
\begin{align}
\label{eq:phiprocd}\phi_d(\tau) &=\phi_d^0(\tau)+\frac{\widetilde u_d}{\W}F(\tau)+\OO(\tfrac{1}{\W^2})\\
\label{eq:phiprocq}\phi_q(\tau) &=\phi_q^0(\tau)+\frac{\widetilde u_q}{\W}F(\tau) +\OO(\tfrac{1}{\W^2}),
\end{align}
where $(\phi_d^0,\phi_q^0)$ is the solution of the system
\begin{align*}
\frac{d\phi_d^0}{dt} &=\bar u_d-Ri_d(\phi_d^0,\phi_q^0)\\
\frac{d\phi_q^0}{dt} &=\bar u_q-Ri_q(\phi_d^0,\phi_q^0)
\end{align*}
obtained by averaging the right-hand side of~\eqref{eq:dynaved}-\eqref{eq:dynaveq}. After an initial transient
$\bigl(\phi_d^0(\tau),\phi_q^0(\tau)\bigr)$ asymptotically reaches the constant value $(\bar\phi_d,\bar\phi_q)$ determined by $\bar u_d=Ri_d(\bar\phi_d,\bar\phi_q)$ and $\bar u_q=Ri_q(\bar\phi_d,\bar\phi_q)$.
Plugging~\eqref{eq:phiprocd}-\eqref{eq:phiprocq} with $t=\frac{\tau}{\W}$ into~\eqref{eq:id}-\eqref{eq:iq}, and expanding along powers of $\frac{1}{\W}$ then yields
\begin{align*}
i_d(t) &=\bar i_d+\frac{F(\W t)}{\W}\biggr(\frac{\widetilde u_d}{L_d}+6\alpha_{3,0}\bar\phi_d\widetilde u_d+2\alpha_{1,2}\bar\phi_q\widetilde u_q\\
&\quad+12\alpha_{4,0}\bar\phi_d^2\widetilde u_d+2\alpha_{2,2}(2\bar\phi_d\bar\phi_q\widetilde u_q+\bar\phi_q^2\widetilde u_d)\biggl)+\OO(\tfrac{1}{\W^2})\\
i_q(t) &=\bar i_q+\frac{F(\W t)}{\W}\biggl(\frac{\widetilde u_q}{L_q}+2\alpha_{1,2}(\bar\phi_d\widetilde u_q+\bar\phi_q\widetilde u_d)\\
&\quad+2\alpha_{2,2}(2\bar\phi_d\bar\phi_q\widetilde u_d+\bar\phi_d^2\widetilde u_q)
+12\alpha_{0,4}\bar\phi_q^2\widetilde u_q\biggr)
+\OO(\tfrac{1}{\W^2}),
\end{align*}
where $\bar i_d=i_d(\bar\phi_d,\bar\phi_q)$ and $\bar i_q=i_q(\bar\phi_d,\bar\phi_q)$
There remains to express $\bar\phi_d,\bar\phi_q$ in function of $\bar i_d,\bar i_q$. Rather than exactly inverting the nonlinear equations~\eqref{eq:id}-\eqref{eq:iq}, we take advantage of the fact the coefficients $\alpha_{i,j}$ are experimentally small. At first order w.r.t. the~$\alpha_{i,j}$ we have $\phi_d=L_di_d+\OO(\abs{\alpha_{i,j}})$ and $\phi_q=L_qi_q+\OO(\abs{\alpha_{i,j}})$. Using this in the previous equations and neglecting $\OO(\tfrac{1}{\W^2})$ and $\OO(\abs{\alpha_{i,j}}^2)$ terms we eventually find~\eqref{eq:iprocd}-\eqref{eq:tiq}. Using directly \eqref{eq:phid}-\eqref{eq:phiq} yields of course the same result.
\section{Experimental Results}\label{sec:experiment}
\subsection{Experimental setup}
The methodology of section~\ref{sec:Identi} is tested on an interior magnet PMSM (IPM) and a surface-mounted PMSM (SPM) with rated parameters listed below.
The setup consists of an industrial inverter with a $400V$ DC bus and a $4kHz$ PWM switching frequency, 3 dSpace boards (DS1005 PPC Board, DS2003 A/D Board, DS4002 Timing and Digital I/O Board) and a host PC. The measurements were sampled also at $4kHz$.
\begin{center}
\begin{tabular}{| l | l | l | }
\firsthline
& IPM & SPM \\
\hline
Pole pairs & 6 & 2 \\
\hline
Rated power & $200~W$ & $1200~W$ \\
\hline
Rated current & $1.2~A$ & $3.4~A$ \\
\hline
Rated speed & $1800~rpm$ & $400~rpm$ \\
\hline
Rated torque & $1.06~N.m$ & $29~N.m$ \\
\hline
Resistance & $12.15~\W$ & $6.69~\W$ \\
\lasthline
\end{tabular}
\end{center}
\subsection{Experimental results}
With the rotor locked in the position~$\theta=0$, a square wave voltage with frequency $\W=500Hz$ and constant amplitude $\widetilde u_d$ or $\widetilde u_q$ ($30V$ for the IPM, $40V$ for the SPM) is applied to the motor. But for the determination of $L_d,L_q$ where $\bar u_d=\bar u_q=0$, several runs are performed with various $\bar u_d$ (resp.~$\bar u_q$) such that $\bar i_d$ (resp.~$\bar i_q$) ranges from $-2A$ to $+2A$ with a $0.3A$ increment (IPM), or from $-8A$ to $8A$ with a $0.5A$ increment (SPM). The estimated parameters are listed below; the uncertainty in the estimation stems from a $\pm10mA$ uncertainty in the current measurements.
\begin{center}
\begin{tabular}{|l | l | l | }
\firsthline
& IPM & SPM \\
\hline
$L_d~(mH$) & $91.9\pm5$ & $155.4\pm10$ \\
\hline
$L_q~(mH)$ & $45.8\pm1$ & $58.6\pm2$ \\
\hline
$\alpha_{3,0}~(A.Wb^{-2})$ & $7.70\pm0.11$ & $5.01\pm0.11$ \\
\hline
$\alpha_{1,2}~(A.Wb^{-2})$ & $5.35\pm0.61$ & $4.83\pm0.27$ \\
\hline
$\alpha_{4,0}~(A.Wb^{-3})$ & $19.42\pm1.34$ & $1.83\pm0.28$ \\
\hline
$\alpha_{2,2}~(A.Wb^{-3})$ & $22.18\pm2.80$ & $8.76\pm1.03$ \\
\hline
$\alpha_{0,4}~(A.Wb^{-3})$ & $6.62\pm0.42$ & $1.18\pm0.17$ \\
\lasthline
\end{tabular}
\end{center}
The good agreement between the fitted curves and the measurements is demonstrated for instance for~\eqref{eq:saturation1} on~Fig.~\ref{fig:Id_id_ud} and for~\eqref{eq:saturation3q} on~Fig.~\ref{fig:Iq_iq_ud}. Notice \eqref{eq:saturation1} illustrates saturation on a single axis, while \eqref{eq:saturation3q} illustrates cross-saturation.
\begin{figure}[t!]
\centering
\subfigure[IPM]{\includegraphics[height=67mm]{Id_id_ud2}\label{fig:Id_id_ud2}}
\subfigure[SPM]{ \includegraphics[height=67mm]{Id_id_ud1}\label{fig:Id_id_ud1}}
\caption{Measured values (circles) and fitted curve (solid line) for~\eqref{eq:saturation1}.}\label{fig:Id_id_ud}
\end{figure}
\begin{figure}[t!]
\centering
\subfigure[IPM]{\includegraphics[height=67mm]{Iq_iq_ud2}\label{fig:Iq_iq_ud2}}
\subfigure[SPM]{\includegraphics[height=67mm]{Iq_iq_ud1}\label{fig:Iq_iq_ud1}}
\caption{Measured values (circles) and fitted curve (solid line) for~\eqref{eq:saturation3q}.}\label{fig:Iq_iq_ud}
\end{figure}
\begin{figure}[t!]
\centering
\subfigure[Interior-magnet PMSM]{
\includegraphics[width=\columnwidth]{i60_ud2}
\label{fig:Id_i60_ud2}
}
\subfigure[Surface mounted PMSM]{
\includegraphics[width=\columnwidth]{i60_ud1}
\label{fig:Id_i60_ud1}
}
\caption{Measured values (circles) compared to model-predicted values (solid line) for a $60^\circ$ current angle.
}\label{fig:i60_ud}
\end{figure}
\subsection{Validation}
The estimation procedure relies on~\eqref{eq:saturation1}--\eqref{eq:saturation4q}, with either $\bar i_d\neq$ or , i.e. current vectors with angles $0^\circ,90^\circ,180^\circ,270^\circ$. To check the validity of the model tests were conducted with current vectors with various angles and magnitudes on the whole operating ($|i|=\sqrt{i_d^2+i_q^2}$ ranging from $0A$ to $2A$ with a $0.3A$ increment for the IPM, and from $0A$ to $5.5A$ with a $0.5A$ increment for the SPM).
Fig.~\ref{fig:i60_ud} shows for instance the results for a $60^\circ$ current angle; there is a good agreement between the measured values and those predicted by the model.
As a kind of cross-validation we also examined the currents time responses to large voltage steps. Fig.~\ref{fig:step} shows the good agreement between the measurements and the time response obtained by simulating the model with the estimated parameters; it also shows the differences with the simulated response when the saturation effects are omitted.
Fig.~\ref{fig:phididm} shows the good agreement also between the ``measured'' flux values (i.e. obtained by integrating the measured currents and voltages) and the flux values obtained by simulation.
\begin{figure}[t!]
\centerline{\includegraphics[width=\columnwidth]{Step3}}
\caption{Time response of $i_d$ to large step voltages in $u_d$ (IPM).}\label{fig:step}
\end{figure}
\begin{figure}[t!]
\centerline{\includegraphics[width=\columnwidth]{phididm}}
\caption{Saturation curve $\phi_d-i_d$ (IPM).}\label{fig:phididm}
\end{figure}
\section{Conclusion}
A simple parametric magnetic saturation model for the PMSM with a simple identification procedure based on high-frequency voltage injection have been introduced. Experimental tests on two kinds of PMSM (IPM and SPM) demonstrate the relevance of the approach. This model can be fruitfully used to design a sensorless control scheme at low velocity.
|
\section{Introduction}
Over the past fifty years the Anderson model of localization has
become a paradigm for investigations of electronic transport in the
presence of static disorder \cite{And58,KraM93}. Depending on the
strength of the disorder the wave functions of non-interacting
electrons vary from delocalized to exponentially localized. This so
called Anderson localization was demonstrated recently in experiments
with matter waves of Bose-Einstein condensates in random or
quasi-periodic optical potentials \cite{BilJZB08,RoaDFF08}. Due to
significant electron-electron and electron-phonon interactions in
realistic materials, e.g. imperfect crystals, the direct observation
of Anderson localization is rather difficult. Therefore the success
of the model is driven to a large extent by the universal properties
of the resulting phenomena.
The characteristics of the disorder potential have a prominent
influence on Anderson localization \cite{KraM93}. For one dimensional
(1D) systems with a spatially uncorrelated potential it has been shown
that all electronic states are exponentially localized \cite{GolMP77}.
Later it was realized that by introducing correlations, one can partly
suppress the localization, at least for weak disorder \cite{IzrK99}.
Moreover, there have been discussions of a delocalization-localization
transition even in 1D for long-range correlated disorder potentials
\cite{deML98, ShiNN04, Kay07}. However, it has early been recognized
that the apparent transition results from a rescaling of the disorder
potential by normalizing the variance, which becomes system-size
dependent for the considered self-affine potential landscapes
\cite{KanRBH00}.
In the present work, we focus on long-range power law correlated
potentials with a correlation function $C(\ell) \propto
\ell^{-\alpha}$ and correlation exponents $\alpha > 0$. The variance
of the resulting potential is independent of the system size. This so
called scale-free disorder is by no means artificial. It appears
naturally in a large variety of physical systems
\cite{Isi92,PenBGH92,VidMNM96}.
The outline of this article is as follows. In the next section we
briefly review the main results for Anderson localization in 1D and in
the presence of correlated disorder. We also discuss the numerical
methods used in our calculations. In Sec.\ \ref{sec:results} we
present and discuss the obtained results. Finally, the article is
summarized in Sec.\ \ref{sec:summary}.
\section{Model and Numerical Techniques}
\subsection{Anderson Model of Localization with Long-Range Correlated Disorder}
The Anderson model \cite{And58,KraM93} is widely used to investigate
the phenomenon of localization in disordered materials. It is based
upon a tight-binding Hamiltonian in site representation,
\begin{equation}
\mathcal{H} = \sum_{i} \varepsilon_i \ket{i}\bra{i}
- t\;\sum_{\mean{i\,j}}\,\ket{i}\bra{j}\;,
\label{eq:AndersonHam}
\end{equation}
where $\ket{i}$ is a localized state at lattice site $i$ and
$\mean{i\,j}$ denotes a restriction of the sum to $i$ and $j$ being
nearest neighbors. The hopping parameter $t=\hbar^2/(2 m^* a^2)$ is
defined in terms of the lattice spacing $a$ and the effective mass
$m^*$. In the following, we set $t\equiv 1$ and thus fix the unit of
energy; lengths are measured in units of $a$. The on-site potentials
$\varepsilon_i$ are random numbers, chosen according to some
probability distribution $P(\varepsilon)$ characterized by the mean
$\mean{\varepsilon_i}$ and the correlation function
$\mean{\varepsilon_i \varepsilon_{i+\ell}} = C(\ell)$. However,
usually the site energies are taken to be statistically independent.
For example, a common choice of $P(\varepsilon)$ is a Gaussian white
noise distribution or a box distribution of width $W$, both with
$\mean{\varepsilon_i}=0$ and $C(\ell)=\frac{W^2}{12} \delta_{\ell,0}$.
Other distributions have also been considered
\cite{BulSK87,KraM93,OhtSK99,RomS03}.
In the present work we are interested in the influence of long-range
correlated disorder potentials on the Anderson model, i.e., using a
correlation function of the form
\begin{equation}\label{eq:PLCorrFunc}
C(\ell) \equiv \mean{\varepsilon_i \varepsilon_{i+\ell}} \propto |\ell|^{-\alpha}
\end{equation}
where $\alpha$ is the correlation exponent which determines the strength of the
correlations. The associated spectral density is given by
\begin{equation}\label{eq:SpecDens}
S(q) = \sum_\ell C(\ell) e^{-\imath q \ell} \propto |q|^{\alpha - 1 }\;.
\end{equation}
In order to generate physically reasonable long-ranged correlated
potentials, the correlation function \eqref{eq:PLCorrFunc} should
decay with distance $\ell$ and therefore $\alpha>0$. For 1D systems
the correlations are considered to be long ranged for $0<\alpha<1$.
The value $\alpha = 1$ corresponds to the case of uncorrelated
disorder. Notice that the delocalization-localization transition
observed in Refs.\ \cite{deML98, ShiNN04, Kay07} relies on power-law
correlations with $\alpha<0$. These correlations increase with
distance $\ell$ and lead to a system-size dependence of the disorder
potential. Thus a rescaling of the results is required which yields
the corrected localization behavior without transition
\cite{KanRBH00}.
In general, it is extremely complicated to obtain analytical results
of transport properties for the Anderson model of localization. For
example, only in the 1D case rigorous proofs of strong
localization for all energies and disorder strengths have been given
\cite{GolMP77}. Moreover, the explicit energy and disorder dependence
of the localization length $\Lambda$ for weak disorder has been
derived \cite{Tho79,PasF92}. There are also some results for 1D
systems with long-range correlated disorder. For energies close to the
band center ($|E| \ll 2$) a weak disorder expansion for $\Lambda$
yields \cite{IzrK99},
\begin{equation}
\Lambda\left(q, W\right) = \left[ \frac{1}{8 \sin^2(q)} \frac{W^2}{12} S(2 q) \right]^{-1}\;,
\label{eq:1DWeakDisExpansion}
\end{equation}
where $E(q) = -2 \cos(q)$. It is important to notice that even for the
correlated case the localization length remains proportional to
$W^{-2}$. Thus, for small disorder strength and in the vicinity of the
band center, the presence of correlations leads to an {\it effective}
disorder strength $W_{\rm eff} = W \sqrt{S(2 q)}$, which depends on
energy via the spectral density $S$.
On the other hand, close to the unperturbed band edge ($|E|=2$) a
different universal behavior can be found \cite{CohES85, Rus02}. Here,
the density of states (DOS) and the localization length are characterized by
units of energy $\epsilon$ and length $\lambda$,
\begin{subequations}\label{eq:BEScaling}
\begin{eqnarray}
\rho_\alpha(E) &=& \lambda^{-1} \epsilon^{-1} \, f_ \alpha( E/\epsilon )\;, \\
\Lambda_\alpha(E) &=& \lambda \, g_ \alpha( E/\epsilon )\;,
\end{eqnarray}
\end{subequations}
where $f_\alpha$ and $g_\alpha$ are universal functions. For
uncorrelated potentials, i.e.\, $\alpha = 1$, there is a closed
analytic form available for both functions \cite{DerG84}. The units
$\epsilon$ and $\lambda$ can be expressed in terms of the hopping
parameter $t$ and the disorder strength $w = W/\sqrt{12}$,
\begin{subequations}
\begin{eqnarray}
\epsilon &=& w^{4/(4-\alpha)}\, \\
\lambda &=& w^{-2/(4-\alpha)}\;.
\end{eqnarray}
\end{subequations}
In particular, it follows that at the unperturbed band edge ($|E|=2$),
\begin{subequations} \label{eq:1DPowerDecay}
\begin{eqnarray}
\Lambda_\alpha( |E|=2, W) &\propto W^{-y}\;,\\
\rho_\alpha( |E|=2, W) &\propto W^{-y}
\end{eqnarray}
\end{subequations}
with $y = 2/3$ for $\alpha=1$ and $y=2/(4-\alpha)$ for $\alpha \le 1$
\cite{DerG84, RusHW98}. Most interestingly, within the so called
white-noise model (WNM) scaling expressions identical to Eqs.\
\eqref{eq:BEScaling} are obtained where $\alpha$ is replaced by the
dimension $d$ of the system. We would like to emphasize that the WNM
does not consider long-range correlated disorder, but the similarity
suggests that the influence of power-law correlations at the band edge
may possibly be interpreted in terms of an {\it effectively reduced
dimensionality}.
\subsection{Numerical Methods}\label{sec:Methods}
In order to generate random numbers with power-law correlations
\eqref{eq:PLCorrFunc} we use the
modified Fourier filtering method ({FFM}) \cite{MakHSS96} with one
additional step. Hereby, the correlation function is given by
\begin{equation}\label{eq:MakseCorrFun}
C(\ell) = \left(1+\ell^2\right)^{-\alpha/2} \propto |\ell|^{-\alpha} \quad (\ell\gg 1)\;,
\end{equation}
which avoids the singularity at $\ell=0$ and therefore improves the
filtering. For large distances $\ell\gg 1$ this form resembles the
desired power-law behavior \eqref{eq:PLCorrFunc}. An additional
benefit of Eq.\ \eqref{eq:MakseCorrFun} consists in avoiding aliasing
effects, which may obscure the long-range character of the
correlations. The spectral density $S(q)$ can be calculated
analytically and is given in terms of a modified Bessel function
\cite{MakHSS96}.
Overall, the special choice of the correlation function
\eqref{eq:MakseCorrFun} leads to a modified behavior with respect to
the correlation exponent $\alpha$. While for the pure power-law the
case of uncorrelated disorder is obtained for $\alpha = 1$ (see also
Eq.\ \eqref{eq:SpecDens}), the modified FFM generates uncorrelated
random numbers only in the limit $\alpha\to\infty$. Therefore, it is
expected to find an influence on the localization behavior even for
$\alpha>1$.
Additionally, we shift and normalize the obtained sequence of
correlated random numbers such that the mean vanishes and the variance
is $W^2/12$ \cite{Kay07}. Thereby, only the strength of the disorder
is adjusted while the correlations of the potential are preserved.
Then we calculate the localization length using a standard
transfer-matrix method (TMM) \cite{KraM93} with a new seed of the
random number generator for each parameter combination ($E, W,
\alpha$). The DOS is calculated by diagonalizing the Hamiltonian
\eqref{eq:AndersonHam} and counting the eigenvalues in finite energy
bins. In each case $100$ realizations are taken into account for
averaging.
Due to the symmetry of the Hamiltionian (\ref{eq:AndersonHam}) and the
chosen symmetric disorder distribution $P(\varepsilon)$ the results
should depend on the absolute value of $E$ only. This is confirmed by
our calculations. Therefore we use the arithmetic average of values
for $-E$ and $+E$ to improve the accuracy of our results.
In the following, we consider chains of fixed length $L=2^{19}$ for
the TMM calculations and $L=2^{13}$ for the DOS computation. The
correlation exponents are $\alpha = \infty, 2.0, 1.0$ and $0.5$.
\section{Results and Discussion}\label{sec:results}
In Fig.\ \ref{fig:1DEngLocW05} we show the energy dependence of
$\Lambda$ and the DOS for fixed disorder strength $W=0.5$ and
different correlation strengths $\alpha$. For energies inside the
unperturbed band ($|E| < 2$) the localization length is increased by
long-range correlations. In this region one observes an overall good
agreement with the weak-disorder result given by Eq.\
\eqref{eq:1DWeakDisExpansion}. On the other hand, near the band edge
($|E| \approx 2$) $\Lambda$ is decreased by correlations, which will
be discussed in more detail below.
\begin{figure}[bt]
\center
\includegraphics[width=\columnwidth]{Fig1_En-Loc-DOS-W05_2.pdf}
\caption{(Color online) Localization length
$\Lambda_\alpha$ vs energy $E$ at $W=0.5$ for a 1D chain of
length $L=2^{19}$ obtained from TMM calculations (a). Full
lines correspond to $\Lambda_\alpha$ given by Eq.\
\eqref{eq:1DWeakDisExpansion}. DOS $\rho_\alpha$ for a 1D
chain of length $L=2^{13}$ (b). The dashed line denotes the exact
DOS for a 1D system without any disorder, $\rho = 1/(\pi
\sqrt{4-E^2})$. Different symbols/colors denote various
correlation parameters $\alpha$.}
\label{fig:1DEngLocW05}
\end{figure}
\begin{figure}[bt]
\center
\includegraphics[width=\columnwidth]{Fig2_EnDOS-W05-Difference-Histo_2.pdf}
\caption{(Color online) Difference of the DOS $\rho_\alpha$
for a 1D chain of length $L=2^{13}$ and $W=0.5$ compared to
the uncorrelated case. Different symbols/colors denote
various correlation parameters $\alpha$ as given in the
legend.}
\label{fig:1DDOS}
\end{figure}
Another important quantity for transport studies in general is the DOS
$\rho(E)$. For an ordered system the DOS can be calculated
analytically. In this case it develops van Hove singularities at the
band edges $|E|=2$, which is typical for 1D systems \cite{Eco06} and
indicates the presence of long-range order \cite{KraM93}. In case of
uncorrelated on-site energies the van Hove singularities are smeared
out while maintaining peaks at $|E|=2$. Inside the unperturbed band
the DOS is only weakly changed compared to the DOS of the ordered
system. For $|E|>2$ there is an exponential band tail \cite{KraM93}.
For correlated on-site energies Fig.\ \ref{fig:1DDOS} illustrates that
the van Hove peak is even less pronounced and shifted towards the band
center, which results in a decreased DOS at $|E|=2$ with decreasing
$\alpha$. On the other hand, in the band tails the DOS is getting
larger for decreasing $\alpha$.
In the following we discuss the two regions separately and in more
detail. Firstly, we focus specifically on the behavior in the center
of the band. Figure \ref{fig:1DDisLocE0} shows the disorder strength
dependence of the localization length at $E=0$. One can see that
introducing correlations leads to a systematically larger localization
length. Only for very strong disorder the influence of the
correlations vanishes and the localization length for finite $\alpha$
approaches the uncorrelated localization length. Also shown is the
localization length obtained from Eq.\ \eqref{eq:1DWeakDisExpansion}
using the same parameters as in the TMM calculation. It is important
to notice, that there is no fitting procedure involved at this point.
Therefore, Eq.\ \eqref{eq:1DWeakDisExpansion} allows for an {\it
independent} check of the numerics and in particular of the correct
behavior of the generated correlations. Figure \ref{fig:1DDisLocE0}
shows a good agreement of $\Lambda$ and the weak disorder expansion
result \eqref{eq:1DWeakDisExpansion} for $W < 1$.
\begin{figure}[tb]
\center
\includegraphics[width=\columnwidth]{Fig3_DisLoc-L524288-E0.pdf}
\caption{(Color online) Localization length $\Lambda_\alpha$
vs disorder strength $W$ at the band center $E=0$ for a 1D
chain of length $L=2^{19}$ obtained from TMM calculations.
Different symbols denote various correlation parameters
$\alpha$. Dashed lines show $\Lambda_\alpha$ given by Eq.\
\eqref{eq:1DWeakDisExpansion}.}
\label{fig:1DDisLocE0}
\end{figure}
In contrast to $\Lambda$ the DOS in the band center has no particular
dependence on the correlation strength for weak disorder as
demonstrated in Fig. \ref{fig:1DDOS}. Therefore the influence of the
long-ranged correlations can be interpreted in accordance with Eq.\
\eqref{eq:1DWeakDisExpansion} as an effective disorder of the
potential. Hereby, the effective disorder strength becomes smaller
with decreasing $\alpha$.
A qualitatively different behavior of the localization length is
observed for energies close to the unperturbed band edge ($|E|=2$).
This is shown in Fig.\ \ref{fig:1DDisLocE2}. Here the localization is
enhanced in case of a correlated potential and $\Lambda$ is
consequently smaller compared to the uncorrelated case. However, for
weak disorder $W < 1$ one can still observe a power-law decay
according to Eq.\ \eqref{eq:1DPowerDecay}, but with an exponent
$y<2/3$. The results of least-squares fits of Eq.\
\eqref{eq:1DPowerDecay} to the numerical data are summarized in Tab.\
\ref{tab:1DDisLocFit} and also shown in Fig.\ \ref{fig:1DDisLocE2}.
The obtained exponents for the uncorrelated potential as well as for
$\alpha=2.0$ and $0.5$ agree with the values predicted by Eqs.\
\eqref{eq:1DPowerDecay}. Only for $\alpha=1.0$ we observe a distinct
deviation from the expected value $y=2/3$ which could be a consequence
of the modified FFM \cite{MakHSS96}. From Eq.\ \eqref{eq:SpecDens} one
expects the FFM to generate an uncorrelated potential for $\alpha=1$
since $S(q)$ becomes independent of $q$ in this case. However, due to
the short-range cutoff introduced in Eq.\ \eqref{eq:MakseCorrFun}, the
spectral density is given in terms of a modified Bessel function
\cite{MakHSS96}, which is independent of $q$ only in the limit
$\alpha\to\infty$. Therefore, the behavior of the localization length
and the DOS at the band edge is sensitive to the specific choice of
the short-range cutoff for sufficiently weak disorder.
On the other hand, for very strong disorder the localization lengths
for the correlated cases coincide with the uncorrelated case.
\begin{figure}[bt]
\center
\includegraphics[width=\columnwidth]{Fig4_DisLoc-L524288-E2.pdf}
\caption{(Color online) Localization length $\Lambda_\alpha$
vs disorder strength $W$ at the unperturbed band edge $E=2$
for a 1D chain of length $L=2^{19}$. Different symbols
indicate results obtained from TMM calculations. Error bars
are smaller than the symbol size. The full line shows the
exact result for uncorrelated disorder \cite{DerG84}.
Different colors denote various correlation parameters
$\alpha$. Dashed lines show results of a least squares fit
of $\Lambda_\alpha \propto W^{-y}$ to the data. }
\label{fig:1DDisLocE2}
\end{figure}
Figure \ref{fig:1DDisDOSE2} shows the DOS at the band edge for small
disorder strengths. Similar to the localization length one observes a
power-law decay of the DOS, which is also in accordance with the
scaling law \eqref{eq:BEScaling}. The fitted exponents are listed in
Tab.\ \ref{tab:1DDisLocFit}. Their values agree with the results
obtained from the localization length but are less accurate.
\begin{table}[pb!]
\caption{Results of a least squares fit of $\Lambda_\alpha \propto W^{-y_\Lambda}$ and
$\rho_\alpha \propto W^{-y_\rho}$ to the numerical data for weak disorder. For
$\Lambda_\alpha$ and $\rho_\alpha$ the range $W\in [0.1, 0.5]$ and $W\in [0.4,1.0]$ was used, respectively.
The obtained exponents
are compared to the expected values $y$ according to Eqs.\ \eqref{eq:1DPowerDecay}.}
\begin{tabular}{lcccc}
\hline
\hline
$\alpha$ & $\infty$ & $2$ & $1$ & $0.5$ \\
\hline
$y$ & $2/3\approx0.67$ & $2/3\approx0.67$ & $2/3\approx0.67$ & $4/7\approx0.57$\\
\hline
$y_\Lambda$ & $0.68\pm0.01$ & $0.66\pm0.01$ & $0.60\pm0.02$ & $0.55\pm0.01$ \\
$y_\rho$ & $0.65\pm0.01$ & $0.61\pm0.03$ & $0.56\pm0.02$ & $0.54\pm0.03$\\
\hline
\hline
\end{tabular}
\label{tab:1DDisLocFit}
\end{table}
\begin{figure}[tb]
\center
\includegraphics[width=\columnwidth]{Fig5_DisDOS-E2.pdf}
\caption{(Color online) DOS $\rho_\alpha$ vs disorder strength
$W$ at the unperturbed band edge $E=2$ for a 1D chain of
length $L=2^{13}$. Different symbols indicate various
correlation parameters $\alpha$. The full line shows the
exact result for uncorrelated disorder \cite{DerG84}.
Dashed lines denote results of a least squares fit of
$\rho_\alpha \propto W^{-y}$ to the data.}
\label{fig:1DDisDOSE2}
\end{figure}
\section{Summary and Conclusions}
\label{sec:summary}
In summary, we have numerically investigated the role of long-range
correlated disorder on Anderson localization in 1D systems.
Specifically, we have studied the behavior of the localization length
and the DOS as functions of energy and disorder strength for different
correlation exponents. Hereby, we found two regions in the electronic
band, where universal behavior can be observed. In the vicinity of the
band center and for weak disorder the localization length is
determined by an effective disorder strength. The localization length
remains proportional to $W^{-2}$ in this regime. The explicit
expression, which has been derived in Ref.\ \cite{IzrK99}, is given by
Eq.\ \eqref{eq:1DWeakDisExpansion}.
A qualitatively different behavior is found at the band edge. Here,
the localization length and DOS are given by Eqs.\
\eqref{eq:BEScaling}, which implies a power-law behavior for both
quantities as functions of disorder strength. A comparison with the
WNM suggests that in this region the long-range correlations lead to
an effectively reduced dimension.
Overall, we find good agreement of our numerical results with the
respective analytical expressions. However, we also find that the
particular realization of the long-range correlations plays an
important role. In the present case the FFM \cite{MakHSS96} makes use
of a modified correlation function given by Eq.\
\eqref{eq:MakseCorrFun}, which contains the desired long-range but
also short-range contributions. Our numerical results indicate that
the modification leads to observing an unexpected influence of the
correlations even for correlation exponents $\alpha>1$. This behavior
should be taken into account for investigations involving realistic
disorder potentials with long-range correlations.
|
\section{Introduction}
\label{SX:INTR}
\noindent
The past few years have witnessed a growing interest in the design and application of coarse-graining methods to simulate complex fluids.\cite{complex} This effort has been motivated by the need for improving computational efficiency with the purpose of investigating complex systems on the numerous lengthscales on which their properties develop.\cite{Mueller-Plathe,DePablo,Klein,Klein1,Voth} Computer simulations have the capability of providing detailed microscopic information on the static and dynamics of the systems under study,\cite{Tisley} but they are limited in the range of timescales and in the number of molecules that can be simulated because the precision of the calculations degrades with the number of computer iterations with a behavior that depends on the Lyupanov exponent of the system. Once the number of particles is set, the window of achievable timescales that can be investigated becomes defined.\cite{Frenkel,shadowing} Because the maximum number of iterations decreases with increasing number of simulated particles, it is particularly difficult to simulate systems where characteristic lengthscales are diverging, such as a system approaching a second order phase transition.\cite{binderbook,blendjmg}
Recent improvements of computational machines has lead to a considerable extension of the maximum time- and length scales that can be reached by simulations where the system is described at the atomistic level. However, for many complex systems, including liquids of high-molecular weight macromolecules, the computational power is still inadequate to describe, at the atomistic level, the long-time dynamics. For example, the most recent and advanced simulations of long chains that have an extended number of entanglements adopt a simplified model, which treats the structure of the polymer as a collection of beads and springs interacting through a FENE potential. This model allows for the simulations of a large number of polymers, which is important for the proper calculation of viscoelastic properties, and reaches full relaxation for all but the longest chains simulated.\cite{Grest}
Progress has been made when the focus is on qualitative behavior and scaling exponents.\cite{DPD,binderbook} For example, if the complex intra- and inter-molecular non-bonded interactions are simplified into an identical potential, the computational efficiency improves dramatically as the code does not need to identify and treat uniquely different pairs of interacting sites. This strategy, however, has the disadvantage that the thermodynamics of the system is not properly described because the interactions are too drastically simplified.
The need for methods that are fully predictive of the physical properties of a system on the basis of the specific chemical structure of the sample and its thermodynamic conditions has stimulated new interest in developing fast quantitative simulations. Such predictive approaches are useful, for example, to evaluate \textit{a priori} the structure and dynamics of newly synthesized polymeric materials, in relation to their technological applications. Following this perspective, several procedures have been proposed to speed up atomistic simulations, while conserving their power of predicting quantitative properties.\cite{Mavran,Mavran1} A few simulations of long entangled chains have been performed using united atoms (UA).\cite{Mavran,Putz,Tsolou,Harm}. For UA the effective unit is very close to the atom is size, i.e. $CH_x$ with $x=1,2,3$, which allows for some gain in the computational time.
A useful strategy to improve the outcome of simulations on the long timescale and large lengthscale is the use of coarse-graining procedures.\cite{MSMD, Mueller-Plathe} A coarse-graining procedure averages out irrelevant degrees of freedom, which occur on lengthscales smaller than a designated cutoff length, and this allows for the extension towards large scales of the simulations. Another way to put it is that, because the interaction potentials become softer, the maximum time and lengthscale increase as the basic timestep of the mesoscale (MS) simulations becomes larger. The characteristic lengthscale of coarse-graining has to be defined on the basis of the properties that need to be investigated. In this paper we discuss a first-principles way of selecting meaningful lengthscales for the structural and dynamical coarse-graining.
Several considerations need to be made to properly develop a coarse-graining procedure. As the coarse-grained liquid is represented as a function of new coordinates, an effective potential needs to be derived to be used as an input to the MS simulation. Care has to be taken to make the potential reproduce the structure of the system, namely pair distribution functions, and to be thermodynamically consistent. A common procedure to optimize the coarse-grained description is to use self-consistent numerical methods that are optimized to reproduce atomistic descriptions through iterative procedures. Usually the target is the optimization of specific quantities, such as the pair distribution function,\cite{pdf} the forces generated by the soft potential,\cite{Voth} or directly the thermodynamic properties.\cite{kroger}
We recently proposed an approach that starts from the Ornstein-Zernike (OZ) equation where the atomistic sites are defined as real sites, and the coarse-grained sites are defined as auxiliary sites.\cite{YAPRL,melt,archi,blend,block,anthony,jaypaper} Because our procedure is analytical, and no optimization of parameters is needed in our approach as the potential is explicitly dependent on the thermodynamic and molecular parameters, it opens up the possibility of deriving a formal solution to key problems. For example, it is straightforward to show that the structural properties are consistent between the two levels of description, i.e. atomistic and coarse-grained.\cite{YAPRL} Moreover, the thermodynamic properties of the coarse-grained polymer liquid (e.g. isothermal compressibility,\cite{YAPRL} pressure in the virial and in the compressibility routes, total and cohesive energy) are shown to be formally consistent in the two levels of coarse-graining.\cite{mccarty} Local structure is easily included \textit{a posteriori} through a multiscale modeling procedure.\cite{jaypaper,multis} Finally, it is possible to derive an analytical rescaling factor for the dynamics, which is the main focus of this paper.
While the structure is well described by simulations of the coarse-grained system on the scale larger than the scale of coarse-graining, the dynamics in MS simulations is unrealistically fast. Because local degrees of freedom are averaged out, the coarse-grained molecules move rapidly over a simplified free energy landscape. As the system explores efficiently this ``reduced" configurational landscape, the measured dynamics is artificially sped up by the smoothness of the potential. This is useful when coarse-grained representations are used to rapidly reach an equilibrated state of the system before starting the atomistic molecular dynamics (MD) simulation. However, to directly collect information on the dynamics of systems from MS MD simulations, it is necessary to develop formalisms that rescale the unrealistically fast dynamics into the slower dynamics at atomistic resolution. In this paper we discuss in details an analytical procedure we recently proposed to rescale the mesoscale dynamics. The procedure is able to predict center-of-mass dynamics in quantitative agreement with experiments and atomistic simulations.\cite{shortivan}
The common strategy to rescale the dynamics is to build a ``calibration curve". The latter is obtained through the numerical fitting of dynamical quantities and optimization of the related parameters until the agreement of dynamical properties calculated in an all-atom and in a MS simulations is obtained.\cite{Nielsen,Kremer} However, the numerical calculation of optimized calibration curves for the dynamics is quite difficult to achieve for macromolecular systems, as the dynamics is mode dependent: there are in principle $N$ internal modes in any molecule formed by $N$ units and the degree of polymerization of a long chain can be of the order of one million monomers. Moreover, numerically optimized parametric quantities are in general not transferable between systems in different thermodynamic conditions or with different chemical structure or increasing degree of polymerization. To overcome this problem, it is common to select as coarse-grained units ones that are very close in size to the atomistic units, so that the needed corrections to reach consistency in dynamic properties are minimal. In this case, corrections to the measured dynamics can be evaluated through a perturbative formalism which should rapidly converge to the desired value. The resulting computational gain is, however, still limited. Recently a numerical Ornstein-Zernike-based approach, with atomistic-level coarse-graining and the Percus -Yevick closure approximation, has been proposed, which shows different rescaling factors depending on the time correlation function under study.\cite{Wang} Another coarse-grained approach for polyethylene melts describes a polymer chain as a collection of soft blobs connected by elastic bands, which enforce chain-chain uncrossability. Simulations follow an effective Langevin equation, whose parameters, i.e. effective potential, frictions and random forces, are obtained from an atomistic MD simulation. The optimized equation of motion (eom) reproduces well experimental data.\cite{PB}
Our approach is different from others in several ways. First of all it is analytical rather than numerical, providing the formal rescaling factor by solving the eoms in the two levels of representation. In this way, there is no need of performing an atomistic simulation to input numerical quantities in our formalism. The dynamics measured in MS simulations of coarse-grained systems is directly rescaled into its atomistic counterpart using approximate closed-form expressions of friction and entropy. The two levels of coarse-graining, which allow for a straightforward analytical solution, are here two simple isotropic models: a soft sphere description for the molecular coarse-graining, and a bead-and-spring description for the monomer level coarse-graining. More sophisticated coarse-grained models can be developed for intermediate lengthscales\cite{block,anthony}; however, the formalism can become more involved.\cite{Akkerman} At the atomistic level the polymer is described as a "bead-and-spring" type of approach where the chain is a collection of friction points connected by harmonic springs. This is an implementation of the most popular model to treat unentangled polymer melt dynamics, i.e. the Rouse model, and maps well into the dynamics of polymers described not only by UA simulations, but also by atomistic simulations and experiments, as it contains both local chemical structure, semiflexibility, and finite size effects.\cite{jpcmreview,blumen,Richter} It is a very accurate and molecular specific model, which has been shown to describe well, for example, the dynamics of the protein CheY, by testing its predictions of NMR relaxation against experiments.\cite{jpcmreview,Guenza}
In our coarse-grained model a polymer chain is represented as a soft-colloidal particle.\cite{Hall,Kremer1,Likos,YAPRL} Because the lengthscale of the coarse-graining is of the order of the molecular radius of gyration, i.e. the size of the molecule, the \textit{direct} predictions of the rescaling procedure are suitable for properties on lengthscales larger than $R_g$ and on timescales longer than the longest time of intramolecular relaxation, i.e. the longest correlation time in the Rouse theory. Internal dynamics cannot be obtained directly from the coarse-grained simulation; however, the rescaled diffusion coefficient leads to the monomer friction coefficient, which can be used as an input to well-tested theories of polymer dynamics, and \textit{indirectly} recovering the dynamics in the complete spectrum of polymer relaxation. An example of this kind of calculation is presented in this paper in Section \ref{SX:rft}. The extended level of coarse-graining provides a good test of our procedure, as large deviations could result from the rescaling if the method were not correct. Furthermore, our procedure can be useful in the study of long-time relaxation, given that large length scales and long timescales are most difficult to simulate for polymeric systems.
Although the outline of our rescaling theory has been published recently in a short paper,\cite{shortivan} this paper presents a detailed derivation and discussion of our approach, which includes the prediction of the dynamics for new samples. After introducing our coarse-grained model input to the mesoscale simulations, we formally derive the rescaling approach for the dynamics, starting from the Liouville equation and using projection operators. Friction coefficients in the two descriptions are derived from the solution of the memory functions, while the rescaling of the simulation time is obtained from the entropic contribution, which accounts for the intramolecular degrees of freedom neglected in the soft colloid representation. Theoretical predictions compare well against UA MD simulations,\cite{MONDE,JARAM,HEINE,Mavran} and experiments,\cite{Richter,maybeothers,maybeothers1,maybeothers2,maybeothers3}. We also calculate the diffusion coefficient for new PE samples in thermodynamics conditions for which UA-MD data are not available. The purpose of these calculations is to show that our method is not a simple rescaling of the mesoscale data through a shift of the diffusion coefficient to bring dynamical results to coincide with atomistic simulations, as is conventionally done. Instead our approach is fully predictive and can be used to calculated the diffusion coefficient, and the monomer friction, for new samples.
The paper is structured as follows: after introducing our coarse-grained model in Section \ref{SX:cg} and the projection operator technique to derive the equations of motion in the two levels of coarse graining in Section \ref{SX:dc}, we formally derive the rescaling approach for the dynamics from the solution of the memory functions in Section \ref{SX:ar}. We then present the MS simulations (Section \ref{SX:ms}), as well as the results obtained from the same, and apply the rescaling procedure to the data from MS simulations (Section \ref{SX:ao}). Predictions of dynamical quantities and direct comparison for several samples, both from atomistic simulations and from experiments, provide a stringent test of the approach and show good quantitative agreement. A brief discussion in Section \ref{SX:da} concludes the paper.
\section{Coarse-graining of polymeric liquids: structural properties}
\label{SX:cg}
In this section we briefly review the theoretical background of the pair distribution functions that are input to our rescaling equation. The structure of a polymeric liquid, at lengthscales equal or larger than the monomer size, is fully specified by the momomer total distribution function, $h(r)$, which for polymer melts depends on two characteristic lengthscales, namely the density fluctuation lenthscale, which is the atomic length, and De Gennes' correlation hole lengthscale.\cite{GENNES} The latter is of the order of the molecular radius-of-gyration, $R_g=\sqrt{N/6}l $, which is the overall dimension of the polymer, where $N$ is the degree of polymerization and $l$ is the statistical bond length. We select $l$ and $R_g$ because these are the two lengthscales that define the structural properties of the polymeric liquid.
At the monomer level traditional dynamical approaches, such as the Rouse model and semiflexible models, adopt a bead-and-spring representation where each monomer can be modeled as a friction point connected by springs (see Fig.~\ref{models}). A similar model, where the polymeric chain is described as a collection of ``sites" centered at the center of the monomeric unit, is also in conventional theories of polymer liquids.\cite{PRISM,PRISM1} Although ``site" is the word most used in the liquid state community and ``monomer" or ``bead" is the common wording in the literature on polymer dynamics, in this paper they identify the same $CH_x$ unit and henceforth they will be used interchangeably. It is important to notice that all the $CH_2$ units are assumed to be equivalent and independent of the position along the chain.
The coarse-graining of a polymer at the $R_g$ lengthscale represents each molecule as an interacting soft colloidal particle with symmetric or asymmetric shape.\cite{Hall,Kremer1,Likos} In our model,\cite{YAPRL,melt,archi,blend} the macromolecular liquid is represented as a liquid of symmetric soft colloidal particles interacting through a pair potential. This potential has a range of the order of few $R_g$, and each soft-colloidal particle is centered at the center-of-mass of a polymer (see Fig.~\ref{models}).
\begin{figure}
\centering
\includegraphics[scale=0.5]{FIGURE1.pdf}
\caption[]{Illustration of monomer and overall coarse-graining of a homopolymer linear chain.}
\label{models}
\end{figure}
The coarse-graining procedure that translates the monomer description into the solf-colloidal representation is performed starting from an Ornstein-Zernike equation where monomers are assumed to be real sites, while the center-of-mass (cm) are auxiliary sites.\cite{Hansen} The cm-cm total intermolecular correlation function is expressed as a function of the polymer parameters as \cite{YAPRL,melt}
\begin{eqnarray}
\label{EQ:1}
h^{cc}(r) & = & \frac{3}{4} \sqrt{\frac{3}{\pi}} \frac{\xi'_\rho}{R_g} \left( 1-\frac{ \xi^2_c}{\xi^2_\rho} \right) e^{-3r^2/(4R^2_g)} \nonumber \\
&& -\frac{1}{2} \frac{\xi'_\rho}{r} \left( 1-\frac{\xi^2_c}{\xi^2_\rho} \right)^2 e^{R^2_g/(3\xi^2_\rho)} \Biggl[ e^{r/\xi_\rho} \mbox{erfc} \Biggl( \frac{R_g}{ \sqrt{3} \xi_\rho} + \nonumber \\
&& \frac{ \sqrt{3}r}{2R_g} \Biggr) -e^{-r/\xi_\rho} \mbox{erfc} \left( \frac{R_g}{ \sqrt{3} \xi_\rho}- \frac{\sqrt{3}r}{2R_g} \right)\Biggr] \ ,
\end{eqnarray}
where erfc$(x)$ is the complementary error function. Here $\xi'_\rho=R_g/(2\pi\rho^*_{ch})=3/(\pi \rho l^2)$ with $\rho^*_{ch}\equiv\rho_{ch}R^3_g$ being the reduced molecular number density, $\rho_{ch}=\rho /N$ is the molecular density, $\rho$ the site number density, and $l$ is the statistical segment length. The length scale of density fluctuations, $\xi_\rho$, is defined as $\xi^{-1}_\rho=\xi^{-1}_c +\xi'^{-1}_\rho$, and $\xi_c=R_g/\sqrt{2}$ is the length scale of the correlation hole.\cite{GENNES}
In the limit of long chains, $N\rightarrow \infty$, Eq.(\ref{EQ:1}) reduces to
\begin{equation}
h^{cc}(r)\approx-\frac{39}{16}\sqrt{\frac{3}{\pi}}\frac{\xi_\rho}{R_g}\left(1+\sqrt{2}\frac{\xi_\rho}{R_g}\right)
\left(1 - \frac{9 r^2}{26 R_g^2} \right) e^{-\frac{3 r^2}{4 R_g^2}} \, .
\label{EQ:2}
\end{equation}
For polymers with $N \geq 30$, a plot of $h(r)$ shows that the two equations, Eqs.(\ref{EQ:1}) and (\ref{EQ:2}), are indistinguishable.\cite{YAPRL,melt}
The structure of the liquid on the lengthscale of the polymer radius-of-gyration and larger, as represented by $h^{cc}(r)$, is in quantitative agreement with the output of both the atomistic UA MD and the MS MD simulation of the coarse-grained liquid. The theory recovers identical analytical expressions of the compressibility in the atomistic and the coarse-grained representations, indicating thermodynamic consistency between the two levels of description.\cite{YAPRL,melt}
Eqs.(\ref{EQ:1}) and (\ref{EQ:2}) are \textit{de facto} coarse-graining equations, which translate the atomistic description of a polymer liquid, onto its representation as a liquid of interacting soft colloidal particles of size $R_g$. The advantage of our coarse-graining approach is that it is analytical and general as it applies to systems with different thermodynamic conditions, different degree of polymerization and different bond length.\cite{YAPRL,melt,archi,blend,block,anthony,jaypaper,mccarty,multis}
\section{Dynamical coarse-graining: from the Liouville to the Langevin equations}
\label{SX:dc}
While the structure of the polymeric liquid, as represented by the total correlation function, is identical in the atomistic and coarse-grained descriptions,\cite{YAPRL,melt} the dynamics of the coarse-grained system, as measured in the MS MD simulations of the soft-colloidal particles, is unrealistically accelerated. In Fig.~(\ref{figure2}) we show, for a polyethylene chain with $N=44$, the mean-square-displacement of the center-of-mass obtained in MS MD simulations of the polymer liquid represented as soft-colloidal particles and the mean-square-displacement directly measured in UA MD simulations . The dynamics in the coarse-grained representation is several orders of magnitude faster than the atomistic description. Because the level of coarse-graining of the model presented here is extended, this effect is more evident than in other models; however, accelerated dynamics is present in any simulation of coarse-grained systems.
It has been argued that there are two main effects of coarse-graining that accelerate the dynamics: namely, the change in entropy and the change in the friction coefficient. \"Ottinger has presented an approach for systems far from equilibrium that accounts for those effects.\cite{Ottinger} We propose here a procedure, based on first-principles theory to properly account for both contributions through the introduction of the necessary corrections for systems where the fluctuation dissipation theorem applies, e.g. close to equilibrium.
\begin{figure}[]
\begin{center}
\centering
\scalebox{0.35}{\includegraphics{PE44_both_rescales.pdf}}
\caption[Plot of the mean-square displacement for polyethylene melt 44]{Cm mean-square displacement, for a polyethylene melt with $N=44$, from MS MD simulations (triangles) and UA MD simulations (squares). Dashed lines show the diffusive limits of the two samples.}
\end{center}
\label{figure2}
\end{figure}
To coarse-grain the dynamics of the polymeric liquid on the lengthscale of the radius of gyration, we adopt a Mori-Zwanzig projection operator technique, where the selected slow variables are the position and momentum coordinates of the polymer center-of-mass. This description should represent well center-of-mass diffusion.\cite{SMPLQ,Ackasu,Opp}
The atomistic level representation is obtained following the same Mori-Zwanzing procedure, but choosing as the slow relevant variables the ensemble of position and momentum coordinates of the center of mass of the monomeric unit, which for a polyolefin is the $CH_x$ unit, with $x=1,2,$ or $3$. This model is consistent with the representation of the polyethylene chain in UA MD simulations,\cite{manychains,schweizer} and it has been shown to describe at a high level of accuracy the dynamics of polyolefins at the monomer lenghtscale.\cite{jpcmreview,Marina,marinamacrom}
In the long-time regime the two descriptions, soft-colloid and monomeric/UA, should be identical as they both recover the diffusive dynamics of the center-of-mass.\cite{DoiEdw} In fact, they are not, as the soft-colloidal description is heavily coarse-grained and its dynamics is accelerated. The analytical rescaling factor is derived directly from the comparison between the soft-colloid and the monomer dynamical equations.
As this coarse-graining and rescaling procedure is general, it can be adopted to formalize the dynamics of the molecular liquid at the desired level of coarse-graining. However, the projection operator technique rests on a separation of timescales between the slow relevant variables onto which the dynamics is projected, and the fast irrelevant variables that are averaged out. If no separation of timescales is observed, it is necessary to include corrections to the projected dynamics, which appear as contributions to the friction coefficient, expressed as memory functions. In the system investigated here, polymer melt dynamics, no clear separation of timescales occurs between the dynamics of the "tagged" chain and the dynamics of the surrounding molecules.\cite{manychains} For this reason, the Generalized Langevin Equation generated from this procedure needs to account for the correction terms to the projected dynamics, which are represented by the memory function contributions.\cite{Zwanzing}
For a liquid of $n$ macromolecules containing $N$ monomers, the first-principle Liouville equation is simply written as
\begin{eqnarray}
\frac{\partial f \left( \mathbf{R, P}, t \right)}{\partial t} = iLf \left( \mathbf{R, P}, t \right) \ ,
\label{EQ:Liouville}
\end{eqnarray}
with
\begin{eqnarray}
f \left( \mathbf{R, P}, t \right)= \prod_{j=1}^{n} \left[ \prod_{a=1}^{N} \delta (\mathbf{r}_a^j(t) - \mathbf{R}_a^j) \delta (\mathbf{p}_a^j(t) - \mathbf{P}_a^j) \right] \ ,
\end{eqnarray}
the instantaneous distribution in reduced phase space, and $\mathbf{R}_a^i$ and $\mathbf{P}_a^i$ are the phase-space variables associated with the Cartesian position and momentum coordinates of the bead $a$ belonging to molecule $i$, namely $\mathbf{r}_a^i(t)$ and $\mathbf{p}_a^i(t)$.
The formal solution of Eq.(\ref{EQ:Liouville}) is
\begin{eqnarray}
f \left( \mathbf{R, P}, t \right)= e^{-iLt} f \left( \mathbf{R, P} \right) \ ,
\end{eqnarray}
with the shorthand notation $ f \left( \mathbf{R, P} \right) = f \left( \mathbf{R, P}, 0 \right) $
The Liouville operator is defined as
\begin{eqnarray}
iL=- \sum_{j=1}^n \sum_{a=1}^N \left[\frac{\partial U_j}{\partial \textbf{r}_a^j} \cdot \frac{\partial}{\partial \textbf{p}_a^j} - \frac{\textbf{p}_a^j}{m}\cdot \frac{\partial}{\partial \textbf{r}_a^j} \right] \ ,
\end{eqnarray}
where the total energy $U_j$ in the Hamiltonian, $H$, contains both intramolecular, $U^0_j$, and intermolecular, $W_{ij}$, pairwise decomposable potential contributions. The intermolecular potential contains both interactions between the $n$ tagged chains, $W^0_{jk}$, and between the tagged chains and the surrounding ones, $W_{jk}$, so that the usual condition applies that $L_0f(\mathbf{R,P})=0$.
The statistical average of the phase space density is defined as
\begin{eqnarray}
\Bigl< f \left( \textbf{R},\textbf{P} \right)\Bigr> = \int d \textbf{r} \int d \textbf{p} f \left( \textbf{R},\textbf{P} \right) \psi \left( \textbf{r},\textbf{p} \right) \ ,
\end{eqnarray}
with the equilibrium distribution of particle positions and coordinates
\begin{eqnarray}
\psi \left( \textbf{r},\textbf{p} \right) = e^{-\beta H} \left[ \int d \textbf{r} \int d \textbf{p} e^{-\beta H} \right]^{-1} \ ,
\end{eqnarray}
where $\beta=(k_B T)^{-1}$, $k_B$ the Boltzman's constant, and $T$ the absolute temperature.
Following Mori-Zwanzig, we define the projection operator, $\hat{P}$, for the coarse-grained model we adopt, namely the monomer and the soft-colloidal.
\subsection{Monomer level representation of the polymer chain}
\label{SX:ml}
In our atomic-level description each macromolecule is represented as a collection of connected beads, or friction points.
In the field variables for one molecule ($n=1$),
\begin{eqnarray}
g \left( \mathbf{R, P}, t \right)= \left[ \prod_{a=1}^{N} \delta \left( \mathbf{r}_a(t) - \mathbf{R}_a \right) \delta \left(\mathbf{p}_a(t) - \mathbf{P}_a \right) \right] \ ,
\end{eqnarray}
the projection operator is defined as
\begin{eqnarray}
\hat{P}h \left( \textbf{R},\textbf{P},t \right) = \int d \textbf{R}' \int d \textbf{P}' \int d \textbf{R}'' \int d \textbf{P}'' \nonumber \\
\bigl<h \left( \textbf{R},\textbf{P},t \right) g \left( \textbf{R}',\textbf{P}' \right)\bigr> \nonumber \\
\times\bigl<g \left( \textbf{R}',\textbf{P}' \right) g \left( \textbf{R}'',\textbf{P}'' \right)\bigr>^{-1} g \left( \textbf{R}'',\textbf{P}'' \right) \ ,
\end{eqnarray}
where $\hat{P}=(\hat{P})^2$ and $\hat{P} g \left( \textbf{R},\textbf{P} \right)=g \left( \textbf{R},\textbf{P} \right) $. Here we use for the field variable the symbol $g \left( \mathbf{R, P}, t \right)$ to indicate that the slow variables in the projection operator can be different than the ones in the general formalism of the preceding section. By applying the projection operator to both the left and the right sides of the Liouville equation, one recovers a generalized Langevin equation.\cite{Ackasu,Opp,manychains,schweizer}
Briefly, the generalized Langevin equation in the phase space is then transformed into its analog equation in space coordinates, yielding
\begin{eqnarray}
\label{inertiala}
m\frac{d^2 \textbf{r}_a(t)}{dt^2} & = &\beta^{-1} \frac{\partial}{\partial \textbf{r}_a(t)} \ln \psi(\textbf{r})\nonumber \\
&& - \int_0^t d \tau \sum_{b=1}^N \frac{\beta \textbf{p}_b}{3m} \bigl<\textbf{F}_a(t) \cdot \textbf{F}_b^{\hat{Q}} (t-\tau)\bigr> \nonumber \\
&& + \, \textbf{F}_a^{\hat{Q}}(t) \ ,
\end{eqnarray}
where $\psi(\textbf{r}(t))$ is the intramolecular distribution function.
The inertial contribution in Eq.(\ref{inertiala}) can be discarded, as the liquid has a low Reynolds number and the dynamics is overdamped.
The Generalized Langevin Equation is simply written as
\begin{eqnarray}
\zeta_m \frac{d \textbf{r}_a(t)}{dt}=\frac{1}{\beta} \frac{\partial}{\partial \textbf{r}_a(t)} \ln \psi(\textbf{r}) +
\textbf{F}_a^{\hat{Q}}(t) \ ,
\end{eqnarray}
with the averaged friction coefficient, in the Markov limit,
\begin{eqnarray}
\zeta_m \approx \beta /3 \ N^{-1} \sum_{a,b=1}^N \int_0^\infty d \tau \bigl<\textbf{F}_a(t) \cdot \textbf{F}_b^{\hat{Q}} (t-\tau)\bigr> \ .
\label{cJ}
\end{eqnarray}
This equation describes how the monomer friction coefficient is generated from the space and time correlation of the random forces that act on two different segments of the "tagged" polymer chain, $a$ and $b$. The extent of the correlation depends on the propagation of the forces through the macromolecule, its structure and local flexibility. The forces are generated by the monomers of the surrounding molecules randomly colliding with the monomers of the tagged chain: the collision strength depends on the structure of the liquid and on the interparticle potential.
A more explicit definition of the friction coefficient is given in the following sections.
\subsection{Solution of the Generalized Langevin Equation in the monomer representation}
\label{SX:so}
The intramolecular distribution function is approximated in our description by a Gaussian distribution
\begin{eqnarray}
\psi (\textbf{r})=\left[(2\pi)^N \det(\textbf{A}^{-1}) \right]^{-3/2} e^{-\frac{3}{2l^2}\textbf{r}^T\textbf{A}\textbf{r}} \ ,
\end{eqnarray}
which holds for polymer chains longer than about $30$ monomers.\cite{jpcmreview}
This leads to a Generalized Langevin Equation where the intramolecular contribution is linear in the monomer coordinates
\begin{eqnarray}
\zeta_m \frac{d \textbf{r}_a(t)}{dt} = -\frac{3k_B T}{l^2} \sum_{b=1}^N \textbf{A}_{a,b}\textbf{r}_b(t)+\textbf{F}^{\hat{Q}}_a(t) \ ,
\label{GLEmonomer}
\end{eqnarray}
and is simply solved through transformation into normal modes of motion.\cite{DoiEdw,jpcmreview}
The matrix $\textbf{A}$ is defined, for a semiflexible polymer represented as a Freely Rotating Chain (FRC), as the product of two matrices, $\textbf{M}$ and $\textbf{U}$,
\begin{eqnarray}
\textbf{A}=\textbf{M}^T \left( \begin{array} {cc} 0 & 0 \\ 0 & \textbf{U}^{-1} \end{array} \right) \textbf{M} \ ,
\label{amatrix}
\end{eqnarray}
with the connectivity matrix, with dimensions $N \times N$, defined as
\begin{eqnarray}
\textbf{M}= \left( \begin{array} {ccccc} N^{-1} & N^{-1} & N^{-1} & ... & N^{-1} \\ -1 & 1 & 0 & ... & 0 \\
0 & -1 & 1 & ... & 0 \\
... & ... & ... & ... & ... \\ 0 & ... & 0 & -1 & 1 \end{array} \right) \ ,
\end{eqnarray}
and the $\textbf{U}$ matrix defined as a function of the stiffness parameter $g$ as
\begin{eqnarray}
\textbf{U}_{ij}=\Biggl<\frac{\textbf{l}_i\cdot \textbf{l}_j}{|\textbf{l}_i| |\textbf{l}_j|}\Biggr>= g^{|j-i|} \ .
\label{semiflexibility}
\end{eqnarray}
Here, $g=-\bigl<\cos \theta\bigr>$ and $\theta$ is the angle between two consecutive bonds in the FRC representation of a homopolymer.\cite{DoiEdw}
The stiffness parameter, $g$, is specific of the chemical structure and thermodynamic conditions of the sample under study.
\subsection{Center-of-mass level representation of the polymer chain}
\label{SX:co}
In the soft colloidal particle representation the projection operator targets the center-of-mass of the polymer. The field variable ($n=1$, $N=1$, $a=cm$) is simply defined as
\begin{eqnarray}
g \left( \mathbf{R, P}, t \right)= \left[ \delta \left( \mathbf{r}_{cm}(t) - \mathbf{R} \right) \delta \left(\mathbf{p}_{cm}(t) - \mathbf{P} \right) \right] \ .
\end{eqnarray}
Applying the projection operator in the new field variable to the Liouville equation, where $U=0$ and $W_{ij}\neq 0$, leads to the generalized Langevin equation
\begin{eqnarray}
\frac{\partial}{\partial t} g \left( \textbf{R},\textbf{P},t \right) & = & - \int_0^t ds \int d \textbf{R}' \int d \textbf{P}' M(\textbf{R}, \textbf{P},\textbf{R}',\textbf{P}') \nonumber \\
&& \times g \left( \textbf{R}',\textbf{P}',(t-s) \right) + F \left( \textbf{R},\textbf{P},t \right) \ ,
\end{eqnarray}
which reduces, following the procedure briefly outlined in Section \ref{SX:ml}, to
\begin{eqnarray}
\label{inertial}
m\frac{d^2 \textbf{r}_{cm}(t)}{dt^2} &=& -\int_0^t d \tau \frac{\beta \textbf{p}_{cm}}{3m} \bigl<\textbf{F}_{cm}(t) \cdot \textbf{F}_{cm}^{\hat{Q}} (t-\tau)\bigr> \nonumber \\
&& +\, \textbf{F}_{cm}^{\hat{Q}}(t) \ .
\end{eqnarray}
In the overdamped regime,
\begin{eqnarray}
\zeta_{soft} \frac{d \textbf{r}_{cm}(t)}{dt} = \textbf{F}^{\hat{Q}}_{cm}(t) \ ,
\label{GLEsoft}
\end{eqnarray}
where $\zeta_{soft}$ is the friction coefficient for the colloidal particle, $\zeta_{soft}\cong \beta /3 \int_0^\infty d \tau \bigl<\textbf{F}_{cm}(t) \cdot \textbf{F}_{cm}^{\hat{Q}} (t-\tau)\bigr>$.
Eq.(\ref{GLEsoft}) obeys the fluctuation-dissipation relation $\left<\mathbf{F}_{cm}(t) \cdot \mathbf{F}_{cm}(t')\right>= \delta_{t-t'} 6 k_BT \zeta_{soft}$.
The choice of the field variables in the projection operator defines the length scale of coarse-graining and the variables in which the resulting Generalized Langevin Equation is expressed. Because the derivation just presented depends on the basic assumption that the correlation function of the bath variables are short lived in the presence of heavy particles, and correction terms represented by the memory functions are minimized when a clear separation of timescale is observed between the slow variables in the projection operator and the fast variables that are averaged out, this criteria provides a way of selecting the relevant lengthscales for the coarse-graining, when dynamical properties are under study.
For example, as far as polymer dynamics is concerned, we know that for times longer than the longest Rouse correlation time, $\tau_R\approx R_g^2/D$, polymer internal dynamics is fully relaxed and the monomer dynamics follows the motion of the center-of-mass, which is long lived. This suggests that the center-of-mass coordinates are a good choice to represent the projected slow dynamics for time $t >> \tau_R$. This reasoning holds for both unentangled and entangled polymer dynamics as the longest relaxation time, after which free diffusion and Brownian motion set in, is $\tau_R$ with the proper diffusion coefficient, i.e. for unentangled chains $D_{unent}\propto N^{-1}$ and for entangled chains $D_{ent}\propto N^{-2}$.
\section{Analytical rescaling of the coarse-grained dynamics}
\label{SX:ar}
The two Langevin equations, Eqs.(\ref{GLEmonomer}) and (\ref{GLEsoft}), display the two levels of coarse-graining of the macromolecular liquid, which are adopted in this paper. The comparison of the two equations, which in the longtime regime should predict identical dynamics for the polymer center-of-mass, shows that the two equations differ because of the presence of the intramolecular free energy in the monomer description, which is absent in the soft colloidal approximation, and because of the different friction coefficients in the two representations.
\subsection{Free energy rescaling}
\label{SX:fe}
The elimination of degrees of freedom increases the entropy of the system, as every coarse-grained state corresponds to a number of preaveraged microstates. In an extreme picture we can imagine that the preaveraging due to the coarse-graining procedure is in effect transforming the energy of the system, expressed for example in the Liouville equation by an Hamiltonian, into a free energy in the corresponding Langevin equation. While the Hamiltonian contains kinetics and potential energy, the free energy includes an entropic contribution due to the preaveraged microstates for each coarse-grained state.
As far as the free energy correction is concerned, the system described by the larger cutoff lenghtscale is the one where the level of coarse-graining is most extensive and the highest entropic correction has to be included. This correction can be calculated from the comparison of the two equations. Because the system described at the monomer level is exploring in time the intramolecular energy states of the configurational landscape, its dynamics is slowed down with respect to the colloid representation where intramolecular degrees of freedom are not present. To take this effect into account we calculate the correction that has to be included in the soft-colloid representation to take into account the time spent by the atomic system to explore the internal degrees of freedom.
Consistent with the monomer-level model adopted in our study and with UA MD simulations, the polymer is described as a collections of beads, or friction points, connected by harmonic springs. Each bead corresponds to a $CH_x$ moiety, with $x=2$ or $3$, depending if the unit is imbedded in the chain or is terminal. This model has been shown to provide a realistic representation of the dynamics of numerous polymeric systems with different chemical structure.\cite{Guenza,jpcmreview,Marina,marinamacrom,blumen}
The intramolecular potential is defined as
\begin{eqnarray}
U(r)=\frac{3k_BT}{2l^2}\sum_{i,j=1}^N \textbf{A}_{i,j} \textbf{r}_i \cdot \textbf{r}_j \ ,
\label{EQ:3}
\end{eqnarray}
with $U(r)$ not to be confused with the semiflexibility matrix of Eq.(\ref{semiflexibility}). Here $\bf{A}$ is the connectivity matrix of Eq.(\ref{amatrix}), which represents the structure and local flexibility of the polymer \cite{Yamakawa,bixonzwanzig}, $\textbf{r}_i$ the position of bead $i$ in a chain of $N$ beads or united atoms, and $\textbf{l}_i=\textbf{r}_{i+1}-\textbf{r}_i$ the bond vector connecting two adjacent beads.
The statistically averaged internal energy for one molecule consisting of $N$ monomers is given by
\begin{equation}
\left< \frac{U}{k_B T} \right> = N \int U e^{-\frac{3}{2 l^2} \mathbf{r}^T \mathbf{A} \mathbf{r}} d\, r = \frac{3N}{2l^2} \int \mathbf{r}^T \mathbf{A} \mathbf{r} e^{-\frac{3}{2l^2} \mathbf{r}^T \mathbf{A} \mathbf{r}} \, .
\end{equation}
After solving the integral by normal mode transformation, as reported in Appendix I, this model predicts the average energy dissipated in the internal modes to be $\left< U/(k_B T) \right> = 3N/2$. The soft-colloidal representation, instead, has no internal degrees of freedom.
The simulation time $\tilde{t}$, as measured in the MS simulation of the coarse-grained system, translates into the real time $t$ after including the rescaling due to the energy, which is reduced by the amount of energy dissipated in the fluctuations due to internal degrees of freedom.\cite{Rice} For our model
\begin{eqnarray}
t=\tilde{t} R_g \sqrt{\frac{m}{k_BT} \frac{3}{2} N} \ ,
\label{EQ:4}
\end{eqnarray}
with the particle mass, $m$, and size $R_g$.
This rescaling slows down the coarse-grained dynamics, but only partially accounts for the observed phenomenon because the rescaling of the friction needs to be included.
\subsection{Monomer friction coefficient}
\label{SX:mf}
The rescaling of the friction coefficient is calculated considering the friction of the polymer center-of-mass in the monomer/UA representation, and comparing the result with the friction of the cm of a soft colloidal particle. The expression for each of the friction coefficients is derived from its definition as the integral of the memory function contribution to the Generalized Langevin Equation (GLE) in the two levels of representation.
The effect of coarse-graining the Liouville equation, or projection onto the slow degree of freedoms, is the appearance in the Langevin equation of the dissipation terms, given by the random force and the friction coefficient. Systems with different levels of coarse graining have different friction and, as a consequence, different diffusion coefficients.
For a particle in a liquid, the center of mass mean-square displacement is defined as
\begin{equation}
\langle \Delta R^2(t) \rangle = 6\,D\,t \ ,
\end{equation}
with $D$ the diffusion coefficient. For a polymer, the cm diffusion coefficient is given by $D=k_BT/(N \zeta_{m})$, where $\zeta_m$ is the friction coefficient of a monomer, while for a liquid of soft colloidal particles $D_{soft}=k_BT/\zeta_{soft}$, with $\zeta_{soft}$ the friction coefficient of the colloidal particle. The two should be identical in the long-time limit, but they are not, as the diffusion coefficient obtained from MS MD simulation is much larger (much faster dynamics) than the one obtained from UA MD. The correction factor to scale down the MS MD diffusion coefficient, $D^{MS}$, is
$\zeta_{soft}/(N\zeta_{m})$, which yields the rescaled mean-square displacement
\begin{equation}
\langle \Delta R^2(t) \rangle = 6\,D^{MS} \frac{\zeta_{\text{soft}}}{N \zeta_m}\,t \ .
\label{dmsris}
\end{equation}
The thermodynamic conditions of the system under study, i.e. density and temperature, and its molecular structure, i.e. the radius-of-gyration, enter the equation above both directly through the definitions of the friction coefficients, Eqs.(\ref{ciccio}) and (\ref{EQ:zeta_soft_mean_force}) and indirectly through the mesoscale simulation from which the diffusion coefficient, $D^{MS}$, is measured.
To solve Eq.(\ref{dmsris}) we start from the definition of the monomer friction coefficient, $\zeta_m$, which is given in the Markov limit by the memory function
\begin{eqnarray}
\zeta_{m} & \cong & \frac{1}{N} \sum_{a,b=1}^{N} \int_0^{\infty} d \tau \Gamma_{a,b} (\tau) \ .
\label{EQ:zetamo}
\end{eqnarray}
$\Gamma_{a,b}(t)$ is the function that
describes the correlation, through the polymer chain between monomers $a$ and $b$, of the random forces generated from the random collisions of the surrounding molecules undergoing Brownian motion, with \cite{Opp,manychains,schweizer}
\begin{multline}
\Gamma_{a,b}(t) \cong \frac{\beta}{3} \rho \int d\mathbf{r}
\int d\mathbf{r'} g(r) g(r') F(r) F(r')\, \mathbf{\hat r} \cdot \mathbf{\hat r'} \\
\times\int d \mathbf{R} \,
S^Q_{a,b}(R;t) S^Q(|\mathbf{r}-\mathbf{r'}+\mathbf{R}|;t) \, ,
\label{EQ:MFMON}
\end{multline}
\noindent where $g(r)=h(r)+1$ is the monomer radial distribution function, $F(r)$ is the total force exerted by all the matrix polymer on the monomer, and $S^Q(r;t)$ is the projected dynamic structure factor of the matrix fluid surrounding the polymer. The unit vectors $\mathbf{\hat r}$ and $\mathbf{\hat r'}$ define the directions of the total exerted forces. The derivation of Eq.(\ref{EQ:MFMON}) is not completely new and is briefly reported in Appendix II.
Eq.(\ref{EQ:MFMON}) rests on the approximations that the fluid is isotropic and that many-body correlation functions can be described with good accuracy as products of pair distribution functions. The solution of this equation is sometimes carried on by introducing a mode-coupling approximation,\cite{manychains,schweizer,fuchs} however we follow a different procedure. The dynamic structure factor, which is ruled by the projected dynamics, is approximated by its real dynamics counterpart, $S^Q(r;t) \approx S(r;t)$, simplifying the solution of Eq.(\ref{EQ:MFMON}). This is an acceptable approximation when the Langevin equation is expressed as a function of the slow variables\cite{Zwanzing} and holds for our system in the long-time, diffusive regime\cite{Marina}.
In order to separate the spatial coordinates of $S(|\mathbf{r}-\mathbf{r'}+\mathbf{R}|;t)$ in Eq.\ (\ref{EQ:MFMON}) it is convenient to use the Fourier transform
\begin{equation}
S(r;t)=\frac{1}{(2 \pi)^3} \int e^{i \mathbf{k r}} S(k;t) \, d \mathbf{k} \ ,
\label{EQ:FT}
\end{equation}
where the dynamic structure factor is calculated in reciprocal space as the sum of intra- and inter-molecular contributions
\begin{multline}
S(k,t) =\frac{1}{N}\sum_{\alpha\gamma}S_{\alpha\gamma}(k,t)=\frac{1}{N}\sum_{\alpha\gamma}\omega_{\alpha\gamma}(k,t)\\
+\rho \frac{1}{N}\sum_{\alpha\gamma}h_{\alpha\gamma}(k,t) \, .
\end{multline}
Here $\omega_{\alpha\gamma}(k,t)$ is the time dependent intramolecular probability distribution functions for monomers $\alpha$ and $\gamma$, on the same molecule, to be separated by a reciprocal distance $k$, while $h_{\alpha\gamma}(k,t)$ is the corresponding intermolecular contribution.
Given that the dynamics on the global scale is driven by the polymer diffusion, the intramolecular probability distribution function in reciprocal space can be expressed, in the limit of large lengthscales, $k \le 1/R_g$, as
\begin{equation}
\omega_{\alpha\gamma}(k;t) \approx \exp\left[{-\frac{k^2 l^2|\alpha-\gamma|}
{6}}\right] \exp\left[{-k^2 D t}\right] \, ,
\end{equation}
where $D$ is the polymer center-of-mass diffusion coefficient and $l=N^{-1}\sum_{i=1}^N |{\bf{l}_i} |$ is the average segmental length.
Because in Eqs.(\ref{EQ:zetamo}) and (\ref{EQ:MFMON}) the order of the summation and time integrals can be changed, the double summation reduces the inter- and intramolecular distributions to their averages over the bead distribution.
The site-averaged intramolecular probability distribution function, $\omega_0(k)$, is well approximated by the Debye formula,\cite{DoiEdw}
\begin{equation}
\omega_0(k) = \frac{1}{N}\sum_{\alpha\gamma}\omega_{\alpha\gamma}(k)=
\frac{2 N (e^{k^2 R_g^2}+k^2 R_g^2-1)}{k^4 R_g^4} \ ,
\label{EQ:Ave_wofk}
\end{equation}
\noindent or by its Pade' approximant
\begin{equation}
\omega_0(k) \approx \frac{N}{1+k^2 \xi_c^2} \ .
\label{EQ:PADE}
\end{equation}
The site-averaged intermolecular probability distribution is defined by the Ornstein-Zernike equation
\begin{equation}
h(k)=\frac{1}{N}\sum_{\alpha\gamma}h_{\alpha\gamma}(k)=\frac{\omega_0^2(k)c(k)}{1-\rho c(k)\omega_0(k)} \, ,
\label{EQ:OZ}
\end{equation}
where $c(k)$ is the direct correlation function. At the monomer level we follow Curro and Schweizer's PRISM thread approach,\cite{PRISM,PRISM1} where the polymer chain is modeled as a thread of vanishing thickness, $c(k) \approx c_0$, with $c_0=-(1-2 \xi_{\rho}^2/R_g^2)/(2 N^2 \rho_{ch} \xi_\rho^2/R_g^2)$. Substitution of $c_0$ and Eq.\ (\ref{EQ:PADE}) into Eq.\ (\ref{EQ:OZ}) gives
\begin{equation}
h(k)=\frac{h_0}{(1+ k^2 \xi_\rho^2 )(1+ k^2 \xi_c^2)} \ ,
\label{EQ:OZnew}
\end{equation}
where $h_0=h(k=0)= (\xi_\rho^2/\xi_c^2 -1)/\rho_{ch}$ is related to the compressibility of the system.\cite{YAPRL,melt}
Because in the large length scale regime, of interest here, the relaxation of the liquid is dominated by the polymer diffusion, the dynamic structure factor is approximated as
\begin{equation}
S(k;t)\approx S(k) \exp{\left[- k^2 D t \right] } \, .
\end{equation}
Finally, after introducing the integral representation of the delta function
\begin{eqnarray}
\int d \mathbf{R} \ e^{i \mathbf{R}(\mathbf{k}_1+\mathbf{k}_2)}=
(2\pi)^3 \ \delta(\mathbf{k}_1+\mathbf{k}_2) \, ,
\end{eqnarray}
the last integral in Eq.(\ref{EQ:MFMON}) simplifies to
\begin{multline}
\int d \mathbf{R} \, S(R;t) S(|\mathbf{r}-\mathbf{r'}+\mathbf{R}|;t) =
\frac{1}{(2\pi)^6} \int d \mathbf{k}_1 \int d \mathbf{k}_2 \\
S(k_1;t) \ S(k_2;t) e^{i \mathbf{k}_2(\mathbf{r}-\mathbf{r'})}
\int d \mathbf{R} \ e^{i \mathbf{R}(\mathbf{k}_1+\mathbf{k}_2)} \ .
\end{multline}
Because the functions $\omega_0(k)$ and $h(k)$ are even with respect to $k$, the equation reduces to three contributions: the first is due to intramolecular interactions $\omega_0^2(k)$, the second includes the cross product $\omega_0(k) h(k)$, and the last is due to the intermolecular contribution $h^2(k)$.
This leads to the following expression
\begin{multline}
\int d \mathbf{R} \, S(R;t) S(|\mathbf{r}-\mathbf{r'}+\mathbf{R}|;t) =
\frac{1}{(2 \pi)^3} \int d \mathbf{k}\ \\
(\omega_0^2(k)+2 \rho h(k)\omega_0(k)+\rho^2 h^2(k)) e^{-2 k^2D t} e^{i \ \mathbf{k}(\mathbf{r}-\mathbf{r'})} \ .
\label{EQ:k_int_mon}
\end{multline}
Because we are assuming that monomers are interacting through a hard core potential, which is consistent with the PRISM thread model,\cite{PRISM,PRISM1} the force is a delta function and therefore
\begin{equation}
g(r)F(r)=g(d)\beta^{-1}\delta(r-d)\, .
\label{EQ:Hard_core}
\end{equation}
where $d$ is a hard core diameter, identical for any $CH_2$ bead in the chain, in the spirit of the UA-MD description and PRISM approach. When we compare our equations with data of experimental or simulated systems, where monomers interact through a Lennard-Jones potential, the latter has to be mapped onto a hard-core potential with the effective diameter, $d$.\cite{mcquarrie}
The final expression for the monomer friction coefficient is given by
\begin{multline}
\zeta_m = \frac{1}{48\, \pi^3}\, \rho g^2(d)\, (\beta D)^{-1} \\
\times \left\{ J[\omega_0(k),\omega_0(k)]+2\rho J[\omega_0(k),h(k)]+\rho^2 J[h(k), h(k)] \right\} \ ,
\label{EQ:zetam}
\end{multline}
with the function
\begin{eqnarray}
\label{jei}
J[\alpha(k),\beta(k)] &=& \int d \mathbf{r} \int d \mathbf{r'} \int_0^\infty d k\, \frac{\sin(k |\mathbf{r}-\mathbf{r'}|)}{k|\mathbf{r}-\mathbf{r'}|}\, \mathbf{\hat r} \cdot \mathbf{\hat r'}\, \nonumber \\
&&\times \delta(r-d) \delta(r'-d)\, \alpha(k) \beta(k) \ .
\end{eqnarray}
The solution of Eqs.(\ref{EQ:zetam}) and (\ref{jei}) is given by a lengthy but analytical expression, which is a function of the molecular parameters, $\xi_{\rho}$, $R_g$, thermodynamic parameters, $\rho$, $\beta$, the diffusion coefficient, $D$, and of the hard-core diameter $d$, as
\begin{widetext}
\begin{eqnarray}
\label{ciccio}
\zeta_m & \approx & \frac{2}{3} (D\beta)^{-1} \rho g^2(d) \Biggl(\frac{1}{12}\pi N^2 d^2 R_g \Biggl[ 15\sqrt{2} + 40 \frac{d}{R_g} + 12\sqrt{2}\left(\frac{d}{R_g}\right)^2 \Biggr]+ \rho \pi N h_0 \frac{1}{3\sqrt{2}(R_g^2-2\xi_\rho^2)^2} \Biggl[ 12\sqrt{2}\xi_\rho^7 \\
&& + 12 d^4 R_g^3 \left(1-2\left(\frac{\xi_\rho}{R_g} \right)^2\right) + \nonumber 4\sqrt{2} d^3 R_g^4 \left( 5-14 \left(\frac{\xi_\rho}{R_g}\right)^2 + 2 \left(\frac{\xi_\rho}{R_g}\right)^4 \right) + 3 d^2 R_g^5 \Biggl( 5- 14 \left(\frac{\xi_\rho}{R_g}\right)^2 - 4\sqrt{2} \left(\frac{\xi_\rho}{R_g} \right)^5 \Biggr) \nonumber \\
&& - 12\sqrt{2} e^{-\frac{2d}{\xi_\rho}} \xi_\rho^7 \left(1+\frac{d}{\xi_\rho} \right)^2 \Biggr] +
\rho^2 \pi h_0^2 \frac{1}{12(R_g^2-2\xi_\rho^2)^3} \Bigg[ 40 d^3 R_g^6 + 15\sqrt{2} d^2 R_g^7 - 24\sqrt{2}d^4 R_g^3 \xi_\rho^2 - 144 d^3 R_g^4 \xi_\rho^2 \nonumber \\
&& + 6 \sqrt{2} d^4 R_g^5 \left(2 - 9 \left(\frac{\xi_\rho}{d}\right)^2 \right) + 12 R_g^2 \xi_\rho^7 \left(4 \left(\frac{d}{\xi_\rho}\right)^3 - 7 \left(\frac{d}{\xi_\rho}\right)^2 + 9 \right) - 8\xi_\rho^9 \left(4 \left(\frac{d}{\xi_\rho}\right)^3 - 9 \left(\frac{d}{\xi_\rho} \right)^2 + 15 \right) - \nonumber\\
&& e^{-\frac{2d}{\xi_\rho}} 12 \xi_\rho^4 (d + \xi_\rho) \left( R_g^2 (d+3\xi_\rho)(2d+3\xi_\rho)-2\xi_\rho^2(2d^2+5\xi_\rho d + 5 \xi_\rho^2) \right) \Biggl] \Biggl) \ . \nonumber
\end{eqnarray}
\end{widetext}
This expression is general and holds for any homopolymer melt represented as a collection of identical beads interaction through a hard-core potential of range $d$. The value of $d$ is specific of the monomeric structure of the homopolymer.
\subsection{Friction coefficient for a liquid of interacting soft colloidal particles}
\label{SX:fc}
The friction coefficient for a point particle interacting through a soft repulsive potential is much smaller than the friction of the macromolecule before coarse-graining. In fact, the friction coefficient of an object can be estimated using Stokes' formula where $\zeta=6 \pi \eta r_H$, with $\eta$ the fluid viscosity and $r_H$ the hydrodynamic radius. The latter can be evaluated from the surface area of the object exposed to the solvent, which can be estimated by "rolling" a solvent molecule on the object. It is evident that the surface available to the solvent in a bead-spring representation of a polymer is much higher than the surface available to the solvent for a point particle interacting through a soft, long-ranged potential.
To calculate the friction coefficient for a soft colloidal particle, we start from the Generalized Langevin Equation that describes the time evolution for the position coordinate of the molecular center-of-mass, i.e. Eq.(\ref{GLEsoft}) where the friction coefficient for soft particles is given by
\begin{eqnarray}
\zeta_{\text{soft}}&\cong& (\beta/3) \ \rho_{ch} \int_0^{\infty}d\,t \ \int d\mathbf{r} \int d\mathbf{r'} g(r) g(r') F(r) F(r')\, \nonumber \\
&&\times\mathbf{\hat r} \cdot \mathbf{\hat r'} \int d \mathbf{R} \, S(R;t) S(|\mathbf{r}-\mathbf{r'}+\mathbf{R}|;t) \, .
\label{EQ:zeta_soft}
\end{eqnarray}
Eqs.(\ref{EQ:MFMON}) and (\ref{EQ:zeta_soft}) look identical, with just a different form of the density prefactor. In reality the form of the pair-distribution function, $g(r)$, the force exerted by the surrounding molecules on the tagged chain, $F(r)$, and the dynamic structure factors, $S(\mathbf{R},t)$, are different quantities in the monomer and soft-colloid representations.
We assume that the dynamic structure factor in reciprocal space has the form
\begin{equation}
S(k;t)\approx S(k)\ e^{-D t k^2}=(1+\rho_{ch} h^{cc}(k))\ e^{-D t k^2}
\label{EQ:SFSC}
\end{equation}
where $h^{cc}(k)$ is the center of mass total pair correlation function
\begin{equation}
h^{cc}(k) = h_0 \left[\frac{1+k^2 R_g^2/2}{1+k^2 \xi_\rho^2} \right] e^{-\frac{k^2 R_g^2}{3}} \, .
\label{EQ:hcck_OZ}
\end{equation}
with $h_0=(\xi_\rho^2/\xi_c^2 -1)/\rho_{ch}$, as defined in the previous section.
Eq.(\ref{EQ:hcck_OZ}) is just the Fourier transform of Eq.(\ref{EQ:1}).
Eq.(\ref{EQ:SFSC}) indicates that in the long-time regime, which is of interest here, the relaxation of the liquid is largely driven by the center-of-mass diffusion, while internal dynamics and local modes of motion are already fully relaxed. This is a reasonable assumption given that the lengthscale of our treatment is the overall polymer dimension, and no structural or dynamical information is retained on the local scale.
To perform our calculation we need to defined an approximate analytical form of the effective force. To do so we adopt the simplified form of $h^{cc}(r) $, Eq.(\ref{EQ:2}). Then, we reduce further the expression by neglecting the small attractive component of the potential. Finally we approximated the real potential, $v(r)$, with its mean-force counterpart $w(r)\approx - k_B T \ln{[h(r)+1]}$ properly rescaled. The real potential, calculated through the HNC approximation as described in the following section, is a complicated function of $h(r)$. However it can be related, in an approximated way, to the simpler potential of mean force through the equation
\begin{equation}
v(r)\approx \frac{v(0)}{w(0)}w(r) \ ,
\end{equation}
where $v(r) \approx \sqrt{3} w(r)$ for all the samples considered in this study.
These approximations define the force $F(r)$ and the pair distribution function, $g(r)=h(r)+1$, entering the equation for the friction coefficient.
The resulting expression for the friction coefficient of the soft colloidal particle is expressed as a function of the diffusion coefficient $D$, $\beta$, $\rho_{ch}$ and the two length scales $R_g$ and $\xi_\rho$ as
\begin{eqnarray}
\zeta_\text{soft} &\cong& \frac{4}{3} \sqrt{\pi} (D \beta)^{-1} \rho_{ch} R_g \xi_\rho^2 \left(1+\frac{\sqrt{2}\xi_\rho}{R_g} \right)^2 \frac{507}{512} \nonumber\\
&&\times\left[ \sqrt{\frac{3}{2}} + \frac{1183}{507} \rho_{ch} h_0 + \frac{679\sqrt{3}}{1024} \rho_{ch}^2 h_0^2 \right] \, .
\label{EQ:zeta_soft_mean_force}
\end{eqnarray}
This expression is an approximated analytical form for the friction coefficient of a soft colloidal particle.
\section{Mesoscale Simulations}
\label{SX:ms}
Here we present numerical calculations to illustrate and discuss the rescaling procedure of the preceding sections. We first perform MS MD simulations of the coarse-grained polymer liquid, where each chain is represented as a soft colloidal particle, centered at the center-of-mass of a chain, and interacting with the surrounding particles through a soft repulsive potential of the order of few times the chain dimension, $R_g$. The simulations of the soft colloidal liquid produce dynamical properties that are accelerated due to the soft nature of the potential in the coarse-grained representation. These properties are rescaled following our procedure, and then compared with existing data, when they are available.
In a previous paper we briefly presented calculations of the rescaled dynamics for a variety of systems including UA MD simulations and experimental data of PE diffusion available in the literature, that we use to test the accuracy of our procedure. We selected UA MD simulations as our test (see Table \ref{TB:parameters_UAMDa}) because they have been shown to reproduce with a high level of accuracy the dynamical properties of PE melts, such as diffusion and viscosity.\cite{PaulSmith,PaulSmith1} We also compared predictions of rescaled MS-MD simulations with experiments for samples with temperature $T=509 \ K$, monomer site density $\rho =0.0315302$~sites/\AA$^3$ and $N= 36, \ 72, \ 106, \ 130, \ 143, \ 192,$ and $ 242 $. \cite{Richter,maybeothers,maybeothers1,maybeothers2,maybeothers3}. Our MS-MD simulations, properly rescaled, provided good quantitative predictions of the diffusion coefficient for those systems.\cite{shortivan}
In this paper we use those same systems to illustrate our procedure. Moreover we present new results for PE samples, not present in the literature, to underline the predictive power of the theory, where no calibration curve is necessary. Once a system is selected, its structural and thermodynamic parameters are defined and are used as input to the MS MD simulation so that the whole procedure is free of adjustable parameters, with the exception of the parameter $d$ that is fixed for PE once and for all samples.\cite{MONDE,JARAM,HEINE,Mavran}
Systems that we simulated include liquids of chains with increasing degree of polymerization, as described above.
As the molecular weight of the polymer increases, the systems cross the threshold from unentangled to entangled dynamics. For entangled systems the dynamical rescaling approach that we propose is modified to include a one-loop perturbation that accounts for the presence of entanglements. Simulations of soft colloidal liquids are performed for entangled systems, and the rescaling applied to predict diffusion.
\begin{table}[h]
\centering
\caption{Polyolefin melts UA MD simulation parameters}
\bigskip
\begin{tabular}{lcccc}\hline \hline
\mbox{{\it System}}
& $T$[K] & $\rho$[sites/\AA$^{3}$] & $(R_g^{UA})^2$[\AA$^2$] & $(R_g^{FRC})^2$[\AA$^2$] \\%& $\xi_\rho$[\AA] & $h_0$[\AA$^{3}$] \\
\hline
PE 30$^a$ & 400 & 0.0317094 & 63.5695 & 81.7544 \\
PE 44$^a$ & 400 & 0.0323951 & 110.3197 & 127.6856 \\
PE 48$^b$ & 450 & 0.0314487 & 111.0832 & 140.8119 \\
PE 66$^a$ & 448 & 0.0328993 & 177.5348 & 199.8812 \\
PE 78$^b$ & 450 & 0.0321465 & 205.9221 & 239.2607 \\
PE 96$^a$ & 448 & 0.0328194 & 281.7989 & 298.3301 \\
PE 122$^b$ & 450 & 0.0325479 & 346.2655 & 383.6526 \\
PE 142$^b$ & 450 & 0.0326600 & 420.7070 & 449.2852 \\
PE 174$^b$ & 450 & 0.0327680 & 525.1816 & 554.2975 \\
PE 224$^b$ & 450 & 0.0328835 & 690.5038 & 718.3791 \\
PE 270$^b$ & 450 & 0.0329520 & 856.4648 & 869.3342 \\
PE 320$^b$ & 450 & 0.0330034 & 980.1088 & 1033.4158 \\
\hline \hline
\multicolumn{4}{l}{$^a$ from Refs. \cite{MONDE,JARAM,HEINE}; $^b$ from Ref. \cite{Mavran,Mavran1}}
\end{tabular}
\label{TB:parameters_UAMDa}
\end{table}
Details about our MS MD simulations have been reported in previous papers of ours and will not be repeated here.\cite{jaypaper,shortivan,multis} Briefly, MS MD simulations were implemented in the microcanonical $(N,V,E)$ ensemble on a cubic box with periodic boundary conditions. We used reduced units such that all the units of length were scaled by $R_g$ ($\tilde{r}=r/R_g$) and energies were scaled by $k_B T$. Temperature and radius-of-gyration were utilized for dimensionalizing the results obtained from the MS MD simulations, after they were performed.
The number of particles, i.e. polymer chains, in our simulations varies from $1728$ ($N=40$) to $85184$ ($N=1000$) depending on the system. This number is determined by the box size, which is larger than twice the range of the potential, and by the liquid density. The potential is long-ranged, due to the many-body effects entering through the OZ equation.
Each simulation evolves for about $50,000$ computational steps. For the entangled melts the potential is longer ranged than for the unentangled systems, and therefore it is cut at larger distances requiring a bigger box size. The reduced density used in simulations, $\rho^{sim} = \rho R_g^3$ where $\rho$ is the site density, varies around $1$ for unentangled melts, and exceeds $2$ for weakly entangled melts. A typical MS simulation takes between $2$ hours ($N=40$) to $4$ days ($N=200$) on one CPU workstation, while using the code that works in parallel, the computational time is further reduced.
\subsection{Interparticle potential}
\label{SX:ip}
The pair potential acting between two effective coarse-grained units is formally derived from the colloidal representation of the liquid, specifically $h(r)$, using an hyper-netted-chain (HNC) closure approximation to the Ornstein-Zernike equation.\cite{SMPLQ} This approximation is known to work well for liquids of particles interacting through a soft potential.\cite{mcquarrie}
The potential input to the MS MD simulation, $v^{cc}(r)$, is derived from the total correlation function for the soft colloidal representation of the liquid,
$h^{cc}(r)$, defined in the limit of long chains, $N\rightarrow \infty$, as in Eq.(\ref{EQ:2}).
The potential is calculated using the hypernetted chain approximation as
\begin{equation}
\beta v^{cc}(r) = h^{cc}(r)-ln[1+h^{cc}(r)]-c^{cc}(r) \ .
\label{EQ:HNC}
\end{equation}
Here the direct correlation function, $c^{cc}(r)$ is given in reciprocal space in terms of $h^{cc}(k)$ as:
\begin{equation}
c^{cc}(k) = \frac{h^{cc}(k)}{1+\rho_{ch} h^{cc}(k)} \ .
\label{EQ:ccck}
\end{equation}
It is important to define the correct potential acting between the coarse-grained units to achieve a realistic representation of the large scale properties of a system through MS MD. Because coarse-grained potentials result from the mapping of many-body interactions into pair interactions, through the averaging over microscopic degrees of freedom, they are parameter dependent. During the coarse-graining procedure, the potential acting between microscopic units, which is given by the Hamiltonian of the system, reduces to an effective potential, which is a free energy in the reference system of the microscopic coordinates. The coarse-grained potential so obtained contains contributions of entropic origin due to the microscopic, averaged-out degrees of freedom and is therefore state-dependent. This can be observed in the form of the total correlation function between coarse-grained sites, Eq.(\ref{EQ:2}) from which the potential is derived. The correlation function explicitly includes the structural and thermodynamic parameters of the polymer, i.e. the radius-of-gyration, density screening length, and number density. The temperature enters directly through Eq.(\ref{EQ:HNC}) and indirectly through the molecular parameters, such as $R_g$.
\subsection{Results from mesoscale molecular dynamics simulations}
\label{SX:rf}
Before entering the details of applying our rescaling approach we focus on the "raw" dynamics obtained directly from the MS MD simulations. Fig.~\ref{FG:MS_MSD_vtcf} displays the mean-square displacement for the MS MD simulation of a polyethylene melt with $N=44$. At short times the inertial term in the Langevin equation is dominant as the particles undergo ballistic dynamics, while in the long-time regime the system crosses over to diffusive dynamics. The diffusion coefficient is higher than the value measured in UA MD simulations, and the transition from ballistic to diffusive regime happens after about $5000$ simulation steps (dot-dot vertical line on top panel of Fig.~\ref{FG:MS_MSD_vtcf}), which corresponds to a distance of roughly $30 R_g$. Such a large distance reflects the fact that in MS MD simulation the point particles interact through a very soft potential and the density is also very low. Because the particle has to "collide" many times to undergo the crossover to diffusive dynamics the latter takes place at a large lengthscale.
Moreover, the bottom panel in Fig.~\ref{FG:MS_MSD_vtcf} displays the velocity correlation function $\left<(v(t)-v(0))^2\right>$ and shows that, consistent with the mean-square-displacement, once more the inertial term becomes negligible at the same crossover time that the diffusive regime sets in. Since our MS MD simulation are performed at equilibrium, the avarage kinetic energy per particle $\left<m v^2(t)/2\right> = 3/2\, k_BT$, and therefore
\begin{eqnarray}
\lim_{t\rightarrow\infty} \left<(v(t)-v(0))^2\right> &=& \lim_{t\rightarrow\infty} \left[2\left<v^2(t)\right>-2\left<v(t) v(0)\right>\right] \nonumber\\
&=& 6 k_BT/m\, .
\end{eqnarray}
The dynamical transition is displayed as a dashed line on the bottom panel of Fig.~\ref{FG:MS_MSD_vtcf}, taking into account that our simulations are in reduced units and $m=1$, $k_BT=1$. The figure shows that the velocity autocorrelation function reaches it's asymptotic value at about the same time as the diffusive regime sets in for the mean-square-displacement.
\begin{figure}[]
\centering
\includegraphics[scale=0.35]{MS_MSD_vtcf.pdf}
\caption{Top panel: cm mean-square displacement (solid line) from MS MD simulation in reduced units as a function of simulation time steps for the $PE 44$ melt sample. The slopes for the ballistic and the diffusive regimes are shown as dashed and dot-dashed lines correspondingly. Bottom panel: cm velocity time correlation function showing when the inertial term becomes negligible. The asymptotic value of $6 \ k_BT/m$ is depicted as a dashed line.}
\label{FG:MS_MSD_vtcf}
\end{figure}
\section{Application of the Rescaling Procedure}
\label{SX:ao}
As stated above, the accelerated dynamics that is a consequence of the coarse graining of the system can be rescaled by taking into account the two main effects of the procedure, namely the change in the entropic contribution to the free energy of the system due to the averaging of the internal degrees of freedom and the change in the friction coefficient due to the different shapes of the molecule in the two different levels of coarse-graining. The difference in shape relates to the change in the molecular surface available to the surrounding molecules, and to the correlation of the random forces generated
by intermolecular collisions.
The first rescaling is given by the inclusion \textit{a posteriori} of the internal degrees of freedom, averaged out during the coarse-graining procedure, as a correction term in the free energy of the system, which accounts for the difference in entropy. The energy correction affects the time of the measured dynamics as the change from a bead-spring description to a soft-colloid representation leads to the rescaling of the time reported in Eq.(\ref{EQ:4}), also taking into account the fact that because the potential is expressed in normalized quantities, the simulation runs using reduced units of energy, $k_BT=1$ and the normalized length $r/R_g$.
The second rescaling of the dynamics is calculated starting from the ratio between the friction coefficients in the two coarse-grained representations, as described in Eq.(\ref{dmsris}).
\subsection{Calculations of the monomer friction coefficient,~$\zeta_m$}
\label{pip}
Because our formalism maps the Lennard-Jones liquid described by the UA MD simulation into a liquid of polymers interacting through a monomer hard-core repulsive potential, it is necessary to define an effective hard-core diameter, $d$. This is done by requiring the friction of the chain with $N=44$ to follow the expected scaling behavior for the diffusion of an unentangled polymer chain, $D=k_B T/ (N \zeta_m)$. Since all except two of the atoms in our PE chains are $CH_2$ monomers, we assume that the potential is identical for all the units along the homopolymer chain. Moreover we assume that the range of the repulsive interaction, $d$, is independent of liquid density.\cite{hardsphere}
Among the different samples, we selected the chain with $N=44$ to optimize $d$, because this sample follows unentangled dynamics while the polymer is long enough to obey the Gaussian intramolecular distribution of monomer positions, which justifies the analytical form of the intramolecular structure factors used in our formalism.
Fig.~\ref{FG:zeta_of_d} displays the monomer friction coefficient, from Eq.(\ref{ciccio}), expressed as the dimensionless quantity $D \beta \zeta_m$, as a function of hard sphere diameter $d$, for polyethylene melts of three different degrees of polymerization. The $1/N$ scaling is reported as a dot-dashed line in the figure.
In these calculations, the numerical values of $N$, $\rho$ and $R_g$ were taken from the data of the UA MD simulation against which the proposed approach is tested. The value of the radial distribution function at the contact was set to $g(d)=1/2$, which is the conventional value assumed in the PRISM thread theory for polyethylene chains. This value is intermediate between zero and the first solvation shell value. The optimized hard-core diameter for $N=44$ is $d=2.1\AA$, which is an intermediate value between the bond length, $l=1.54$ \AA, and the Lennard-Jones $\sigma$-parameter, $\sigma=3.95$ \AA, in the intermonomer potential of the UA MD.\cite{JARAM,HEINE} The unentangled scaling is fulfilled for PE30 at $d=2.07$ \AA \ and for PE96 at $d=1.96$ \AA, which are close to the one for PE44. Because the PE96 sample has a degree of polymerization that is close to the entanglement value of $N_e=130$, its dynamics is likely to be in the crossover regime where the effect of entanglements start to be felt, since the transition from the unentagled to the entangled dynamics is very broad.
\begin{figure}[]
\centering
\includegraphics[scale=0.30]{zeta_mon_of_d_3in1.pdf}
\caption[Plot of $\zeta(d)$]{Plot of $D \beta \zeta_m$ as a function of hard core diameter $d$. From left to right, the three panels show curves for polyethylene melt with $N=30$, $N=44$, and $N=96$.}
\label{FG:zeta_of_d}
\end{figure}
Table \ref{TB:mon_hard_sphere} displays the numerical values of the dimensionless monomer friction coefficient, $D \beta \zeta_m$, for polymeric liquids with different degree of polymerization, $N$, across the untentengled-to-entangled transition. For both unentangled and entangled systems the hard sphere diameter has been fixed to the value of the unentangled ones, $d=2.1$\AA, so that the intermolecular monomer potential is not changed as a function of $N$. While for unentangled systems the monomer friction coefficient was calculated from Eq.(\ref{ciccio}), for entangled chains we adopted a perturbative approach to account for the effect of entanglements.
Let's denote
\begin{equation}
D\beta \zeta_m = J(\rho, N, R_g, d) \, ,
\label{EQ:bulky2}
\end{equation}
where for unentangled systems $J(\rho, N, R_g, d) \approx N^{-1}$. Following a one-loop perturbation, and including the definition of the diffusion coefficient for a macromolecule comprised of $N$ monomers with $\zeta_m$ the monomer friction coefficient $D=(\beta \zeta_m N)^{-1}$, the normalized and perturbed friction coefficient becomes
\begin{equation}
D\beta\zeta_m' = N D\beta \zeta_m J(\rho, N, R_g, d) = N (D\beta \zeta_m)^2\ .
\label{entangly}
\end{equation}
The one loop perturbation is in the spirit of the reptation model where both the chain reptating and the chains involved in the entanglements relax with the same diffusive mechanism: each brings a $N^{-1}$ scaling contribution, which is the trademark of polymer Brownian motion. Interestingly, in our model the diffusion coefficient of entangled polymers under certain fixed monomer density and temperature shows apparent scaling exponents different from ${- 2}$ (see Fig.~\ref{FG:diffusion1}). The resulting scaling exponents emerge cumulatively from output of mesoscale simulations and both steps of rescaling.
Because Eq.(\ref{entangly}) applies only when the systems are entangled, to predict the diffusive behavior of new samples it is necessary to estimate a priori the crossover degree of polymerization, $N_e$. Several methods have been presented in the literature to estimate $N_e$ from thermodynamic conditions and molecular parameters.\cite{ennee} Those methods provide similar values of $N_e$. Moreover, the expressions for the unentangled and entangled frictions, Eqs.(\ref{ciccio}) and (\ref{entangly}), predict values that differ only slightly in the crossover region, as it is shown in Fig.~\ref{FG:diffusion1} in this paper. In this way, selecting the unentangled expression to represent entangled systems, or viceversa, in the crossover region would result in small inconsistencies in the calculated diffusion coefficients.
Table \ref{TB:mon_hard_sphere} includes intra- ($D\beta\zeta_m(\omega_{\alpha \gamma})$) and inter-molecular ($D\beta\zeta_m(h_{\alpha \gamma})$) contributions to the monomer friction coefficient, as well as the self intramolecular contribution, $D\beta\zeta_m(\omega_{\alpha\alpha})$. In general, the calculated total friction is comparable in magnitude to the self intramolecular contribution, $D\beta\zeta_m(\omega_{\alpha\alpha})$. Moreover, the total intramolecular contribution, $D\beta\zeta_m(\omega_{\alpha,\gamma})$, is of the same order of magnitude of the itermolecular contribution, $D\beta\zeta_m(h_{\alpha,\gamma})$, but with the opposite sign, which is reasonable as the liquid is almost incompressible. This result shows that the conventional approximation of replacing the structure factor, $S_{\alpha,\gamma}(k)$, by the single chain analog, $\omega_{\alpha,\gamma}(k)$, can lead to errors in the evaluation of the memory function for macromolecular liquids.\cite{Hess} The table also displays the value of the dimensionless friction coefficient $N D \beta \zeta_m$, which for unentangled systems should be $\approx 1$. As expected, we see deviation from the unentangled behavior in the very short chains and in the crossover to entangled dynamics at $N \approx 100$.
\begin{table}[h]
\centering
\caption{Monomer friction coefficient contributions with hard sphere potential for $d = 2.1$\AA}
\bigskip
\begin{tabular}{lccccc}\hline \hline
\mbox{{\it System}}
& $D\beta\zeta_{m}(\omega_{\alpha \alpha}) $ & $D\beta\zeta_{m}(\omega_{\alpha \gamma}) $
& $D\beta \zeta_{m}(h_{\alpha \gamma})$ & $D\beta\zeta_{m}$ & $N D\beta\zeta_{m}$\\
\hline
PE 30 & 0.03378 & 0.08484 & -0.04883 & 0.03601 & 1.0804 \\
PE 44 & 0.02744 & 0.06244 & -0.03991 & 0.02253 & 1 \\
PE 48 & 0.03101 & 0.07984 & -0.04904 & 0.03080 & 1.4782 \\
PE 66 & 0.02539 & 0.05703 & -0.03823 & 0.01880 & 1.2409 \\
PE 78 & 0.02632 & 0.06287 & -0.04179 & 0.02107 & 1.6439 \\
PE 96 & 0.02239 & 0.04766 & -0.03325 & 0.01441 & 1.3834 \\
PE 122 & 0.02400 & 0.05533 & -0.03831 & 0.01701 & 2.0757 \\
PE 142 & 0.02245 & 0.04967 & -0.03501 & 0.01466 & 2.0815 \\
PE 174 & 0.02193 & 0.04827 & -0.03438 & 0.01389 & 2.4170 \\
PE 224 & 0.02133 & 0.04664 & -0.03359 & 0.01304 & 2.9219 \\
PE 270 & 0.02039 & 0.04344 & -0.03164 & 0.01180 & 3.1849 \\
PE 320 & 0.02184 & 0.04956 & -0.03585 & 0.01371 & 4.3874 \\
\hline \hline
\end{tabular}
\label{TB:mon_hard_sphere}
\end{table}
\subsection{Calculation of the friction coefficient of a soft colloid, $\zeta_{soft}$}
Starting from Eq.(\ref{EQ:zeta_soft_mean_force}) we calculated the friction coefficient, $D \beta \zeta_{soft}$, for polymer liquids represented as soft colloidal particles. Table \ref{TB:Soft_MFP} shows the dimensionless friction coefficient for several systems. The molecular parameters, $N$ and $R_g$, and the thermodynamic conditions of density $\rho$ and temperature $T$, are taken from the UA MD simulations, see Table \ref{TB:parameters_UAMDa}.
\begin{table}[h]
\centering
\caption{Soft colloids friction coefficient contributions}
\bigskip
\begin{tabular*}{230pt}{@{\extracolsep{\fill}}lccc}\hline \hline
\mbox{{\it System}}
& $D\beta\zeta_{\text{soft}}^{\text{self}} $ & $D\beta\zeta_{\text{soft}}(h^{cc})$ & $D\beta\zeta_{\text{soft}}(1 + h^{cc})$\\
\hline
PE 30 & 0.044273 & -0.029932 & 0.014341 \\
PE 44 & 0.020769 & -0.012441 & 0.008328 \\
PE 48 & 0.024639 & -0.015289 & 0.009349 \\
PE 66 & 0.012619 & -0.006659 & 0.005960 \\
PE 78 & 0.012019 & -0.006254 & 0.005765 \\
PE 96 & 0.008055 & -0.003712 & 0.004344 \\
PE 122 & 0.007423 & -0.003334 & 0.004089 \\
PE 142 & 0.006159 & -0.002612 & 0.003546 \\
PE 174 & 0.005218 & -0.002108 & 0.003109 \\
PE 224 & 0.004298 & -0.001648 & 0.002650 \\
PE 270 & 0.003660 & -0.001351 & 0.002310 \\
PE 320 & 0.003521 & -0.001288 & 0.002233 \\
\hline \hline
\end{tabular*} \\
\label{TB:Soft_MFP}
\end{table}
The dimentionless friction coefficient for these systems is $D\beta\zeta_{\text{soft}}(1 + h^{cc})\approx0.002-0.01$, while we would expect $D\beta\zeta_{\text{soft}}(1 + h^{cc})\approx 1$ for unentangled systems, see Table \ref{TB:Soft_MFP}. These data show that the theoretically calculated friction coefficient (without rescaling) for the soft colloidal systems greatly underestimate the friction coefficient, as also observed in the MS MD simulations, and hence give rise to accelerated dynamics as discussed previously. It also shows that intra- and inter-molecular contributions to the friction coefficient are comparable in magnitude: both of them need to be taken into account when calculating dynamical properties of polymer melts.
\subsection{Results from the rescaling procedure. Comparison with simulation and experimental data}
\label{SX:rft}
Some of the results reported in this section were already briefly presented in our short paper.\cite{shortivan} Our discussion here makes use of some of those data as a starting point to illustrate with an example the details of the proposed rescaling procedure and highlight its strengths and weaknesses.
In order to rescale the unrealistic fast dynamics of MS MD simulations we applied our rescaling procedure and compare the predicted dynamics with data from UA MD simulations and experiments. We use as input parameters the thermodynamic conditions and molecular parameters of each sample under study. The rescaling procedure is given by Eq.(\ref{dmsris}), where $D^{MS}$ is the diffusion coefficient from the MS-MD simulation, the soft colloid friction coefficient is calculated using Eq.(\ref{EQ:zeta_soft_mean_force}), and the monomer friction coefficient is given by Eq.(\ref{ciccio}) for unentangled chains, and Eq.(\ref{entangly})with Eq.(\ref{ciccio}) for entangled ones. Eq.(\ref{dmsris}) depends on the temperature and density of the system investigated and on its molecular radius-of-gyration.
Indirectly those parameter enter our procedure through the diffusion coefficient from mesoscale simulations, $D^{MS}$. Specifically temperature enters through the rescaling of the time, as the time step in the mesoscale simulation is adimensional and becomes dimensional once it is rescaled by the energy, following a well-established procedure. Moreover, thermodynamic and molecular parameters enter indirectly through the soft potential, Eq.(\ref{EQ:HNC}), which is parametric and includes density, temperature, and the molecular radius-of-gyration.
Finally, thermodynamic parameters enter directly through the definitions of the friction coefficients in the monomer and soft-sphere descriptions, Eqs.(\ref{ciccio}) and (\ref{EQ:zeta_soft_mean_force}) respectively. Specifically, the monomer friction coefficient is a function of $\rho$, $N$, $R_g$, plus a hard sphere diameter, $d$, which is used to map the Lennard-Jones potential of the united atom simulation onto a repulsive hard-core potential with an effective bead diameter. The hard sphere diameter $d$ is assumed to be independent of the thermodynamic conditions, for the range of temperature and density simulated here, and constant for all the monomers in the homopolymer chain. The criteria of choosing numerical value for $d$ have been already explained and discussed.
In an analogous way, the soft-sphere friction coefficient depends on the chain number density, which relates to the monomer number density through $N$ as $\rho_{ch} = \rho/N$, and $R_g$ is the radius of gyration of the polymer chain. It also depends on the density fluctuation length scale $\xi_\rho$, which is expressed as a function of $R_g$ and $\rho_{ch}$ as $\xi_\rho = {R_g}/(\sqrt{2}+2\pi\rho_{ch}R_g^3)$, and on the parameter $h_0=h(k=0)=-(1-2\xi_\rho^2/R_g^2)/\rho_{ch}$. In fact, the dimensionless combination $D\beta\zeta_{\text{soft}}$ is determined by only three parameters: $\rho$, $N$ and $R_g$. In conclusion, once thermodynamic parameters, $R_g$ and $d$ are defined, there are no adjustable parameters in our method.
The predicted diffusion coefficients from our rescaled MS MD simulations are in good agreement with the data for all the test systems.
As an example, Table \ref{TB:D_MS_MD} displays the diffusion coefficients obtained directly from the MS MD, $D^{MS}$, once they are rescaled to include the internal degrees of freedom, and after the second rescaling of the friction, $D_{cm}$, as well as the values of the diffusion coefficient from the UA MD simulations, $D^{UA}$, against which we compare our predicted diffusion. The table shows that while the initial values of the diffusion are orders of magnitude larger than the data from UA MDs, the rescaled coefficients are very close to the real values. For the entangled systems we adopt the perturbative approach described in Section \ref{pip} obtaining predicted values that are in quantitative agreement with the UA MD simulations. Entangled samples are from references \cite{Mavran,Mavran1}, and are mostly in the weakly entangled regime. For these samples the UA MD simulations include a small number of chains: $n=40$ for $N=78$, $n=22$ for $N=142$, $n=32$ for $N=174$, $N=224$, $N=270$ and $N=320$, with $n$ the number of chains in the simulation and $N$ the degree of polymerization. These numbers show one advantage of adopting a coarse-grained description as typically our samples include thousands of chains. Simulating a large ensemble of molecules is necessary, for example, when the goal is to investigate large-scale fluctuations or the relative relevance of intra- vs inter-molecular contributions to the dynamics.
\begin{table}[h]
\centering
\caption{Diffusion coefficients in \AA$^2$/ns from MS MD compared with UA MD simulation}
\bigskip
\begin{tabular*}{230pt}{@{\extracolsep{\fill}}lcccc}\hline \hline
\mbox{{\it System}}
& $T$[K]& $D^{MS}$ & $D_{cm}$ & $D^{UA}$ \\
\hline
PE 30 & 400 & 4.44$\times 10^3$ & 58.9 & 82.9 \\
PE 44 & 400 & 5.29$\times 10^3$ & 44.5 & 46.0 \\
PE 48 & 450 & 5.80$\times 10^3$ & 36.7 & 50.8 \\
PE 66 & 448 & 6.04$\times 10^3$ & 29.0 & 31.8 \\
PE 78 & 450 & 6.73$\times 10^3$ & 23.6 & 26.0 \\
PE 96 & 448 & 6.98$\times 10^3$ & 21.9 & 23.3 \\
PE 142& 450 & 8.45$\times 10^3$ & 6.92 & 7.93 \\
PE 174& 450 & 8.51$\times 10^3$ & 4.53 & 5.70 \\
PE 224& 450 & 8.80$\times 10^3$ & 2.73 & 3.28 \\
PE 270& 450 & 9.39$\times 10^3$ & 2.14 & 2.06 \\
PE 320& 450 & 8.73$\times 10^3$ & 1.03 & 1.30 \\
\hline \hline
\end{tabular*}
\label{TB:D_MS_MD}
\end{table}
\begin{figure}
\centering
\includegraphics[scale=0.4]{fig5crop.pdf}
\caption[]{ Plot of mean-square displacement as a function of time for unentangled $PE$ melts. The rescaled MS MD simulation (line) is compared with UA MD simulation (symbols) for $N=44, 66, 96$. Also shown is the outcome of the theory for cooperative dynamics (dashed lines).}
\label{FG:MSD_PE_approximations}
\end{figure}
Fig.~\ref{FG:MSD_PE_approximations} illustrates how our approach can be used to calculate dynamics also in the short time regime. The figure shows the mean-square-displacement of the center-of-mass from UA MD in comparison with the one calculated from the diffusion coefficient rescaled from the MS MD. The agreement is quantitative in the long time regime. In the short time regime, the UA MD simulation data exhibit a subdiffusive behavior, even if polymers are unentangled. In a series of papers we have shown that the subdiffusive regime is a consequence of the presence of cooperative dynamics involving several polymer chains moving in a correlated way inside the dynamically heterogeneous liquid of macromolecules.\cite{manychains,Marina} A detailed discussion of this phenomenon, which is of intermolecular origin, has been provided before and will not be repeated here. Our theory, the Cooperative Dynamics Generalized Langevin Equation (CD-GLE), needs as an input the diffusion coefficient and predicts the subdiffusive behavior for times shorter than the longest correlation time as a function of the number, $n' \propto \sqrt{N}$, of macromolecular chains moving in a cooperative way. Fig.~\ref{FG:MSD_PE_approximations} shows the results from our CD-GLE calculations as dashed lines. Here, the input monomer friction coefficient is calculated from MS MD using the rescaling procedure. The number of chains undergoing cooperative dynamics is $n'=30$ for $N=96$, $n'=25$ for $N=66$, and $n'=14$ for $N=44$.
The subdiffusive behavior shown in UA-MD data is not visible in the rescaled MS MD data as the dynamics are accelerated. The effective temperature experienced by the polymer is much higher than the temperature in UA MD simulations as the energy is not dissipated in the internal degrees of freedom.
Finally we discuss the calculations of the monomer and soft-colloid friction coefficients for a set of polyethylene chains investigated experimentally.\cite{Richter,maybeothers,maybeothers1,maybeothers2,maybeothers3} The experimental data do not report the values of $R_g$ at the desired thermodynamic conditions, $T=509$ K and density $\rho=0.0315302$ sites/\AA$^3$, while it is known that the chain conformation, and $R_g$, are temperature dependent. To calculate the input parameters for our MS MD simulations we adopt a freely rotating chain model, for which the mean-square end-to-end polymer distance is given by\cite{Yamakawa}
\begin{eqnarray}
\left<R_{ete}^2 \right>=N l^2\left[ \frac{1+g}{1-g}-\frac{2g}{N}\frac{1-g^N}{(1-g)^2} \right] \ ,
\end{eqnarray}
and $R_g^2\approx \left<R_{ete}^2 \right>/6$ for a chain with Gaussian statistics. For polyethylene melts at this temperature the stiffness parameter is $g=0.785$. \cite{Richter}
For the samples investigated in the UA MD simulations, a comparison of the theoretical values of $R_g$ using the freely-rotating-chain model (FRC) and the values measured directly from the UA MD, which are both reported in Table \ref{TB:parameters_UAMDa}, show a reasonable agreement. The agreement is particularly good for the samples that have long chains because for them the hypothesis of a Gaussian intramolecular distribution is well justified. In this framework, the FRC model provides a reliable description of the chain intramolecular structure.
The values of the monomer and soft-colloid friction coefficients for the experimental samples, calculated from Eqs. (\ref{ciccio}) and (\ref{EQ:zeta_soft_mean_force}) respectively, are presented in Table \ref{TB:zeta_EXP}. The Table shows the large difference between the predicted dimensionless friction coefficients, $D \beta \zeta_{soft}$ and $N D \beta \zeta_m$, for the same macromolecule coarse-grained at two different length scales. From the values displayed in Table \ref{TB:zeta_EXP} we calculate the rescaling factor for the friction coefficient measured in MS MD simulations, following the procedure described in this paper. Because the data have different thermodynamic parameters of density and temperature, their scaling behavior cannot be inferred from their plot, even if an apparent $N^{-1}$ scaling is followed by the unentangled samples and the typical reptation $N^{-2}$ scaling by the entangled ones.
\begin{table}[h]
\centering
\caption{Theoretically calculated dimensionless friction coefficient for monomer ($d=2.1$\AA) and soft colloid with $R_g^{FRC}$ for experimental samples}
\bigskip
\begin{tabular}{lccc}\hline \hline
\mbox{{\it System}}
& $(R_g^{FRC})^2$[\AA$^2$] & $D\beta \zeta_{soft} $ & $N D\beta \zeta_m$ \\
\hline
PE 36 & 101.4350 & 0.007846 & 0.5153 \\
PE 72 & 219.5710 & 0.004927 & 0.8946 \\
PE 106 & 331.1465 & 0.004543 & 1.8407 \\
PE 130 & 409.9056 & 0.003497 & 1.5354 \\
PE 143 & 452.5669 & 0.003318 & 1.6822 \\
PE 192 & 613.3669 & 0.002828 & 2.2412 \\
PE 242 & 777.4485 & 0.002500 & 2.8186 \\
\hline \hline
\multicolumn{3}{l}{$T=509$K, $\rho=0.0315302$ [sites/\AA$^3$]}\\
\end{tabular}
\label{TB:zeta_EXP}
\end{table}
\begin{table}[h]
\centering
\caption{Predicted diffusion coefficients in \AA$^2$/ns from MS MD and experimental data from
Refs. \cite{Richter,maybeothers,maybeothers1,maybeothers2,maybeothers3}}
\bigskip
\begin{tabular}{lcccccccc}\hline \hline
{\it System} & PE36 & PE72 & PE106 & PE130 & PE143 & PE192 & PE242 \\
$D_{cm}$ & 111 & 50 & 17 & 14 & 11 & 5.6 & 3.2 \\
$D^{exp}$ & 120 & 41 & 14 & 12 & 8.6 & 6.5 & 4.5 \\
\hline \hline
\end{tabular}
\label{TB:parameters_EXP}
\end{table}
\subsection{Theoretical predictions of diffusion coefficients for polyethylene samples}
In this section we report theoretical predictions from rescaled mesoscale simulations of the diffusion coefficients for a series of PE samples for which data of chain dynamics, either from simulations or from experiments, are not available in the literature. The degree of polymerization of each sample is not larger than the ones already investigated. However, because there are no data to fit any parameter, these calculations illustrate the predictive power of our method. Diffusion coefficients calculated by combining the mesoscale simulations with the rescaling procedure presented in this manuscript, are displayed in Fig.~\ref{FG:diffusion1} as a function of the degree of polymerization.
The set of MS MD simulations is performed for $N=40, 60, 80, 100, 200, 300, 400, 500$ and $1,000$, at constant monomer density $\rho_m = 0.0329497$ [sites/\AA$^3$] and temperature $T=450 \ K$ for all samples. Values of $R_g$ are calculated using a freely rotating chain model. The hard sphere diameter is fixed to the value reported in the previous sections for PE, $d=2.1$\AA, and the pair distribution function at contact is $g(0)=1/2$ as described early on in this paper. While the simulations of the small samples can be performed on a single CPU machine, for systems with a higher degree of polymerization is convenient to adopt parallel computing. For those systems, simulations were run using the LAMMPS code\cite{LAMMPS}, with our potential as an input, remotely on a 64 CPU machine available through the TeraGrid\cite{TERAGRID}. For the PE1000 sample, which included 85,184 molecules, results were obtained after one week of calculations. By comparison with our single CPU calculations, running the simulation in parallel reduces the computer time by a factor of $10^2$. The number of particles in the simulation is determined by the length of the box size, which for PE1000 sample is equal to $24$ $R_g$, i.e. larger than twice the range of the potential, to eliminate molecular self-interaction through the periodic boundary conditions.
While we assume that small changes of density and temperature do not affect the hard-core diameter, $d$, even a small difference in $\rho_m$ can noticeably change the prefactor in Eq.(\ref{dmsris}). The monomer friction coefficient is calculated using Eq.(\ref{ciccio}) for unentangled systems and Eq.(\ref{entangly}) for entangled ones. The full lines in Fig.~\ref{FG:diffusion1} show the equations used in the calculation, while the dashed lines represents the prediction of the equation for the entangled system in the unentangled region, and the prediction of the unentangled equation in the entangled region. In the crossover regime, $N\approx 100$ for PE, both expressions lead to very similar results, providing a smooth crossover between the two equations.
The predicted values of the diffusion coefficient appear to be consistent with the known experimental behavior. The diffusion coefficients of unentangled chains ($N < 130$) follow the scaling behavior of the Rouse approach, while the entangled chains show a scaling with degree of polymerization of $-2.5$. Although the latter scaling exponent disagrees with the ``reptation model", it is known that experimental samples of weakly entangled chains also show a scaling exponent of $-2.5$. For those polymer chains, which are just across the transition from unentangled to entangled dynamics, constraint release and "tube" fluctuations are relevant. The observed scaling behavior is also consistent with the scaling of the viscosity observed experimentally.\cite{Lodge} The advantage of our method with respect to UA-MD simulations is that even in the case of long entangled chains it is possible to include a large number of molecules, improving the statistics of calculated correlation functions. Overall this plot shows that it is possible to provide reasonable predictions of large scale dynamical properties by properly rescaling mesoscale simulations.
\begin{figure}
\centering
\includegraphics[scale=0.28]{DofN_new_set_Xrenorm.pdf}
\caption[]{Plot of diffusion coefficients (symbols) as a function of degree of polymerization, $N$. Also shown are the scaling exponents for our unentangled, $N^{-1}$ (dotted line), and entangled systems, $N^{-2.5}$ (dotted line).}
\label{FG:diffusion1}
\end{figure}
\section{Discussion and Conclusions}
\label{SX:da}
The need for developing a fundamental approach to rescale dynamical data obtained from MS simulations of coarse-grained systems has been a long-standing problem from the time that coarse-graining approaches started being developed. Because MS simulations are less computational demanding than atomistic simulations, it is possible to investigate larger systems for longer times than in all-atom simulations, allowing one to extend the maximum time and length scales accessible through simulations and to improve the statistics of measured averaged quantities. Considering that the number of particles in a simulation should be large enough to ensure proximity to the thermodynamic limit, MS simulations of coarse-grained systems could become an indispensable tool to investigate the structure and dynamics of macromolecular liquids.
One advantage of a MS simulation of a coarse-grained system is that the simulation speeds up because of the averaging of the internal degrees of freedom, leading to a softer potential and allowing the study of longer timescales than in a fully atomistic simulation. This implies, however, that the dynamical properties resulting from the MS MD are faster than their real counterpart, for example the ones from UA MD and need to be rescaled.
It is the common procedure to rescale the measured dynamics numerically by bringing a time correlation function to agree with the one measured in atomic level simulations, however we adopt a different strategy. We have proposed a first principle approach to derive an analytical form of the rescaling procedure to be applied to the dynamics measured directly from MS MD of a coarse-grained polymer liquid. Our approach allows for the reliable prediction of the long-time diffusion of a polymer melt as it would be measured in an atomistic or UA-MD simulation. The rescaling procedure has been tested so far against simulations and experiments of polyethylene liquids both unentangled and entangled. Calculated diffusion coefficients for samples for which we do not have data either experimental or simulated, show consistent behavior.
We start by running MS MD simulations of coarse-grained polyethylene melts where each polymer is represented as a point particle. The analytical intermolecular potential, input to the MS MD, is derived from the Ornstein-Zernike equation with the hypernetted closure approximation. The correction term to the measured dynamics of the MS MD simulations, is calculated from the solution of the Generalized Langevin Equations written for the coarse-grained and for the monomer-level representations of the macromolecular liquid. Those equations are formally derived from the Liouville equation by assuming two different lengthscales characterizing the relevant slow dynamics, i.e. monomer and center-of-mass, onto which the Liouville equation is projected.
While the Mori-Zwanzig projection operator technique suggests a reliable criteria to select the proper length scale of coarse-graining for dynamical properties, the GLEs thus generated allow one to derive analytical forms of the rescaling contributions associated with the coarse-grained dynamical equations. The rescaling procedure includes two contributions, given by the changes in entropy and in the friction coefficient during coarse-graining. The entropic contribution emerges from the averaging of the internal degrees of freedom, while the friction is due to the change in shape, and as a consequence the change of the molecular surface exposed to the surrounding molecules. Both corrections depend on the thermodynamic conditions of the system simulated, and on the molecular structure through the radius-of-gyration of the macromolecule. Thermodynamic and molecular quantities enter both directly through the rescaling equations and indirectly through the effective potential in the mesoscale simulations. In this way the dynamics predicted from the rescaling of each mesoscale simulation is specific of the system under study.
A feature of the coarse-graining models we study is the mapping of the polymeric liquid onto simple representations, which are isotropic. At the molecular level the polymer is described as a soft isotropic sphere. At the monomer level, the bead-and spring description affords equivalent beads in the chain, which is a reliable approximation due to the high number of statistically equivalent structural configuration of the molecule. Chain end effects enter in the model through the finite size of the polymer in the matrix representation of the equations. Moreover, because the monomers in a homopolymer are structurally identical, with the exception of the two end monomers, the intermolecular monomer-monomer hard-core interaction potential is assumed to be identical for any pair of monomers, and each monomer is supposed to have identical friction coefficient.
Although the theoretical picture is straightforward, our approach has the advantage of being described in closed-form expressions, even if approximated, which allows for an analytical solution of the rescaling formalism. This has the potential of being useful in improving our understanding of the nature of coarse-graining procedures.
\section*{APPENDIX I: INTERNAL ENERGY CALCULATION FOR A FREELY-ROTATING-CHAIN MODEL}
\label{Appendix1}
The effective mean-force potential for one homopolymer composed of $N$ monomers can be expressed through the structural matrix $\mathbf{A}$ as
\begin{eqnarray}
U = \frac{3k_B T}{2 l^2} \sum_{i,j} \mathbf{A}_{i,j}\, \mathbf{r}_i \cdot \mathbf{r}_j& = &\frac{3k_B T}{2 l^2} \sum_{x,y,z} \sum_{i,j} x_i \mathbf{A}_{i,j} x_j \nonumber \\
&=& \frac{3 k_B T}{2\, l^2} \mathbf{r}^T \mathbf{A} \mathbf{r} \, ,
\end{eqnarray}
with the matrix $\mathbf{A}$ being real and symmetric, and diagonalized by the orthonormal matrix of the eigenvectors $\mathbf{Q}^{-1}=\mathbf{Q}^T$, so that
\begin{equation}
\mathbf{r}^T \mathbf{A} \mathbf{r} = \mathbf{\xi}^T \mathbf{Q}^{-1} \mathbf{A} \mathbf{Q} \mathbf{\xi} = \mathbf{\xi}^T \mathbf{\Lambda} \mathbf{\xi} \, ,
\end{equation}
where $\mathbf{\Lambda}$ is the matrix of the eigenvalues, and $\xi$ is the matrix of the normal modes defined by $\mathbf{r}=\mathbf{Q} \mathbf{\xi}$.
In this model, the equilibrium distribution function is
\begin{eqnarray}
\Psi_{eq}(\textbf{r}) &=& \textit{N}_x e^{-\frac{3}{2 l^2} x^T A x} \textit{N}_y e^{-\frac{3}{2 l^2} y^T A y} \textit{N}_z e^{-\frac{3}{2 l^2} z^T A z} \nonumber \\
&=& \textit{N} e^{-\frac{3}{2\, l^2} \mathbf{r}^T \mathbf{A}\, \mathbf{r}} = \textit{N} e^{-\frac{1}{2} \mathbf{r}^T \mathbf{A'}\, \mathbf{r}} \, ,
\label{EQ:distribution}
\end{eqnarray}
where for convenience of notation we introduced the matrix $\mathbf{A'}=3 \mathbf{A}/l^2$.
Here $\textit{N}_x$ is the normalization factor, defined by enforcing $\int dx \Psi_x =1$, as
\begin{equation}
N_x = \left(\frac{3}{2 \pi l^2} \right) ^{N/2} [det(\mathbf{A})]^{1/2} \ ,
\label{EQ:normalization}
\end{equation}
with $\textit{N}_x=\textit{N}_y=\textit{N}_z=\textit{N}^{1/3}$.
The statistically averaged internal energy for one molecule consisting of $N$ monomers simplifies to
\begin{equation}
\left< \frac{E}{k_B T} \right> = N \int U e^{-\frac{1}{2} \mathbf{r}^T \mathbf{A'} \mathbf{r}} d\, r = N \int \frac{1}{2} \mathbf{r}^T \mathbf{A'} \mathbf{r} e^{-\frac{1}{2} \mathbf{r}^T \mathbf{A'} \mathbf{r}} \, .
\end{equation}
In one dimension,
\begin{eqnarray}
\left< \frac{E}{k_B T} \right>_x
& = &\frac{3 N_x}{2l^2} \int d\mathbf{x}\, \, \mathbf{x}^T \mathbf{A} \mathbf{x}\, e^{-\frac{3}{2l^2} \mathbf{x}^T \mathbf{A} \mathbf{x}}\nonumber \\
&=& N_x \frac{3}{2 l^2} \left[ \frac{N_x l^2}{3} \prod_{i=1}^{N_x} \sqrt{\frac{2 \pi l^2}{3 \lambda_i}} \right]=\frac{N}{2} \ ,
\end{eqnarray}
which gives, as the final result for the internal energy of one molecule consisting of $N$ monomers,
\begin{equation}
\left< \frac{E}{k_B T} \right> = \frac{3 N}{2} \ .
\end{equation}
\section*{APPENDIX II: THE DYNAMIC MEMORY FUNCITION}
\label{Appendix2}
We briefly report here the derivation of Eq.(\ref{EQ:MFMON}) starting from Eq.(\ref{cJ}).
The product of the direct and projected forces is expressed as a function of the density field variables as
\begin{eqnarray}
\bigl< \mathbf{F}(0) \cdot \mathbf{F}^{\hat Q}(t)\bigr> \cong \mathbf{F}(\mathbf{r}) \cdot \mathbf{F}(\mathbf{r}')
\, \bigl<\rho_\alpha (\textbf{r};0)\rho_\gamma (\textbf{r};t)\bigr> \ .
\end{eqnarray}
Because the fluid is uniform and isotropic, the density fields can be replaced by their fluctuation variables, $\Delta \rho_\alpha (\mathbf{r},t)=\rho_\alpha (\mathbf{r},t)-\bigl<\rho_\alpha (\mathbf{r})\bigr>$, where the ensemble-averaged density field is approximated by $\bigl<\rho_\alpha (\mathbf{r})\bigr>\approx \rho g(r)$.
The correlation of the random forces is then expressed as
\begin{eqnarray}
\label{uhu}
\bigl< \mathbf{F}(0) \cdot \mathbf{F}^{\hat Q}(t)\bigr> &\cong& {\hat r} \cdot {\hat r'} \rho^2 g(r)g(r') F(r) F(r') \nonumber \\
&&\times \frac{\bigl<\Delta\rho_\alpha (\mathbf{r})\Delta\rho_\gamma (\mathbf{r}',t)\bigr>}{\bigl<\rho_\alpha (r)\bigr>\bigl<\rho_\gamma (r')\bigr>} \ ,
\end{eqnarray}
where we adopt a kind of "dynamical" Kirkwood superposition approximation in a weighted average form
\begin{eqnarray}
\label{aha}
\bigl<\Delta\rho_\alpha (\mathbf{r})\Delta\rho_\gamma (\mathbf{r}',t)\bigr> &\approx& \rho \int d \mathbf{R} \ g(r) g(r') S(R,t)\nonumber \\
&&\times S(|\mathbf{r}- \mathbf{r}' + \mathbf{R}|t) \ .
\end{eqnarray}
Eq.(\ref{aha}) describes the multipoint correlation between the density fluctuations at a distance $\mathbf{r}$ from segment $\alpha$ at time zero, and the density fluctuations a distance $\mathbf{r}'$ from segment $\gamma$ at time $t$. Because $\alpha$ and $\gamma$ can be on the same or on different polymer chains, no assumptions are made a priori about the relative importance of intra and intermolecular correlations. In this way, the chain connectivity does not play a dominant role in our description from the very beginning. We then calculate both intra and intermolecular contributions and show that both need to be included in the calculation of the memory function as intramolecular contributions are comparable in size to the intermolecular ones.
Substitution of Eq.(\ref{aha}) into Eq.(\ref{uhu}) leads to Eq.(\ref{EQ:MFMON}).
\section*{ACKNOWLEDGEMENTS}
We acknowledge support from the National Science Foundation through grant DMR-0804145.
Computational resources were provided by LONI through the TeraGrid project supported by NSF.
We thank James McCarty for his careful reading of the manuscript.
We are grateful to G. S. Grest, V. G. Mavrantzas and coworkers for sharing the UA-MD computer simulation trajectories.
|
\section{Introduction
The problem of heat conductivity in low dimensional systems
attracts much attention in last decades (see review \cite{Lep03})
and is motivated by the discovery of quasi-one-dimensional
(nanotubes, nanowires, etc.) and two-dimensional (graphen,
graphan, etc.) systems.
The modern theory of heat conductivity was initiated by the
celebrated preprint of E.~Fermi, J.~Pasta, and S.~Ulam
\cite{Fer55}, though the primary aim was ``{\it of establishing,
experimentally, the rate of approaching to the equipartition of
energy among the various degrees of freedom}''. Subsequent
investigations demonstrated wide area of consequences in many
physical and mathematical phenomena (see reviews in special issues
of journals CHAOS \cite{Cha05} and Lecture Notes in Physics
\cite{Lec08} devoted to the 50th anniversary of the FPU preprint).
The dynamical properties of nonlinear systems in microcanonical
ensemble (total energy $E =$ const) were thoroughly analyzed. It
allows to investigate the dynamics and to get exact results
(soliton \cite{Kru64, Zab65, Dod82} and breather \cite{Cam96,
Sie88, Dau93, Aub94, Mac94} solutions), to analyze regular and
stochastic regimes and to find the corresponding thresholds. The
FPU preprint also initiated the investigations in the field of
``experimental mathematics'' \cite{Por09} .
About ten decades ago P.~Debye argued that the nonlinearity can be
responsible for the finite value of heat conductivity in
insulating materials \cite{Deb14}. But modern analysis shows that
it is not always the case. There are many examples where the
coefficient of heat conductivity $\varkappa$ diverges with the
increasing system size $L$ as $\varkappa \propto L^{\alpha}$ where
$\alpha > 0$, and $\varkappa \to \infty$ in the thermodynamic
limit ($L \to \infty$). Most of momentum conserving
one-dimensional nonlinear lattices with various types of
nearest-neighbor interactions have this unusual property (see,
e.g., \cite{Lep03, Cas05, Lic08}\,). Moreover, some other systems,
-- two- \cite{Bar07, Lip00, Sai10} and three-dimensional lattices
\cite{Shi08}, polyethylene chain \cite{Hen09}, carbon nanotubes
\cite{Yao05, Yu_05, Mar02, Min05, Cao04} have analogous property
-- diverging heat conductivity with the increasing size of the
system.
There were some conjectures explaining the anomalous heat
conductivity. Generally speaking, whenever the equilibrium
dynamics of a lattice can be decomposed into that of independent
``modes'' or quasi-particles, the system is expected to behave as
an ideal thermal conductor \cite{Lep05}. Thereby, the existence of
stable nonlinear excitations is expected to yield ballistic rather
than diffusive transport. At low temperatures normal modes are
phonons. At higher temperatures noninteracting ``gas'' of solitons
or/and breathers starts to play more significant role, and M.~Toda
was the first who suggested the possibility of heat transport by
solitons \cite{Tod79}.
Though analytical expressions for solitons can be derived only for
few continuum models described by partial differential equations,
Friesecke and Pego in a series of recent papers \cite{Fri99,
Fri02, Fri04a, Fri04b} made a detailed study of the existence and
stability of solitary wave solutions on discrete lattices with the
Hamiltonian $H = \sum_i \frac12 p_i^2 + u(y_i)$, where $y_i = x_i
- x_{i-1}; \ p_i = \dot x_i$. It has been proven that the systems
with this Hamiltonian and with the following generic properties of
nearest-neighbor interactions: $u'(0) = 0; \ u''(0) > 0; \ u'''(0)
\neq 0$ has a family of solitary wave solutions which in the small
amplitude, long-wavelength limit have a profile close to that of
the KdV soliton. It was also shown \cite{Hof08} that these
solutions are asymptotically stable. Thus most acceptable point of
view on the origin of anomalous heat conductivity in nonlinear
lattices is as follows: phonons are responsible for heat
conductivity at low temperatures, and at high temperatures --
solitons \cite{Li_05,Vil02}.
Rather confusing experimental and numerical results are
demonstrated in literature about the dependence of heat
conductivity on different parameters, -- model under
consideration, types of boundary conditions, used thermostat and
temperature. For instance, temperature dependence of heat
conductivity in carbon nanotubes decreases as $\varkappa \sim 1/T$
at $T > 10$ K \cite{Mar03}; experimentally is found \cite{Yu_05}
that $\varkappa$ also decreases with the growth of temperature.
Different temperature dependencies $\varkappa$ vs. $T$ were found
in 1D nonlinear lattices. For $\beta$-FPU lattice: $ \varkappa
\sim N^{\alpha} T^{-1}$ at $T \lesssim 0.1$ and $ \varkappa \sim
N^{\alpha} T^{1/4}$ at $T > 50$ \cite{Aok01} what is usually
observed in insulating crystals. For the interparticle harmonic
potentials and on-site potentials (e.g. Klein-Gordon chains)
$\varkappa \sim T^{-1.35}$, i.e. heat conductivity decreases with
the growth of temperature \cite{Aok00}. But there exists firm
theoretical background \cite{Nar02} that the exponent $\alpha$ in
the dependence $\varkappa \propto N^{\alpha}$ is the universal
constant $\alpha \approx 1/3$ in momentum-conserving systems.
The calculation of heat conductivity at small temperature
gradients is an additional problem. Usually these calculations are
very time consuming because of great fluctuations of heat current
and statistical averaging over large number of MD trajectories is
necessary.
The paper organized as follows. In section \ref{sec:heat} the heat
conductivity is considered when the temperature gradient $\nabla
T$ is small. The explicit contribution to the heat conductivity is
extracted by the decomposing of the total process into two parts.
The first one is the equilibrium process at temperatures $T$ of
both lattice ends, and the second -- non-equilibrium, when one
lattice end has temperature $\Delta T$ and the other -- zero
temperature. Namely the latter process is responsible for the heat
conductivity. This method allows to find the threshold temperature
$T_{\rm thr}$. And though $T_{\rm thr}$ scales with the lattice
length $N$ as $T_{\rm thr} \sim N^{-3}$, unusual dynamics can be
revealed on nanoscale when both $T$ and $N$ are small. The low
temperature thermodynamics can be adequately described in terms of
harmonic lattice with temperature renormalized rigidity
coefficients.
Section \ref{sec:sol_breath} is independent of the previous one
and is aimed at elucidating the role of solitons and breathers in
the heat conductivity. Solitons and breathers are found as the
solutions of the modified Korteweg -- de~Vries equation. This
equation is obtained in the continuum limit from the discrete
$\beta$-FPU lattice. Soliton and breather solution are checked in
numerical simulation and demonstrate very high stability.
Necessary details of the derivation of accurate soliton and
breather solutions in the continuum limit are given in Appendix.
\section{Heat conductivity in the $\beta$-FPU lattice}
\label{sec:heat}
We consider the one-dimensional $\beta$-FPU lattice of $N$
oscillators with the interaction of nearest neighbors
\begin{equation}
\label{2-1}
U = \sum\limits_i \dfrac{\alpha}{2} (x_i - x_{i-1})^2 +
\dfrac{\beta}{4} (x_i - x_{i-1})^4
\end{equation}
(usually the dimensionless potential will be used below, that is
$\alpha = \beta = m = 1$). Nonequilibrium conditions are necessary
for the heat transport simulation. The most abundant method is the
placement of the lattice into the heat bath with different
temperatures of left $T_+$ and right $T_-$ ends ($T_+ > T_-$).
Different types of heat reservoirs are thoroughly analyzed in
\cite{Lep03}. The usage of the Langevin forces with the noise
terms and friction forces acting on the left $F_+ = \xi_+ - \gamma
\dot x_1$ and right $F_- = \xi_- \gamma \dot x_N$ oscillators is
the common practice ($\gamma = 1$ is also put for brevity).
$\left\{ \xi_{\pm} \right\}$ are independent Wiener processes with
zero mean and the correlator $\left< \xi_{\pm}(t_1) \,
\xi_{\pm}(t_2) \right> = 2 T_{\pm} \, \delta(t_1 - t_2)$. $\Delta
T = (T_+ - T_-)$ is the temperature difference. The generalized
Langevin dynamics with a memory kernel and colored noises is also
suggested \cite{Wan07} to correctly account for the effect of the
heat baths.
The following set of stochastic differential equations (SDEs)
\begin{equation}
\label{2-2}
\ddot x _i = - \dfrac{\partial U}{\partial x_i} + \delta_{i1}
F_+ + \delta_{iN} F_-
\end{equation}
is usually solved to find the heat flux $J$. And the local heat
flux (power transmitted from $i$th to $(i+1)$th oscillator) is
\cite{Zha02}
\begin{equation}
\label{2-4}
J_{i \to i+1} = F_{i \to i+1}\, \dot x_{i+1} ;
\qquad F_{i \to i+1} \equiv -U'(x_{i+1} - x_i),
\end{equation}
where $F_{i \to i+1}$ is a shorthand notation for the force
exerted by the $i$th on the $(i+1)$th oscillator. The total heat
flux $J$ can be found as the mean $J = (N-1)^{-1} \sum_i^{N-1}
J_{i \to i+1}$.
\subsection{Equilibrium and non-equilibrium contributions
to the heat conductivity}
If $T_- \neq 0$ then the process of heat conductivity can be
formally decomposed into two contributions: the first one --
equilibrium process with equal temperatures $T_-$ of both lattice
ends; and the second -- nonequilibrium process with temperature
$\Delta T$ of the left lattice end and zero temperature of the
right end (see Fig.~\ref{fig_01}) (by `process' we hereafter
assume for brevity the solution $ {\bf x} (t) = \left\{ x_1(t),
x_2(t), \ldots , x_N(t) \right\}; \, {\bf v} (t) = \left\{
v_1(t), v_2(t), \ldots , v_N(t) \right\} $ of the corresponding
SDEs).
\begin{figure}
\begin{center}
\includegraphics[width=100 mm,angle=0]{fig_01.eps}
\end{center}
\caption{
\label{fig_01}
Schematic representation of the total
process ${\bf x}(t)$ as sum of equilibrium ${\bf x}^0(t)$ and
non-equilibrium ${\bf x}^1(t)$ processes.
}
\end{figure}
Namely the second process is responsible for the heat transport
taking place on(?) the background of equilibrium process. Once
this approach is utilized then the noise terms in \eqref{2-2},
owing to their independence, are $\left\{ \xi_+ \right\} = \left\{
\xi^0 \right\} + \left\{ \xi^1 \right\} $ and $ \left\{ \xi_-
\right\} = \left\{ \xi^0 \right\} $ for the left and right
lattice ends, correspondingly Superscripts `0' and `1' refer to
equilibrium and nonequilibrium processes. The total process ${\bf
x}(t) $ can be represented as the sum
\begin{equation}
\label{2-5}
{\bf x}(t) = {\bf x}^0(t) + {\bf x}^1(t),
\end{equation}
where ${\bf x}^0(t)$ is the equilibrium (Gibbs's) process at
temperature $T_-$, and ${\bf x}^1(t)$ -- nonequilibrium,
responsible for the energy transport, process. The corresponding
stochastic dynamics is
\begin{equation}
\label{2-6}
\hspace{-1.3 cm} \ddot x_i^0 = - \dfrac{\partial U^0}{\partial x_i^0} +
\delta_{i1} (\xi^0 - \dot x^0_1) +
\delta_{iN} (\xi^0 - \dot x^0_{N}),
\end{equation}
\begin{equation}
\label{2-7}
\ddot x_i^1 = - \left[
\dfrac{\partial U}{\partial x_i} -
\dfrac{\partial U^0}{\partial x_i^0}
\right] +
\delta_{i1} (\xi^1 - \dot x^1_1) +
\delta_{iN} (- \dot x^1_N),
\end{equation}
and the sum of equations \eqref{2-6} and \eqref{2-7} is identical
to the parent equation \eqref{2-2}. Random values $\left\{ \xi^0
\right\}$ and $\left\{ \xi^1 \right\}$ obey the identities $\left<
\xi^0(t_1) \xi^0(t_1) \right> = 2 T_- \delta(t_1 - t_2)$ and
$\left< \xi^1(t_1) \xi^1(t_1) \right> = 2 \, \Delta T \delta(t_1 -
t_2)$; $U^0$ is the total energy \eqref{2-1} where the arguments $
x_1(t), x_2(t), \ldots, x_N(t) $ of the total process are replaced
by coordinates of the equilibrium process $x_1^0(t), x_2^0(t),
\ldots, x_N^0(t)$. Expression in the square brackets in
\eqref{2-7} is the difference of forces acting on the $i$th
oscillator from the total process ${\bf x}(t)$ and equilibrium
process ${\bf x}^0(t)$. It is worth mentioning that this force is
the random value, and the process ${\bf x}^1(t)$ (heat transport)
is realized in the lattice with {\it time-dependent random
potentials}. The problem of heat conductivity in the random
time-independent potentials was analyzed in \cite{Joh08}.
Equation \eqref{2-6} describes the system embedded in the heat
reservoir at temperature $T_-$ of both lattice ends. And ${\bf
x}^0(t)$ is the stationary equilibrium process described by the
canonical Gibbs distribution. Process ${\bf x}^1(t)$ is
responsible for the heat transport and the Wiener process
$\left\{ \xi^1(t) \right\}$ on the left lattice end defines
temperature $\Delta T$. Right lattice end has zero temperature
(only the friction force acts on this oscillator). An expression
for the local heat flux is
\begin{equation}
\label{2-8}
J_{i \to i+1} =
\left[ F_{i \to i+1}({\bf x}) - F_{i \to i+1}({\bf x}^0) \right]\, \dot
x_{i+1}^0 ,
\end{equation}
and the equilibrium process ${\bf x}^0$ does not transfer energy:
$\left< F_{i \to i+1}({\bf x}^0) \, \dot x^0_{i+1} \right> \equiv
0$, where $\left< \ldots \right>$ stands for the time average. It
is essential that the heat flux \eqref{2-8} is the small
difference of large values from processes ${\bf x}(t)$ and ${\bf
x}^0(t)$. This is the reason why MD simulation gives large
fluctuation when $\Delta T \to 0$ and $T$ is not low. It can be
shown that the time of computation increases $\propto (\Delta
T)^{-2}$ if the standard error is fixed. The comparison of two
approaches (solving of standard SDEs \eqref{2-2} and
\eqref{2-6}-\eqref{2-7}) is shown in Fig.~\ref{fig_02} and results
coincide with very good accuracy.
\begin{figure}
\begin{center}
\includegraphics[width=80mm,angle=0]{fig_02.eps}
\end{center}
\caption{
\label{fig_02}
Temperature dependence of heat conductivity for the lattice of $N=5$
oscillators. Filled circles: solution of standard SDEs \eqref{2-2};
empty circles: SDEs
\eqref{2-6}-\eqref{2-7}. Averaging over $100$ MD trajectories
$10^4$ time units (t.u.) each. $T_- = 0{.}2$,
$\Delta T = 0{.}01 T_-$. Triangle up at
$T = 0$ is the exact value in the harmonic approximation ($\beta = 0$).
}
\end{figure}
Some results in this paper are obtained for the number of
oscillators $N = 5$. It may appear that this value is too small.
For instance, ``standard'' simulations require up to $\sim 10^4$
particles and $\sim 10^8$ integration steps plus ensemble
averaging \cite{Lep03}. But our results are aimed at establishing
some new issues where number of particles is unessential. Lattices
with larger number of oscillators were tested when necessary.
Relative displacements $(x_i^0 - x_{i-1}^0)$ and $(x_i^1 -
x_{i-1}^1)$ for processes ${\bf x}^0(t)$ and ${\bf x}^1(t)$ are
shown in Fig.~3. These values characterize the energy fluxes in
the lattice. Energy fluxes to the left and to the right are equal
on average for process ${\bf x}^0(t)$ as it is the equilibrium
process without energy transfer. But the energy flux is directed
mainly to the right for process ${\bf x}^1(t)$ (right panel). Low
temperature is chosen for the better illustration.
\begin{figure}
\begin{center}
\includegraphics[width=60mm,angle=0]{fig_03a_color.eps}
\includegraphics[width=68mm,angle=0]{fig_03b_color.eps}
\end{center}
\caption{
\label{fig_3}
Spatiotemporal evolutions of relative displacements $(x_i^0 - x_{i-1}^0)$
and $(x_i^1 - x_{i-1}^1)$ for the equilibrium
${\bf x}^0(t)$ (left panel) and nonequilibrium ${\bf x}^1(t)$ (right panel) processes.
$N = 100$, $T_- = 0{.}05$, $\Delta T = 0{.}0001$.
}
\end{figure}
The dependence of heat conductivity on the oscillators number $N$
is shown in Fig.~\ref{fig_4} at two value of temperature $T_-$.
Results coincide with very good accuracy. Inharmonicity becomes
negligible at low temperature and heat conductivity at $T_- =
0{.}1$ (cirles in Fig.~\ref{fig_4}) coincides with the heat
conductivity of the harmonic lattice (dashed line) with good
accuracy. The analytical solution of the heat conductivity for the
harmonic lattice is given in \cite{Rie67}.
\begin{figure}
\begin{center}
\includegraphics[width=80mm,angle=0]{fig_04.eps}
\end{center}
\caption{
\label{fig_4}
Coefficient of heat conductivity for the $\beta$-FPU lattice for
$N = 7-150$ oscillators. Squares: $T_-=1$, circles:
$T_-=0{.}1$. Filled symbols -- results obtained by the solution
of standard SDEs \eqref{2-2}, empty symbols -- SDEs
\eqref{2-6}-\eqref{2-7}. Averaging over 200 MD trajectories
$3\,10^4$ t.u. $\Delta T = 0{.}01T_-$. Dashed line -- harmonic
approximation. Filled symbols are practically fully covered by
empty symbols and are invisible.
}
\end{figure}
There should be solved twice as large SDEs \eqref{2-6}-\eqref{2-7}
in suggested approach as that in standard scheme \eqref{2-2}. But
this approach has some undoubted merits as discussed below.
\subsection{Heat conductivity at small temperature gradients}
One of the goals of the present paper is the computation of heat
conductivity at very small temperature gradients. With this in
mind an expression for the heat flux is analyzed in more details.
The expression for the local heat flux \eqref{2-8} can be
rewritten as
\begin{equation}
\label{2-9}
\begin{split}
J_{i \to i+1} = & \left[ F_{i \to i+1}({\bf x}) -
F_{i \to i+1}({\bf x}^0) \right]
\dot x_{i+1}^1 = \\
{} & \left[ -(x_{i+1} - x_i) -(x_{i+1} - x_i)^3 +
(x^0_{i+1} - x^0_i) + (x^0_{i+1} - x^0_i)^3 \right]
\dot x_{i+1}^1 ,
\end{split}
\end{equation}
where $\dot x_{i+1}^1$ -- velocity of $(i+1)$th oscillator in
process ${\bf x}^1$. Taking in mind that $x_i = x_i^0 + x_i^1$,
the expression \eqref{2-9} can be transformed to
\begin{equation}
\label{2-90}
J_{i \to i+1} = \left\{ - (x_{i+1}^1 - x_i^1)
\left[ 1 + 3 \, (x_{i+1}^0 - x_i^0)^2
\right] +
(x_{i+1}^1 - x_i^1)^2
\left[1 + 3 \, (x_{i+1}^0 - x_i^0)
\right]
\right\} \dot x_{i+1}^1 .
\end{equation}
Processes ${\bf x}^0(t)$ and ${\bf x}^1(t)$ have different ranges
of specific energies. Noise terms $\left\{ \xi^1 \right\}$, which
provide temperature $\Delta T$, are of the order $ \xi^1 \sim
\sqrt{\Delta T}$ (as $\left< \xi^1(t_1) \xi^1(t_2) \right> \sim
\Delta T$). And one can expect that process ${\bf x}^1(t)$ has the
same order ${\bf x}^1(t) \sim \sqrt{\Delta T}$ because equations
\eqref{2-7} become linear in the limit $\Delta T \to 0$ when
$\xi^1 \to 0$. Expression in curly brackets in \eqref{2-90} is the
polynomial of the third degree in the square root of temperature
difference $\sqrt{\Delta T}$. Taking into account that the
velocity $\dot x_{i+1}^1$ is also of the order $\sim \sqrt{\Delta
T}$, expression \eqref{2-90} is the polynomial of the forth degree
in $\sqrt{\Delta T}$. But the coefficient of heat conductivity is
determined by the relation $J/\Delta T$. Then terms of the third
and forth orders can be neglected at $\Delta T \to 0$. Then
\eqref{2-90} is simplified to
\begin{equation}
\label{2-91}
J_{i \to i+1} = - \left( x_{i+1}^1 - x_i^1 \right)
\left[ 1 + 3 \, \left( x_{i+1}^0 - x_i^0 \right)^2 \right]
\dot x_{i+1}^1 .
\end{equation}
An expression for the potential energy, corresponding to process
${\bf x}^1(t)$, can be derived analogously. This energy is the
difference of potential energies $U({\bf x}) - U^0({\bf x}^0)$ and
again, using coordinates ${\bf x}$ and ${\bf x}^0$, and preserving
only terms quadratic in $(x_{i+1}^1 - x_{i}^1)$, one can get the
potential energy for process ${\bf x}^1$ in the form
\begin{equation}
\label{2-11}
U^1 = \dfrac12 \sum_i g_{i+1}(t) \, (x_{i+1}^1 - x_{i}^1)^2,
\qquad
g_{i+1}(t) = 1 + 3 \, [x_{i+1}^0(t) - x_{i}^0(t)]^2 \,,
\end{equation}
where $g_i(t)$ are {\it time-dependent random} coefficients of
rigidities determined by process ${\bf x}^0(t)$.
It is illuminating to note that in the limit $\Delta T \to 0$ the
problem of heat conductivity in the $\beta$-FPU lattice is reduced
to the {\it harmonic} lattice with random coefficients.
Corresponding SDEs have noise terms with friction forces on the
left oscillator and zero temperature (only viscous forces) on the
right oscillator:
\begin{equation}
\label{2-12}
\ddot x_i^1 = - g_i (x_i^1 - x_{i-1}^1) + g_{i+1} (x_{i+1}^1 - x_{i}^1) +
\delta_{i1} (\xi^1 - \dot x_1^1) - \delta_{i N} \dot x_{N}^1
\end{equation}
and $g_i, g_{i+1}$ are defined in \eqref{2-11}. If the 1D lattice
with an arbitrary interaction potential is analyzed then the
corresponding equations are the same with rigidities
$g_i=U''(x^0_i - x^0_{i-1}$) where $U$ is the potential energy.
SDE for the system with arbitrary neighbor radius of interaction
can be written in the general form as
\begin{equation}
\label{2-13}
\ddot x_i^1 = - \sum_{j=1}^M {\bf \Lambda}_{ij}^0 \, x_j^1 +
\delta_{i1} (\xi^1 - \dot x_1^1) - \delta_{i N} \dot x_{N}^1
\,,
\end{equation}
where ${\bf \Lambda}_{ij}^0$ -- matrix of second derivatives of
potential energy depending on ${\bf x}^0$, and $M$ is the number
of neighbors.
\subsection{Unusual dynamics of process ${\bf x}^1(t)$ at high
temperatures}
The process ${\bf x}^1(t)$ can be characterized by some time
average correlators. The correlator $\left< \left[ x_1^1(t)
\right]^2 \right>$ was analyzed for the better understanding of
the dynamics of process ${\bf x}^1(t)$. One can expect that
$\left< \left[ x_1^1(t) \right]^2 \right> \sim \Delta T$ as
discussed above. Two temperatures of the background process $T_-$
were tested: $T_1 = 0{.}2$ and $T_2 = 5$ (hereafter subindex `--'
is omitted, that is $T_- = T$). Results are shown in
Fig.~\ref{fig_5}.
As expected, the correlator $\left< \, [x_1^1(t)]^2 \right>$
linearly depends on $\Delta T$: $\left< \, [x_1^1(t)]^2 \right>
\sim \Delta T$ at $T_1 = 0{.}2$. But the case is quite different
at $T_2 = 5$: $\left< \, [x_1^1(t)]^2 \right>$ reaches the
stationary value $\approx 0{.}064$ in the limit $\Delta T \to 0$.
It means that there exists some undamped stationary process ${\bf
x}^1(t)$ at high temperatures $T$ of the background process ${\bf
x}^0(t)$ even in the limit $\Delta T \to 0$. These results also
imply an existence of a threshold temperature $T_{\rm thr}$
separating two regimes -- damped at low temperatures and undamped
at high temperatures.
\begin{figure}
\begin{center}
\includegraphics[width=90mm,angle=0]{fig_05.eps}
\end{center}
\caption{
\label{fig_5}
Dependence of correlator $\left< (x_1^1)^2 \right>$ on the
temperature difference $\Delta T$. Filled circles: $T = 0{.}2$,
empty circles: $T=5$. Asymptotic value $\left< (x_1^1)^2
\right>_{\Delta T \to 0} \approx 0{.}064$ at $T = 5$ (coefficient
of linear regression $0{.}9993$). Averaging over 100 MD
trajectories $10^4$ t.u. each. $N=5$. The range of $\Delta T$:
$10^{-11} \leq \Delta T/T \leq 2 \! \cdot \! 10^{-1}$.
}
\end{figure}
\subsection{Threshold temperature}
Any process ${\bf x}^1(t)$ damps out at low temperatures and
flattens out to a stationary value at higher temperatures even in
the limit $\Delta T \to 0$, and the temperature $T$ of process
${\bf x}^0(t)$ defines different damping rates. Bearing this in
mind, it is convenient to excite some auxiliary process
$\widetilde{\bf x}^1(t)$ over the background process ${\bf
x}^0(t)$ and to analyze it.
Coordinates and velocities of process $\widetilde{\bf x}^1(t)$ get
random increments $\frac12 \sum_i [{\widetilde x}_i^1(t=0)]^2 +
\frac12 \sum_i [ {\widetilde v}_i^1(t=0)]^2 = 0{.}5$. The
particular choice of initial conditions does not influence the
final results. The total dynamics is the sum of two processes
${\bf x}(t) = {\bf x}^0(t) + \widetilde{\bf x}^1(t)$.
Stochastic differential equations for the process ${\bf \widetilde
x}^1(t)$ are
\begin{equation}
\label{2-10}
\ddot {\widetilde x}_i^1 = - \left[
\dfrac{\partial U } {\partial x_i} -
\dfrac{\partial U^0 }{\partial x_i^0}
\right] -
\delta_{i1}\dot {\widetilde x}^1_1 -
\delta_{iN} \dot {\widetilde x}^1_{N}
\end{equation}
and only viscous forces acts at the extreme left and right
oscillators. $U$ and $U^0$ are potential energies with coordinates
${\bf x}(t)$ and ${\bf x}^0(t)$, correspondingly. Stochastic
dynamics \eqref{2-10} is implicitly ruled out by the temperature
$T$ of process ${\bf x}^0(t)$.
To find the threshold temperature we initially consider the case
of small temperature $T$ when process ${\bf \widetilde x}^1(t)$ is
damped out. Its damping is determined by the viscous friction of
left $(-\dot {\widetilde x}^1_1)$ and right $(-\dot {\widetilde
x}^1_{N})$ oscillators in \eqref{2-10}. Gradually increasing the
temperature its threshold value can be found when process ${\bf
\widetilde x}^1(t)$ becomes undamped. The damping of mean squared
displacement of the first oscillator $[{\widetilde x}_1^1(t)]^2$
was calculated. Process $\widetilde{\bf x}^1(t)$ exponentially
decays $\left< \, [\widetilde x_1^1(t)]^2 \right> \propto
\exp(-\alpha t)$ and $\alpha$ depends on $T$ (see
Fig.~\ref{fig_6}a).
\begin{figure}
!\begin{center}
\includegraphics[width = 75mm, angle=0]{fig_06a.eps}
\includegraphics[width = 70mm, angle=0]{fig_06b.eps}
!\end{center}
\caption{
\label{fig_6}
a) Exponential damping of process ${\bf \widetilde x}^1(t)$ at
different temperatures: $T=3{.}5$ (circles), $T=3{.}8$ (squares),
$T=4{.}0$ (triangles up), $T=4{.}2$ (triangles down). Solid
lines -- linear regressions. Averaging time $\sim \! 5\,000-10\,000$
t.u. 20 trajectories ${\bf x}^0$ were used to estimate the standard
error. b) Damping coefficient $(-\alpha)$ as the function of temperature
$T$ of process ${\bf x}^0(t)$. Damping stops $(\alpha = 0)$ at
$T_{\rm thr} \simeq 4{.}07$. $N=5$.
}
\end{figure}
The temperature dependence of coefficient $\alpha$ is shown in
Fig.~\ref{fig_6}b and $T_{\rm thr} \simeq 4{.}07$ when $\alpha =
0$.
Next method to find the threshold temperature is moving ``from up
to down'', going from higher to lower temperatures. At high
temperatures there exists the stationary process arising from
random forces $\Phi_i = \left[ {\partial U }/ {\partial x_i} -
{\partial U^0 }/{\partial x_i^0} \right]$ (see~\eqref{2-10}). And
process ${\bf \widetilde x}^1(t)$ decreases in a sense that all
quadratic mean values tend to zero as temperatures approaches
$T_{\rm thr}$ from above. When the temperature reaches its
threshold value, process ${\bf \tilde x}^1(t)$ is totally damped
(see Fig.~\ref{fig_7}). The found threshold temperature is $T_{\rm
thr} \simeq 4{.}09$.
\begin{figure}
\begin{center}
\includegraphics[width=80mm,angle=0]{fig_07.eps}
\end{center}
\caption{
\label{fig_7}
Stationary values $\left< (\widetilde x_1^1)^2 \right>$ at
$T > T_{\rm thr}$.
Time of averaging $~10^6$ t.u. The temperature dependence is
approximated by the function
$\left< [\widetilde x_1^1(t)]^2 \right>
\sim \exp[-b /(T-T_{\rm thr})]$ (solid line). $N=5$.
}
\end{figure}
Strange behavior of process ${\bf x}^1(t)$ is basically explained
by {\it time-dependent random} forces $\Phi_i = \left[ {\partial
U}/{\partial x_i} - {\partial U^0}/{\partial x_i^0} \right]$ (see
\eqref{2-7}) rather then random Langevin forces $\xi^1 \sim \sqrt{
\Delta T}$. And the plateau for the correlator $\left<
[x^1_1(t)]^2 \right> \approx 0{.}064$ at $\Delta T \to 0$ is
determined exclusively by the background process ${\bf x}^0$.
This conjecture can be additionally supported. Let us consider the
case of low temperatures $T$ when process ${\bf x}^0$ is ``weak''.
Then the rigidity coefficients $g_i$ in \eqref{2-11} are close to
unity. The lattice where actual rigidity coefficients $g_i$
\eqref{2-11} are substituted by the mean values taken from the
equilibrium Gibbs distribution: $\left< g_i \right> = g_0(T)$ and
$g_0(T) = 1 + 3\left< (x^0_i -x^0_{i-1})^2 \right>$ is considered
as an example. This harmonic model is exactly solvable and results
are shown in Fig.~\ref{fig_8}
\begin{figure}
\begin{center}
\includegraphics[width=90mm,angle=0]{fig_08.eps}
\end{center}
\caption{
\label{fig_8}
Temperature dependence of the mean squared displacement
$\left< [x_1(t)]^2 \right>$. Circles -- MD simulation of SDEs
\eqref{2-12}; dashed line -- model of mean rigidities in the harmonic
approximation. Averaging over 20 trajectories $2 \, 10^4$ t.u.
each.
}
\end{figure}
One can see that process $\left< [x_1^1(t)]^2 \right>$ damps out
in the model with constant rigidity in contrast to the case when
actual values \eqref{2-11} are used. And the growth of process
$\left< [x_1^1(t)]^2 \right>$, when temperature increases, is
mainly governed by {\it fluctuations} rather then the increasing
rigidities.
Additional evidence of threshold phenomena is the one-dimensional
analogue of the Mathieu equation
\begin{equation}
\label{Math_1}
\ddot x = -[1 + g \cos^2(t)] \, x.
\end{equation}
Different types of solutions depend on the parameter $g$ and
initial conditions. There exists such critical value $g_{\rm cr}$
that the solution is the superposition of periodic functions at $g
< g_{\rm cr}$, and the solution diverges $\propto \exp (\pm \mu
t)$ at $g > g_{\rm cr}$.
We consider an equation for the harmonic oscillator for one
variable $x$ with friction force
\begin{equation}
\label{Math_2}
\ddot x = - k(t) x - \dot x,
\end{equation}
where $k(t)$ -- stochastic rigidity. This equation is similar to
equation \eqref{2-12} for process ${\bf x}^1(t)$ if $k(t)= 1 + 3
z^2(t)$ and $z(t)$ is the stochastic process generated by the
dynamics of harmonic oscillator with noise term and friction force
at temperature $T$:
\begin{equation}
\label{Math_3}
\ddot z = - z + \xi - \dot z
\end{equation}
and spectral property $\left< \xi(t_1) \, \xi(t_2) \right> = 2 T
\delta(t_1 - t_2)$. The substitution $x \exp(- t/2) \to X$
excludes damping and \eqref{Math_2} becomes
\begin{equation}
\label{Math_4}
\ddot X = - [1 + 3 z^2(t)] X,
\end{equation}
what is similar to the Mathieu's equation \eqref{Math_1}.
Eqs. \eqref{Math_3}-\eqref{Math_4} have rich family of solutions
depending on initial conditions, temperature $T$ and the sequence
$\left\{ {\bf \xi} \right\}$. The solution is nearly harmonic
function at very low temperatures. The superposition of harmonic
functions is the solution at higher temperatures. At last there
exists such threshold temperature $T_{\rm thr}$ that the solution
diverges and is the product of harmonic functions by $\exp(\mu
t)$. The solution is the product of some stochastic process by
$\exp(\nu t)$ at much higher temperatures, and $\nu > \mu$. There
are many interesting intermediate solutions, and this problem
merits more attention. Considered examples show that an existence
of threshold phenomena is not exceptional and can occur in
different dynamical systems.
The threshold temperature $T_{\rm thr} \approx 4{.}1$ was found
for the lattice length $N = 5$. Larger lattice lengths were
considered and the dependence of $T_{\rm thr}$ on the lattice
length $N$ is shown in Fig.~\ref{fig_9}. Fitting gives dependence
$T_{\rm thr} \approx 6 \cdot 10^2 \, N^{-3}$.
\begin{figure}
\begin{center}
\includegraphics[width=90mm,angle=0]{fig_09.eps}
\end{center}
\caption{
\label{fig_9}
Dependence of $T_{\rm thr}$ vs. lattice length $N$ in log-log
coordinates. Solid line is the fitting $T_{\rm thr} \sim
N^{-3}$.
}
\end{figure}
\subsection{Time-resolved dynamics of process ${\widetilde{\bf x}}^1(t)$ }
Dynamics of process ${\widetilde {\bf x}}^1(t)$ at high
temperatures $T$ was analyzed above in terms of time-average
correlators. And time-resolved behavior of process ${\widetilde
{\bf x}}^1(t)$ at two temperatures $T$ of the background process
${\bf x}^0$ is shown in Fig.~\ref{fig_10}. As above, ${\Delta}(t)
= [\widetilde x_1^1(t)]^2$ was calculated.
\begin{figure}
\begin{center}
\includegraphics[width=60mm,angle=0]{fig_10a.eps}
\includegraphics[width=60mm,angle=0]{fig_10b.eps}
\end{center}
\caption{
\label{fig_10}
Dependence of ${\Delta}(t) = [\widetilde x_1^1(t)]^2$ on time at
different temperatures of process ${\bf x}^0(t)$. Left panel:
$T=4{.}3$; right panel: $T = 7{.}0$. $N=5$, integration step $h =
0{.}01$. Time average $\left. \left< \Delta(t) \right>
\right|_{t=0}^{t=10^5}$ are shown in horizontal solid lines.
}
\end{figure}
One can see that ${\Delta}(t)$ behaves highly irregular. And
numbers and heights of observed peaks increase with the growth of
temperature until the process becomes chaotic at high $T$. At
temperatures $T < T_{\rm thr}$ process $\widetilde {\bf x}_1^1(t)$
consists of individual rare peaks what can point to the
possibility that the energy can be transmitted by impulses.
\section{Sound, solitons and breathers in the $\beta$-FPU lattice}
\label{sec:sol_breath}
In accordance with the conjecture on the soliton contribution to
the heat conductivity \cite{Li_05, Vil02}, an attempt was made to
shed some light on this problem. The spatiotemporal correlator
$[y_k(t) \, y_{k+m}(t + \tau)]$ was calculated, where $y_i(t) =
x_i(t) - x_{i-1}(t)$ is the relative displacement of neighboring
particles in time instant $t$. Solitons, being highly correlated
displacements of particles, can leave a trace in correlation
function. Time shift $\tau = 20$ was fixed and spatial correlation
were calculated. Result is shown in Fig.~\ref{fig_11}.
\begin{figure}
\begin{center}
\includegraphics[width=90mm,angle=0]{fig_11.eps}
\end{center}
\caption{
\label{fig_11}
Correlator $y_{50}(t) \, y_{50+m}(t+20) $ as a function of the
lattice coordinate $n$. $\beta$-FPU lattice, $N = 101$, $T=2$.
Arrow points to $n=50$
}
\end{figure}
Two peaks in the correlation function, shifted by $m \approx \pm
25$, are visible. The velocity of their propagation is $v_{\rm p}
= m/\tau \approx 1{.}25$.
\subsection{Sound velocity in the $\beta$-FPU lattice}
The temperature dependence of the sound velocity in the
$\beta$-FPU lattice was discovered about a decade ago
\cite{Lep98}. The sound velocity was estimated $v_{\rm snd} \sim
\sqrt{1 + \alpha}$, where $\alpha$ -- parameter of renormalized
frequencies depending on the temperature. Asymptotic value of the
sound velocity in the high temperature limit $v_{\rm snd} \approx
1{.}22 \, T^{1/4}$ was derived recently in \cite{Li_10}. If this
formula apply to $T =2$ then $v_{\rm snd} \approx 1{.}45$ what
differs from $v_{\rm p} \approx 1{.}25$ found from correlation
functions above.
Below we derive more accurate expression for the sound velocity at
low temperatures. It was shown \cite{Ala01} that in nonlinear
systems there exists a spectrum of frequencies which are
proportional to the harmonic ones, according to a well defined
law. Then the $\beta$-FPU potential can be rewritten as
\begin{equation}
\label{f1}
u(y) = \left( 1 + \dfrac12 y^2 \right)\dfrac12 y^2
\end{equation}
and an expression in brackets can be replaced by an effective
rigidity $k_{\rm eff}$
\begin{equation}
\label{f2}
u(y) = k_{\rm eff}\dfrac{1}{2} y^2.
\end{equation}
As a result the lattice becomes harmonic and it is necessary to
find $k_{\rm eff}$. It can be done in terms of mean field
approximation (MFA). Mean value of potential energy is
\begin{equation}
\label{f3}
\left< u_{\rm p}(y) \right> = k_{\rm eff}\dfrac{1}{2} \left< y^2
\right>,
\end{equation}
where $\left< y^2 \right>$ is the mean of $y^2$. According to the
virial theorem, mean values of potential and kinetic energies are
equal in the harmonic lattice, that is $\left< u_{\rm p} \right> =
\left< u_{\rm k} \right>$. But the identity $\left< u_{\rm k}
\right> \equiv T/2$ holds for 1D systems. Then
\begin{equation}
\label{f5}
k_{\rm eff} \dfrac12 \left< y^2 \right> = \dfrac{T}{2}.
\end{equation}
The condition of self consistency of the MFA is
\begin{equation}
\label{f6}
\left( 1 + \dfrac12 \left< y^2 \right> \right) = k_{\rm
eff}
\end{equation}
and it follows that $\left< y^2 \right> = T/k_{\rm eff}$ and
substitution of this expression into \eqref{f6} gives the
self-consistent equation for $k_{\rm eff}$:
\begin{equation}
\label{f7}
1 + T/(2 k_{\rm eff}) = k_{\rm eff}
\end{equation}
with the solution
\begin{equation}
\label{f8}
k_{\rm eff} = \dfrac12 + \sqrt{\dfrac14 + \dfrac{T}{2}}.
\end{equation}
Eq.~\eqref{f8} defines the rigidity coefficient for the harmonic
lattice with the renormalized spectrum depending on temperature
$T$. Thereby the temperature renormalized sound velocity
\begin{equation}
\label{f9}
v_{\rm snd} = \sqrt{k_{\rm eff}} = \sqrt{\frac12 + \sqrt{\frac14 +
\frac{T}{2}}}, \ \ (m=1)
\end{equation}
and $v_{\rm snd} = 1{.}27$ for $T = 2$ what coincides with good
accuracy with $v_{\rm p} \approx 1{.}25$ found from correlation
functions. The high temperature asymptotic of the sound velocity
$v_{\rm snd} \sim 0{.}84 \, \left. T^{1/4} \right|_{T \gg 1}$. The
temperature dependence of sound velocity for temperatures in the
range $0 \le T \le 10$ is shown in Fig.~\ref{fig_13} and very good
agreement between analytical and ``experimental'' results is
observed.
\begin{figure}
\begin{center}
\includegraphics[width=90mm,angle=0]{fig_12.eps}
\end{center}
\caption{
\label{fig_12}
The temperature dependence of sound velocity $v_{\rm snd}$: solid
line -- dependence \eqref{f9}; empty circles -- MD simulation.
}
\end{figure}
If the mean field approximation is applied to the lattices with
cubic nonlinearity $u(y) = \frac12 y^2 \pm \frac13 y^3$ then no
dependence of the sound velocity on temperature is expected.
Really, $u(y) = \frac12 \left( 1 \pm \frac23 y \right) y^2$ and
$k_{\rm eff} = \left( 1 \pm \frac23 \left< y \right> \right) =1$
as $\left< y \right> = 1$.
The obtained results point to the fact that the nonlinearity can
be also ignored in describing the thermodynamics of the
$\beta$-FPU lattice, at least at $T < 10$, and the harmonic
lattice with renormalized rigidity coefficients \eqref{f8} is an
adequate model. This conjecture was checked at different
temperatures $T = 1, 2, 5, 10$ by the comparison of the total
energies computed in MD simulation of the $\beta$-FPU lattice and
its harmonic model with renormalized rigidity coefficients. Very
good coincidence of both energies testifies this hypothesis.
There exists an accurate expression for the specific potential
energy (mean potential energy of one oscillator) \cite{Lik09}
\begin{equation}
\label{thermodyn}
\dfrac{\left<U_{\rm p} \right>}{N} = \dfrac18
\left[ \dfrac{K_{5/4}(q) + K_{3/4}(q)}{2 \, K_{1/2}(q)} -1
\right]; \qquad q = 1/(8T)
\end{equation}
derived from the thermodynamics of the $\beta$-FPU lattice; $K$ --
modified Bessel functions. Specific potential energy
\eqref{thermodyn} and $\left<U_{\rm p} \right>/N$ computed in
harmonic approximation with rigidity coefficients \eqref{f8}
coincidence with good accuracy.
\subsection{Solitons and breathers in the $\beta$-FPU lattice}
The discrete $\beta$-FPU lattice can be reduced to the modified
Korteweg -- de~Vries (mKdV) equation in the continuum
long-wavelength approximation (see Appendix). The mKdV equation
has solutions in the form of solitons and moving breathers
\cite{Lam80}.
Solitons of compression and elongation has the form
\begin{equation}
\label{sol}
y(z,t) = \pm \dfrac{1}{\sqrt{6}} \, B \,
{\rm sech} \left\{ B \left[ z - \left( 1 + \dfrac{B^2}{24} \right) t
\right] \right\}
\end{equation}
where plus/minus signs stand for elongation/compression solitons;
$B$ -- single parameter which simultaneously determines amplitude,
width and velocity of soliton; $y$ -- local deformation of the
lattice; $z$ -- soliton coordinate at time $t$.
The two-parameteric breather solution is
\begin{equation}
\label{br}
y(z,T) = -4\beta\,
\mathrm{sech}{\Psi} \,
\displaystyle\frac{\cos{\Phi}-(\beta/\alpha)\sin{\Phi} \,
\mathrm{tanh} {\Psi}} {1+(\beta/\alpha)^2\sin^2{\Phi} \,
\mathrm{sech} {\Psi}},
\end{equation}
where
\begin{equation}
\label{phi_psi_2}
\begin{array}{ll}
\Psi \equiv 2 \beta (z - \gamma t) + \psi, \quad &
\Phi \equiv 2 \alpha (z - \delta t) + \phi , \qquad {\rm and} \\
\gamma = 4 \, (3 \alpha^2 -\beta^2), & \delta = 4\, (\alpha^2 -3\beta^2)
\end{array}
\end{equation}
and $\alpha, \beta$ are free parameters (see Appendix for more
details). It is necessary to make transformation from continuous
variables to discrete variables $y \to x_i - x_{i-1}$ and $z \to
i$ in an attempt to use soliton and breather solutions on the
discrete $\beta$-FPU lattice.
If soliton \eqref{sol} and breather \eqref{br} solutions exist in
the continuum limit, then the question arises: whether these
moving localized excitations can be observed on discrete lattice?
The answer is `yes' and below the visualization method is
suggested.
Recall that the visualization method for standing discrete
breathers is well known \cite{Bik99, Aub06}: boundaries with
friction forces absorb thermal noise (phonons) and standing
breathers can be easily seen. Other method is necessary to
visualize the moving excitations. Let we have the thermolized
lattice with $N$ oscillators at temperature $T$. If ``cold''
lattice (with zero velocities and displacements) is switched to
the thermalized lattice then solitons and breathers should ``run
out'' to the cold lattice and could be observed. Results are shown
in Fig.~\ref{fig_13} and solitons and breather are immediately
seen.
\begin{figure}
\begin{center}
\includegraphics[width=120mm,angle=0]{fig_13.eps}
\end{center}
\caption{
\label{fig_13}
Solitons and breather running out of the initially thermalized
lattice. A -- breather, B -- soliton of compression, C -- pair of
antisolitons (solitons of elongation). Arrow at $n = 200$ shows
the border separating initially thermalized and ``cold'' parts of
the lattice. Initial temperature of the left lattice part ($1 \leq
n \leq 200$), $T = 10$.
}
\end{figure}
Solitons and breather (shown in inserts to Fig.~\ref{fig_13}) move
faster then the sound front. At the first glance it is
inconsistent with the relation between velocities of sound and
solitons. The sound velocity at $T = 10$ is $v_{\rm snd} \approx
1{.}67$ what is larger then the maximal soliton velocity $(v_{\rm
sol})_{\rm max} \approx 1{.}3$ (see Appendix). But the temperature
of expanding thermal excitations gradually decreases and the sound
velocity also decreases according to \eqref{f9}. And there comes a
point when solitons, which have constant velocity, keep ahead the
sound front.
There are good grounds for believing that solitons and breathers
do exist in the $\beta$-FPU lattice. Very likely that the soliton
contribution to the heat conductivity increases with the growth of
temperature. Really, the temperature dependence of the soliton
density obeys the relation $n(T) \sim T^{1/3}$ for the Toda
lattice at low temperatures \cite{Mar91}. Conceivably the growth
of the solitons density with temperature might be an inherent
property of nonlinear systems.
Our results on the soliton contribution to the heat conductivity
are inconsistent with previous publication \cite{Li_10} where the
energy carriers are effective phonons rather than solitons.
The possibility of energy transfer by solitons was conjectured
three decades ago \cite{Tod79}. Less studied is the possibility of
energy transfer by breathers. One suggested mechanism is the
Targeted Energy Transfer \cite{Kop01, Man04} when an efficient
energy transfer can occur under a precise condition of nonlinear
resonance between discrete breathers. Various aspects and possible
applications of energy transfer by breathers are considered in
\cite{Aub06}.
\section{Conclusions}
\label{sec:concl}
In conclusion we briefly summarize our results. A new method for
the calculation of the heat conductivity is suggested. This is
done by the decomposing of the total dynamics into two parts:
equilibrium process ${\bf x}^0(t)$ at equal temperatures $T$ of
both lattice ends, and nonequilibrium process ${\bf x}^1(t)$ at
temperature $\Delta T$ of one end and zero temperature of the
other. This approach allows to extract and analyze the heat
conductivity in an explicit form.
The primary goal of the paper was to develop a method which would
allow to decrease the computational time at small temperature
gradients when fluctuations of the heat flux are usually too
large. It was supposed that at small temperature gradients, when
the harmonic approximation is valid and an expression for the heat
flux has the form \eqref{2-91}, an analytical averaging over
random Langevin noise terms can be done. This approach is very
efficient for the calculation of quadratic in ${\bf x}^1(t)$ terms
-- the gain was thousand-fold. But formulaes are very complex for
the linear terms and an efficient algorithm for their realization
was not found yet. By this expedient the objective has not been
met in full: the gain in computational time is obvious for small
$(N \lesssim 100)$ lattices, but decreases as the lattice length
increases. Nevertheless we suppose that the further analysis of
process ${\bf x}^1(t)$ can be useful as it is responsible for the
energy transfer.
The threshold phenomena are familiar in microcanonical ensembles
\cite{Lic08}. There exists two values of specific energy $E$. One
separates dynamical regime and weak chaos, and higher $E$
separates weak and strong chaos. It may be inferred that an
existence of threshold phenomena is also a common occurrence in
canonical ensembles. Really, a threshold temperature $T_{\rm thr}$
was found. The threshold temperature separates the different
behavior of process ${\bf x}^1(t)$: process ${\bf x}^1(t)$ damps
out at $T < T_{\rm thr}$ and reaches the stationary value at $T >
T_{\rm thr}$. It may be conceived that the soliton and breather
contributions to the heat conductivity increases with the growth
of temperature if $T > T_{\rm thr}$. Solitons and breathers can
emerge from either thermal fluctuations or higher order phonon
interactions. Additional experiments for nanosized systems at low
temperatures when $T < T_{\rm thr}$ can reveal some new features
omitted in the present work.
The modified Korteweg -- de~Vries equation is derived in the
continuum approximation for the $\beta$-FPU lattice. mKdV has
solutions in the form of compression/elongation solitons and
breathers. The stability of these quasi-particles was checked in
numerical experiments. Both types of excitations were directly
visualized.
On the other hand, it was found that the non-linear $\beta$-FPU
lattice can be reduced to the harmonic lattice with the
temperature renormalized frequency spectrum. This reduction allows
to reproduce adequately the heat conductivity and thermodynamics
of the parent lattice. These two, mutually contradictory,
properties of the $\beta$-FPU lattice, -- an inherent existence of
solitons and breathers, and its ``harmonic'' behavior, seem to be
very strange. Further analysis is necessary to solve this dilemma.
It is likely that some fraction of total heat conductivity is
conditioned by solitons and moving breathers. But their
contribution to the energy transfer deserves further
investigation.
The $\beta$-FPU lattice is unique in the sense that in the
continuum limit it has stable solutions in the form of solitons of
compression and elongation. If the number of compression and
elongation solitons is equal on average, then no macroscopic
changes in the lattice length appear and an additional energy of
deformation is negligible. It is an additional energetic factor
favoring the solitons existence.
The case is quite different for unsymmetrical potentials which in
the lowest order of the Taylor expansions have the form $u(y) =
\frac{a}{2}y^2 \pm \frac{b}{3}y^3$. $\alpha$-FPU, Toda, Morse,
Lennard-Jones potentials are the examples. All these ``cubic''
potentials can be easily reduced to the ordinary KdV equation with
the solution in the form of soliton of compression. And the large
number of solitons is highly energetically unfavorable due to
macroscopical compression of the lattice length.
|
\section{Introduction}
In this paper, we analyse the long-time dynamics of solutions to the Burgers-type equation
\begin{equation}\label{E:toy}
\phi_t + c\tanh\left(\frac{cx}{2}\right) \phi_x = \phi_{xx} + \phi_x^2,
\qquad c>0
\end{equation}
with small localized initial data, where $x\in\R$, $t>0$, and $\phi(x,t)$ is a scalar function. The key feature of this equation as opposed to the usual Burgers equation is that the characteristic speeds are $c>0$ at spatial infinity and $-c<0$ at spatial minus infinity: hence, transport is always directed away from the shock interface at $x=0$ and not towards $x=0$ as would be the case for the Lax shocks of the standard Burgers equation.
We are interested in (\ref{E:toy}) due to its close connection with the dynamics of coherent structures that arise in reaction-diffusion systems
\begin{equation} \label{E:rd}
u_t = Du_{xx} + f(u), \qquad x\in\R, \qquad u\in\R^n.
\end{equation}
A coherent structure (or defect) is a solution $u^*(x,t)$ of (\ref{E:rd}) that is time-periodic in an appropriate moving frame $y=x-c^*t$ and spatially asymptotic to wave-train solutions, which are spatially periodic travelling waves of (\ref{E:rd}). Such structures have been observed in many experiments and in various reaction-diffusion models, and we refer to \cite{SandstedeScheel04_classify} for references and to Figure~\ref{f:1} for an illustration of typical defect profiles. For the sake of simplicity, we shall assume from now on that the speed $c^*$ of the defect we are interested in vanishes, so that the coherent structure is time-periodic. Coherent structures can be classified into several distinct types \cite{SaarloosHohenberg,Hecke,SandstedeScheel04_classify} that have different stability and multiplicity properties. This classification involves the group velocities of the asymptotic wave trains, and we therefore briefly review their definition and features. Wave trains of (\ref{E:rd}) are solutions of the form $u(x,t)=u_\mathrm{wt}(kx-\omega t;k)$, where the profile $u_\mathrm{wt}(y;k)$ is $2\pi$-periodic in the $y$-variable. Thus, $k$ and $\omega$ represent the spatial wave number and the temporal frequency, respectively, of the wave train. Wave trains typically exist as one-parameter families, where the frequency $\omega=\omega_\mathrm{nl}(k)$ is a function, the so-called nonlinear dispersion relation, of the wave number $k$, which varies in an open interval. The group velocity $c_\mathrm{g}$ of the wave train with wave number $k$ is defined as
\[
c_\mathrm{g} := \frac{\rmd \omega_\mathrm{nl}}{\rmd k}(k).
\]
The group velocity is important as it is the speed with which small localized perturbations of a wave train propagate as functions of time $t$, and we refer to Figure~\ref{f:1}(iii) for an illustration and to \cite{DoelmanSSS} for a rigorous justification of this statement. The classification of coherent structures mentioned above is based on the group velocities $c_\mathrm{g}^\pm$ of the asymptotic wave trains at $x=\pm\infty$. We are interested in sources for which $c_\mathrm{g}^-<0<c_\mathrm{g}^+$ as illustrated in Figure~\ref{f:1}(i) so that perturbations are transported away from the defect core towards infinity. Sources are important as they actively select wave numbers in oscillatory media; examples of sources are the Nozaki--Bekki holes of the complex Ginzburg--Landau equation.
\begin{figure}
\centering\includegraphics{figure-1}
\caption{Panel (i) shows the graph of a source $u^*(x,t)$ as a function of $x$ for fixed time $t$: the group velocities of the asymptotic wave trains point away from the core of the coherent structure. Panel~(ii) illustrates the Floquet spectrum of a spectrally stable source: the two eigenvalues at the origin correspond to translations in space and time, which, in contrast to the essential spectrum, cannot be moved by exponential weights. Panel~(iii) shows the behaviour of small phase $\phi$ or wave number $\phi_x$ perturbations of a wave train: to leading order, they are transported with speed given by the group velocity $c_\mathrm{g}$ without changing their shape \cite{DoelmanSSS}.}\label{f:1}
\end{figure}
From now on, we focus on a given source and discuss its stability properties with respect to the reaction-diffusion system (\ref{E:rd}). Spectral stability of a source can be investigated through the Floquet spectrum of the period map of the linearization of (\ref{E:rd}) about the time-periodic source. Spectral stability of sources was investigated in \cite{SandstedeScheel04_classify}, and we now summarize their findings. The Floquet spectrum of a spectrally stable source will look as indicated in Figure~\ref{f:1}(ii). A source $u^*(x,t)$ has two eigenvalues at the origin with eigenfunctions $u^*_x(x,t)$ and $u^*_t(x,t)$; the associated adjoint eigenfunctions are necessarily exponentially localized, so that the source has a well defined spatial position and temporal phase. There will also be two curves of essential spectrum that touch the origin and correspond to phase and wave number modulations of the two asymptotic wave trains. It turns out that the two eigenvalues at the origin cannot be removed by posing the linearized problem in exponentially weighted function spaces; the essential spectrum, on the other hand, can be moved to the left by allowing functions to grow exponentially at infinity.
The nonlinear stability of spectrally stable sources has not yet been established, and we now outline why this is a challenging problem. From a purely technical viewpoint, an obvious difficulty is related to the fact that there is no spectral gap between the essential spectrum and the imaginary axis. As discussed above, such a gap can be created by posing the linear problem on function spaces that contain exponentially growing functions, but the nonlinear terms will then not even be continuous. To see that these are not just technical obstacles, it is illuminating to discuss the anticipated dynamics near a source from an intuitive perspective. If a source is subjected to a localized perturbation, then one anticipated effect is that the defect core adjusts its position and its temporal phase in response. From its new position, the defect will continue to emit wave trains with the same selected wave number but there will now be a phase difference between the asymptotic wave trains at infinity and those newly emitted near the core. In other words, we expect to see two phase fronts that travel in opposite directions away from the core as illustrated in Figure~\ref{f:2}. The resulting phase dynamics can be captured by writing the perturbed solution $u(x,t)$ as
\begin{equation}\label{e:phase}
u(x,t) = u^*(x+\phi(x,t),t) + w(x,t),
\end{equation}
where we expect that the perturbation $w(x,t)$ of the defect profile decays in time, while the phase $\phi(x,t)$ resembles an expanding plateau as indicated in Figure~\ref{f:2} whose height depends on the initial perturbation through the spatio-temporal displacement of the defect core.
\begin{figure}
\centering\includegraphics{figure-2}
\caption{The left panel contains a sketch of the space-time diagram of a perturbed source. The defect core will adjust in response to an imposed perturbation, and the emitted wave trains, whose maxima are indicated by the lines that emerge from the defect core, will therefore exhibit phase fronts that travel with the group velocities of the asymptotic wave trains away from the core towards $\pm\infty$. The right panel illustrates the profile of the anticipated phase function $\phi(x,t)$ defined in (\ref{e:phase}).}\label{f:2}
\end{figure}
The preceding discussion indicates that localized perturbations of a defect will not stay localized but result instead in phase fronts that propagate towards infinity. As a first step towards a general nonlinear stability result for sources in reaction-diffusion systems, we focus here on the Burgers-type equation
\begin{equation}\label{e:tm}
\phi_t + c\tanh\left(\frac{cx}{2}\right) \phi_x = \phi_{xx} + \phi_x^2, \qquad c>0
\end{equation}
for the phase function $\phi(x,t)$ introduced in (\ref{e:phase}). The rationale for choosing this model problem is as follows. First, as established formally in \cite{HowardKopell} and proved rigorously in \cite{DoelmanSSS}, the integrated viscous Burgers equation captures the phase dynamics of wave trains over long time intervals. Since the Burgers equation does not admit sources, we add the artificial advection term on the left-hand side which creates the characteristic speeds $\pm c$ at $x=\pm\infty$ that account for the outgoing group velocities of the asymptotic wave trains of the source. While the inhomogeneous advection term can be thought of as fixing the position $x=0$ of the core, the equation still has a family of constant solutions, which correspond to different temporal phases of the underlying hypothetical sources. We therefore feel that gaining a detailed understanding of the long-time dynamics of (\ref{e:tm}) for small localized initial data will shed significant light on the expected dynamics of sources and on the techniques needed to analyse their stability. We emphasize that the dynamics of wave trains of reaction-diffusion systems under non-localized phase perturbations was investigated only recently in \cite{SSSUecker}; the methods used there rely on renormalization-group techniques which seem difficult to generalize to sources. On the other hand, the analysis presented here is currently far more limited in terms of the equations it applies to.
The linearization of (\ref{e:tm}) about $\phi=0$ is given by
\begin{equation}\label{E:toy_lin}
\phi_t = \phi_{xx} - c\tanh\left(\frac{cx}{2}\right) \phi_x.
\end{equation}
The spectrum of the operator in (\ref{E:toy_lin}) is as shown in Figure~\ref{f:1}(ii) except that there is only one embedded eigenvalue at the origin that cannot be moved by exponential weights: the eigenfunction of the eigenvalue at $\lambda=0$ is given by the constant function $\frac{c}{4}$, and the associated adjoint eigenfunction $\psi$ is given by
\[
\psi(y) = \mathop\mathrm{sech}\nolimits^2\left(\frac{cy}{2}\right).
\]
Equation (\ref{E:toy_lin}) admits\footnote{Note that (\ref{E:toy_lin}) is the formal adjoint of the linearization of the standard viscous Burgers equation $u_t=u_{zz}-2uu_z$ about the Lax shock $\bar{u}(x)=(c/2)[1-\tanh(cx/2)]$ with $x=z-ct$, whose Green's function can be found via the linearized Cole--Hopf transformation by setting $w(x,t)=\mathop\mathrm{cosh}\nolimits(cx/2)\int_{-\infty}^x u(y,t)\rmd y$. The Green's function of (\ref{E:toy_lin}) can then be constructed by reversing the roles of $x$ and $y$ in the Green's function for the Lax shock linearization.} the Green's function
\begin{eqnarray} \label{E:toy_greens}
\mathcal{G}(x,y,t) & = &
\frac{1}{\sqrt{4\pi t}} \rme^{-\frac{(x-y+ct)^2}{4t}}\frac{1}{1+\rme^{cy}}
+ \frac{1}{\sqrt{4\pi t}} \rme^{-\frac{(x-y-ct)^2}{4t}}\frac{1}{1+\rme^{-cy}}
\\ \nonumber
& & \qquad + \frac{c}{4}\left[ \mathop\mathrm{errfn}\nolimits\left(\frac{y-x+ct}{\sqrt{4t}}\right) - \mathop\mathrm{errfn}\nolimits\left(\frac{y-x-ct}{\sqrt{4t}}\right) \right] \psi(y),
\end{eqnarray}
where the error function is given by
\[
\mathop\mathrm{errfn}\nolimits(z) = \frac{1}{\sqrt{\pi}}\int_{-\infty}^z \rme^{-s^2} \rmd s.
\]
The first two terms in the Green's function are Gaussians that move with speed $\pm c$ away from the core and decay at the rate $1/\sqrt{t}$, while the term comprised of the difference of the two error functions produces a plateau of constant height $\frac{c}{4}$ that spreads outward as indicated in Figure~\ref{f:3}.
\begin{figure}
\centering\includegraphics{figure-3}
\caption{Shown are the graphs of the function $\mathop\mathrm{errfn}\nolimits((-z+ct)/\sqrt{4t})-\mathop\mathrm{errfn}\nolimits((-z-ct)/\sqrt{4t})$ for smaller and larger values of $t$, which resemble plateaus of height approximately equal to one that spread outwards with speed $\pm c$.}\label{f:3}
\end{figure}
For sufficiently localized initial data $\phi_0(x)$, the solution $\phi(x,t)$ to the linear equation (\ref{E:toy_lin}) is therefore given by
\[
\phi(x,t) = \int_\mathbb{R} \mathcal{G}(x,y,t) \phi_0(y) \rmd y,
\]
which converges to a constant:
\[
\phi(x,t) \longrightarrow \frac{c}{4}\int_\mathbb{R}\psi(y) \phi_0(y) \rmd y \quad\mbox{as}\quad t\longrightarrow\infty
\]
for each fixed $x$. In particular, one cannot expect the solution associated with localized initial data to remain localized for all time. In fact, the same is true for the nonlinear equation (\ref{e:tm}): the Cole--Hopf transformation
\[
\tilde\phi(x,t) = \rme^{\phi(x,t)}-1, \qquad \phi(x,t) = \log\Big[1+\tilde\phi(x,t)\Big]
\]
relates solutions $\phi(x,t)$ of (\ref{e:tm}) and solutions $\tilde\phi(x,t)$ of the linearization (\ref{E:toy_lin}), and we conclude that solutions $\phi(x,t)$ of (\ref{e:tm}) are given by
\[
\phi(x,t) = \log\left[1+\int_\mathbb{R}\mathcal{G}(x,y,t)\phi_0(y)\rmd y \right].
\]
For $t\to\infty$, these solutions converge again pointwise in $x$ to the constant
\[
\log\left[ \int_\mathbb{R}\psi(y)\left(\rme^{\phi_0(y)} - 1\right)\rmd y + 1 \right].
\]
The Cole--Hopf transformation does not, however, extend to more general equations, and we therefore pursue here a different approach that, we hope, will also be useful when investigating the nonlinear stability of sources in general reaction-diffusion systems.
In order to analyze the dynamics of (\ref{e:tm}) in a way that may also be applicable in the case of a general reaction-diffusion equation, we need to find an appropriate ansatz for small-amplitude solutions. To do so, we define $\mathcal{B}(x,t)$ by
\[
\mathcal{B}(x,t) := \mathcal{G}(x,0,t+1),
\]
where $\mathcal{G}(x,y,t)$ is the Green's function defined in (\ref{E:toy_greens}), and note that
\begin{equation}\label{def-phia}
\phi^*(x,t,p) := \log\left( 1 + p \mathcal{B}(x,t) \right)
\end{equation}
is then a solution\footnote{In fact, $\phi^*(x,t,p)$ does not need to satisfy (\ref{e:tm}) exactly: our analysis goes through provided $\mathcal{B}(x,t)$ can be chosen such that $\phi^*(x,t,p)$ satisfies (\ref{e:tm}) approximately with an error of order of $(1+t)^{-1}$ times a Gaussian.} of (\ref{e:tm}) for each fixed $p\in\R$. We now investigate the long-time dynamics of solutions to (\ref{e:tm}) with initial data $\phi(x,0)=\phi_0(x)$ using the ansatz
\begin{equation}\label{E:ansatz}
\phi(x,t) = \phi^*(x,t,p(t)) + v(x,t),
\end{equation}
where $p(t)$ is a real-valued function, and $v(x,t)$ is a remainder term. At time $t=0$, we normalize the decomposition in (\ref{E:ansatz}) by choosing $p(0)=p_0$ such that
\[
\int_\mathbb{R} \psi(x) \big[\phi_0(x) - \phi^*(x,0,p_0)\big] \rmd x = 0.
\]
We will prove in \S\ref{S:proof} that a unique $p_0$ with this property exists for each sufficiently small localized initial condition $\phi_0$. Note that the leading behavior $ \phi^*(x,t,p(t))$ corresponds to a plateau with height $\log(1+p(t))$ of length $2ct$ that spreads outward with speed $\pm c$. The main result of this paper is as follows.
\begin{Theorem}\label{Thm:main}
For each $\gamma\in(0,\frac12)$, there exist constants $\epsilon_0,\eta_0,C_0,M_0>0$ such that the following is true. If $\phi_0\in C^1$ satisfies
\begin{equation}\label{e:phi0}
\epsilon := \|\rme^{x^2/M_0}\phi_0\|_{C^1} \leq \epsilon_0
\end{equation}
then the solution $\phi(x,t)$ of (\ref{e:tm}) with $\phi(\cdot,0)=\phi_0$ exists globally in time and can be written in the form
\[
\phi(x,t) = \phi^*(x,t,p(t)) + v(x,t)
\]
for appropriate functions $p(t)$ and $v(x,t)$ with $ \phi^*$ as in (\ref{def-phia}). Furthermore, there is a $p_\infty\in\R$ with $|p_\infty|\leq C_0$ such that
\[
|p(t)-p_\infty| \leq \epsilon C_0 \rme^{-\eta_0 t},
\]
and $v(x,t)$ satisfies
\[
|v(x,t)| \leq \frac{\epsilon C_0}{(1+t)^{\gamma}}
\left( \rme^{-\frac{(x+ct)^2}{M_0(t+1)}} + \rme^{-\frac{(x-ct)^2}{M_0(t+1)}} \right),
\quad
|v_x(x,t)| \leq \frac{\epsilon C_0}{(1+t)^{\gamma+1/2}}
\left( \rme^{-\frac{(x+ct)^2}{M_0(t+1)}} + \rme^{-\frac{(x-ct)^2}{M_0(t+1)}} \right)
\]
for all $t\geq0$. In particular, $\|v(\cdot,t)\|_{L^r}\to0$ as $t\to\infty$ for each fixed $r>\frac{1}{2\gamma}$.
\end{Theorem}
Note that our result assumes that the initial condition is strongly localized in space. We believe that this assumption can be relaxed significantly; for the purposes of this paper, however, phase fronts are created even by highly localized initial data, and the key difficulties are therefore present already in the more specialized situation of Theorem~\ref{Thm:main}.
The exponential convergence of $p(t)$ reflects the intuition that a source has a well-defined position due to the exponential localization of the adjoint eigenfunction $\psi$. The asymptotics of the perturbation $v$ is given by moving Gaussians that decay only like $(1+t)^{-\gamma}$ for each fixed $\gamma\in(0,\frac12)$, rather than with $(1+t)^{-\frac12}$ as expected from the dynamics of the viscous Burgers equation. This weaker result is due to the form (\ref{E:ansatz}) of our ansatz, which effectively creates a nonlinear term that is proportional to $g(x)uu_x$ for some function $g(x)$. While this term resembles the term $2uu_x$ in Burgers equation, it does not respect the conservation-law structure. Thus, we only obtain decay at the above rate. Although this may not be optimal, it allows us to avoid terms that grow logarithmically in the nonlinear iteration in \S\ref{S:proof}. We do not know if it is possible to improve this rate to $\gamma=\frac12$ by adjusting our ansatz appropriately.
The remainder of this paper, \S\ref{S:proof}, is devoted to the proof of Theorem~\ref{Thm:main}.
\section{Proof of the main theorem} \label{S:proof}
To prove Theorem~\ref{Thm:main}, we set up integral equations for $(p,v)$ and solve them using a nonlinear iteration scheme. Throughout the proof, we denote by $C$ possibly different positive constants that depend only on the underlying equation but not on the initial data or on space or time.
\subsection{Derivation of an integral formulation}
Substituting the ansatz
\[
\phi(x,t) = \log\left( 1+p(t)\mathcal{B}(x,t) \right) + v(x,t), \qquad
\mathcal{B}(x,t) := \mathcal{G}(x,0,t+1),
\]
into the Burgers-type equation
\[
\phi_t + c\tanh\left(\frac{cx}{2}\right) \phi_x = \phi_{xx} + \phi_x^2,
\]
we find that $(p,v)$ needs to satisfy the equation
\begin{equation}\label{E:v}
v_t = v_{xx} - c\tanh\left(\frac{cx}{2}\right) v_x + v_x^2 - \frac{\dot{p}}{1+\frac{c }{4}p} \mathcal{G}(x,0,t+1)+ \mathcal{N}(x,t,p,\dot{p},v_x),
\end{equation}
where the nonlinear function $\mathcal{N}$ is given by
\[
\mathcal{N}(x,t,p,\dot{p},v_x) := \frac{2pv_x\mathcal{B}_x(x,t)}{1+p\mathcal{B}(x,t)} + \dot{p} \left( \frac{\mathcal{B}(x,t)}{1+\frac{c}{4}p} - \frac{\mathcal{B}(x,t)}{1+p\mathcal{B}(x,t)} \right).
\]
The idea of the proof is to use an appropriate integral representation for $v$ that will allow us to set up a nonlinear iteration argument to show that solutions $(p,v)$ of (\ref{E:v}) exist and that they satisfy the desired decay estimates in space and time. Recall from (\ref{E:toy_greens}) the expression
\begin{eqnarray} \label{e:gf}
\mathcal{G}(x,y,t) & = &
\frac{1}{\sqrt{4\pi t}} \rme^{-\frac{(x-y+ct)^2}{4t}}\frac{1}{1+\rme^{cy}}
+ \frac{1}{\sqrt{4\pi t}} \rme^{-\frac{(x-y-ct)^2}{4t}}\frac{1}{1+\rme^{-cy}}
\\ \nonumber
& & + \frac{c}{4}\left( \mathop\mathrm{errfn}\nolimits\left(\frac{y-x+ct}{\sqrt{4t}}\right) - \mathop\mathrm{errfn}\nolimits\left(\frac{y-x-ct}{\sqrt{4t}}\right) \right) \psi(y)
\end{eqnarray}
for the Green's function of the linear problem (\ref{E:toy_lin}) and note that it satisfies the identity
\[
\int_\mathbb{R} \mathcal{G}(x,y,t-s)\mathcal{G}(y,0,s+1)\rmd y = \mathcal{G} (x,0,t+1).
\]
Using this identity and the variation-of-constants formula, we can rewrite (\ref{E:v}) in integral form and obtain\footnote{We shall often use the notation $[f+g](x,t)$ to denote $f(x,t)+g(x,t)$.}
\begin{eqnarray}\label{E:v_init}
v(x,t) & = &- \frac{4}{c} \mathcal{G}(x,0,t+1)\left( \log\left(1+\frac{c p(t)}{4} \right) - \log\left(1+\frac{c p(0)}{4}\right) \right) \\ \nonumber &&
+ \int_\mathbb{R}\mathcal{G}(x,y,t)v_0(y)\rmd y
+ \int_0^t \int_\mathbb{R} \mathcal{G}(x,y,t-s) \left[ v_y^2 +\mathcal{N}(\cdot,\cdot,p,\dot{p},v_y)\right](y,s) \rmd y \rmd s.
\end{eqnarray}
The expression (\ref{e:gf}) of the Green's function $\mathcal{G}(x,y,t)$ shows that the terms involving the error functions do not provide any temporal decay. We need to treat these terms separately and therefore write the Green's function $\mathcal{G}(x,y,t)$ as
\[
\mathcal{G}(x,y,t) = \mathcal{E}(x,y,t) + \tilde{\mathcal{G}}(x,y,t),
\]
where
\[
\mathcal{E}(x,y,t) = e(x,t)\psi(y), \qquad
e(x,t) := \frac{c}{4}\left( \mathop\mathrm{errfn}\nolimits\left(\frac{x+ct}{\sqrt{4t}}\right) - \mathop\mathrm{errfn}\nolimits\left(\frac{x-ct}{\sqrt{4t}}\right) \right)
\]
and
\begin{eqnarray}\label{def-tG}
\tilde{\mathcal{G}}(x,y,t) & := & \mathcal{G}(x,y,t) - \mathcal{E}(x,y,t)
\\ & = & \nonumber
\frac{1}{\sqrt{4\pi t}} \rme^{-\frac{(x-y+ct)^2}{4t}}\frac{1}{1+\rme^{cy}}
+ \frac{1}{\sqrt{4\pi t}} \rme^{-\frac{(x-y-ct)^2}{4t}}\frac{1}{1+\rme^{-cy}}
\\ && \nonumber
+ \left( \mathop\mathrm{errfn}\nolimits\left(\frac{y-x+ct}{\sqrt{4t}}\right) - \mathop\mathrm{errfn}\nolimits\left(\frac{y-x-ct}{\sqrt{4t}}\right) \right) \frac{c}{4}\mathop\mathrm{sech}\nolimits^2(\frac{cy}{2})
\\ && \nonumber
- \left( \mathop\mathrm{errfn}\nolimits\left(\frac{-x+ct}{\sqrt{4t}}\right) - \mathop\mathrm{errfn}\nolimits\left(\frac{-x-ct}{\sqrt{4t}}\right) \right) \frac{c}{4}\mathop\mathrm{sech}\nolimits^2(\frac{cy}{2}).
\end{eqnarray}
A calculation shows that
\[
\left| \mathop\mathrm{errfn}\nolimits\left(\frac{y-x\pm ct}{\sqrt{4t}}\right) - \mathop\mathrm{errfn}\nolimits\left(\frac{-x\pm ct}{\sqrt{4t}}\right) \right| \frac{c}{4} \mathop\mathrm{sech}\nolimits^2(\frac{cy}{2}) \leq
C t^{-1/2} \left(\rme^{-\frac{(x-y+ct)^2}{4t}} + \rme^{-\frac{(x-y-ct)^2}{4t}} \right) \rme^{-c|y|/4},
\]
and we conclude that
\begin{equation}\label{e:tildeg}
|\tilde{\mathcal{G}}(x,y,t)| \leq C t^{-1/2} \left( \rme^{-\frac{(x-y+ct)^2}{4t}} + \rme^{-\frac{(x-y-ct)^2}{4t}} \right).
\end{equation}
To analyse (\ref{E:v_init}), we consider the initial condition $\phi_0(x)=\phi^*(x,0,p(0))+v_0(x)$ and show that, if $\phi_0 $ is sufficiently small in $L^\infty$, then $p(0)=p_0$ can be chosen such that
\[
\int_\mathbb{R} \psi(y) v_0(y) \rmd y = 0
\]
or, equivalently,
\begin{equation}\label{choice-p0}
\int_\mathbb{R} \psi(y) [\phi_0(y) - \phi^*(y,0,p_0)] \rmd y = 0.
\end{equation}
To prove this claim, we observe that
\begin{equation}\label{choice-p1}
\phi^*(y,0,p_0) = \log(1 + p_0 \mathcal{B}(y,0)) = p_0 \mathcal{G}(y,0,1) + \rmO(p_0^2).
\end{equation} We note that the term $\rmO(p_0^2)$ in (\ref{choice-p1}) is bounded uniformly in $y\in\R$, and substitution into (\ref{choice-p0}) therefore gives the equation
\[
\int_\mathbb{R} \psi(y) \phi_0(y) \rmd y = p_0 \int_\mathbb{R} \psi(y) \mathcal{G}(y,0,1)\rmd y + \rmO(p_0^2),
\]
which can be solved uniquely for $p_0=p(0)$ near zero for each $\phi_0\in L^\infty$ for which $\| \phi_0\|_{L^\infty}$ is small enough. In particular, there is a constant $C>0$ such that the resulting initial value $p(0)$ satisfies
\[
|p(0)| \leq C \| \phi_0\|_{L^\infty} \leq C\epsilon.
\]
We will construct $p(t)$ such that
\begin{equation}\label{eqs-dotp}
\dot{p}(t) = \left(1+ \frac{cp(t)}{4} \right) \int_\mathbb{R} \psi(y)\left[ v_y^2 +\mathcal{N}(\cdot,\cdot,p,\dot{p},v_y)\right](y,t) \rmd y
\end{equation}
or, equivalently, that
\begin{equation}\label{eqs-p}
\log\left(1+\frac{c p(t)}{4}\right) = \log\left(1+\frac{cp_0}{4}\right) +\frac{c}{4} \int_0^t \int_\mathbb{R} \psi(y)\left[ v_y^2 +\mathcal{N}(\cdot,\cdot,p,\dot{p},v_y)\right](y,s) \rmd y \rmd s
\end{equation}
for all $t\geq0$. Substituting (\ref{eqs-p}) into (\ref{E:v_init}), we obtain the equation
\begin{eqnarray}\label{E:v_p}
v(x,t) & = & -\frac{4}{c}\tilde{\mathcal{G}}(x,0,t+1) \left( \log\left(1+\frac{c p(t)}{4}\right) - \log\left(1+\frac{c p_0}{4}\right) \right) + \int_\mathbb{R}\tilde{\mathcal{G}}(x,y,t)v_0(y)\rmd y \\ \nonumber &&
+ \int_0^t \int_\mathbb{R} \tilde{\mathcal{G}}(x,y,t-s) \left[ v_y^2 +\mathcal{N}(\cdot,\cdot,p,\dot{p},v_y) \right](y,s) \rmd y \rmd s
\\ \nonumber &&
+ \int_0^t \int_\mathbb{R} \left(e(x,t-s) - e(x,t+1)\right) \psi(y) \left[ v_y^2 +\mathcal{N}(\cdot,\cdot,p,\dot{p},v_y)\right](y,s) \rmd y \rmd s
\end{eqnarray}
for $v(x,t)$.
To analyse solutions of (\ref{eqs-dotp}) and (\ref{E:v_p}), we introduce template functions that capture the anticipated spatio-temporal behaviour of solutions. For each fixed choice of $\gamma\in(0,\frac12)$ and $M>0$, we define
\begin{equation} \label{E:def_theta}
\theta_1(x,t) = \frac{1}{(1+t)^{\gamma}} \left( \rme^{-\frac{(x-ct)^2}{M(t+1)}} + \rme^{-\frac{(x+ct)^2}{M(t+1)}}\right), \qquad
\theta_2(x,t) = \frac{1}{(1+t)^{\gamma+1/2}} \left( \rme^{-\frac{(x-ct)^2}{M(t+1)}} + \rme^{-\frac{(x+ct)^2}{M(t+1)}}\right)
\end{equation}
and let
\[
h_1(t) := \sup_{y\in\mathbb{R}, 0\leq s\leq t} \left[\frac{|v|}{\theta_1}+\frac{|v_y|}{\theta_2} \right](y,s), \qquad
h_2(t) := \sup_{0\leq s\leq t} |\dot{p}(s)|\rme^{c^2 s/M}, \qquad
h(t) := h_1(t) + h_2(t).
\]
We will later choose $M\gg1$.
We remark that we do know existence and smoothness of $(v,p)$ for short times: Indeed, we can solve the original PDE for $\phi(x,t)$ for short times and can substitute the resulting expression into (\ref{eqs-dotp}) upon using (\ref{E:ansatz}). The resulting integral equation has a solution $\dot{p}(t)$ for small times and, using again (\ref{E:ansatz}), we find a smooth function $v$ that then satisfies (\ref{E:v_p}). Furthermore, using that $\phi_0$ satisfies (\ref{e:phi0}) by assumption, we see that $h(t)$ is well defined and continuous for $0<t\ll1$. Finally, standard parabolic theory implies that $h(t)$ retains these properties as long as $h(t)$ stays bounded. The key issue is therefore to show that $h(t)$ stays bounded for all times $t>0$, and this is what the following proposition asserts.
\begin{Proposition}\label{P:claim-zeta}
For each $\gamma\in(0,\frac12)$, there exist positive constants $\epsilon_0,C_0,M$ such that
\begin{equation}\label{claim-zeta}
h_1(t) \leq C_0 (\epsilon + h_2(t) + h(t)^2), \qquad
h_2(t) \leq C_0 (\epsilon + h(t)^2).
\end{equation}
for all $t\geq0$ and all initial data $u_0$ with $\epsilon:=\|\rme^{x^2/M}\phi_0\|_{C^1}\leq\epsilon_0$.
\end{Proposition}
Using this proposition, we can add the inequalities in (\ref{claim-zeta}) and eliminate $h_2$ on the right-hand side to obtain
\[
h(t) \leq C_0(C_0+1)(\epsilon + h(t)^2).
\]
Using this inequality and the continuity of $h(t)$, we find that $h(t)\leq 2C_0(C_0+1)\epsilon $ for all $t\geq0$ provided $0<\epsilon\leq\epsilon_0$ is sufficiently small. Thus, Theorem~\ref{Thm:main} will be proved once we establish Proposition~\ref{P:claim-zeta}. The following sections will be devoted to proving this proposition.
\subsection{Estimates of the nonlinear term}
We begin by deriving estimates of the nonlinear term
\[
\mathcal{N}(x,t,p,\dot{p},v_x) = \frac{2pv_x\mathcal{B}_x(x,t)}{1+p\mathcal{B}(x,t)} + \dot{p} \left( \frac{\mathcal{B}(x,t)}{1+\frac{c}{4}p} - \frac{\mathcal{B}(x,t)}{1+p\mathcal{B}(x,t)} \right)
\]
that appears in (\ref{eqs-dotp}) and (\ref{E:v_p}). Recalling that
\[
e(x,t) = \frac{c}{4}\left( \mathop\mathrm{errfn}\nolimits\left(\frac{x+ct}{\sqrt{4t}}\right) - \mathop\mathrm{errfn}\nolimits\left(\frac{x-ct}{\sqrt{4t}}\right) \right)
\]
and using the definition of the error function, we see that
\[
\left| e(x,t) \left(\frac{c}{4} - e(x,t)\right) \right| \leq C \left(\rme^{-\frac{(x + ct)^2}{8t}} +\rme^{-\frac{(x - ct)^2}{8t}} \right).
\]
Since $\mathcal{B}(x,t)=e(x,t+1)+\tilde{\mathcal{G}}(x,0,t+1)$, we conclude from (\ref{e:tildeg}) that
\[
\left|\mathcal{B}(x,t) \left(\frac{c}{4}-\mathcal{B}(x,t)\right) \right| \leq C \left(\rme^{-\frac{(x + ct)^2}{8(t+1)}} +\rme^{-\frac{(x - ct)^2}{8(t+1)}} \right)
\]
and thus
\[
\frac{\mathcal{B}(x,t)}{1+p\mathcal{B}(x,t)} - \frac{\mathcal{B}(x,t)}{1+\frac{c}{4}p} \leq C|p| \left(\rme^{-\frac{(x + ct)^2}{8(t+1)}} +\rme^{-\frac{(x - ct)^2}{8(t+1)}} \right).
\]
In addition, we have
\[
\left| \frac{\mathcal{B}_x(x,t)}{1+p\mathcal{B}(x,t)} \right| \leq C(1+t)^{-1/2} \left(\rme^{-\frac{(x + ct)^2}{8(t+1)}} + \rme^{-\frac{(x - ct)^2}{8(t+1)}} \right).
\]
Combining these estimates, we therefore obtain
\begin{equation}\label{bound-N}
|\mathcal{N}(x,t,p,\dot{p},v_x)| \leq C \left( (1+t)^{-1/2} |p| |v_x| + |p\dot{p}| \right) \left(\rme^{-\frac{(x + ct)^2}{8(t+1)}} +\rme^{-\frac{(x - ct)^2}{8(t+1)}} \right)
\end{equation}
uniformly in $x\in\mathbb{R}$ and $t\geq0$, provided $p$ is sufficiently small.
\subsection{Estimates for $h_2(t)$}
To establish the claimed estimate for $h_2(t)$, recall from (\ref{eqs-dotp}) that
\begin{equation}\label{est-qdot1}
|\dot{p}(t)| \leq C (1+|p(t)|) \int_\mathbb{R} \psi(y) \Big| \left[ v_y^2 + \mathcal{N}(\cdot,\cdot,p,\dot{p},v_y) \right](y,t) \Big| \rmd y.
\end{equation}
We shall show that there is a constant $C_1=C_1(M)$ such that
\begin{equation}\label{c:1}
|\dot{p}(t)|\leq C_1 \rme^{-c^2 t/M} (\epsilon + h^2(t)),
\end{equation}
which then establishes the estimate for $h_2(t)$ stated in Proposition~\ref{P:claim-zeta}. To show (\ref{c:1}), we note that the definitions of $p(t)$ and $h_2(t)$ imply that
\begin{equation}\label{p-estimate}
|p(t)| \leq |p(0)| + \int_0^t |\dot{p}(s)| \rmd s
\leq |p(0)| + \int_0^t \rme^{-c^2 s/M} h_2(t) \rmd s
\leq C_1 (\epsilon + h_2(t))
\end{equation}
and therefore
\begin{equation} \label{E:p_est}
|p(t)\dot{p}(t)| \leq C_1 \rme^{-c^2 t/M} h_2(t)(\epsilon + h_2(t)) \leq C_1 \rme^{-c^2 t/M}(\epsilon + h(t)^2).
\end{equation}
Next, we use the estimate $|\psi(y)|\leq 2\rme^{-c|y|}$, the bound (\ref{bound-N}) on $\mathcal{N}$, and the inequality
\[
\rme^{-\frac{c|y|}{2}} \rme^{-\frac{(y\pm ct)^2}{M(1+t)}} \leq C_1 \rme^{-\frac{c|y|}{4}} \rme^{-\frac{c^2 t}{M}},
\]
which holds for each $M\geq8$, to conclude that
\begin{equation}\label{est-phiN}
|\psi(y)v_y^2(y,t)| \leq \frac{C_1}{(1+t)^{1+2\gamma}} C \rme^{-\frac{c}{2} |y|} \rme^{-\frac{2c^2}{M}t} h_1(t)^2
\end{equation}
and therefore
\begin{eqnarray}
|\psi(y)\mathcal{N}(y,t,p,\dot{p},v_y)| & \leq & C_1 (1+t)^{-1/2} \left( \rme^{-\frac{(x + ct)^2}{8(t+1)}} + \rme^{-\frac{(x - ct)^2}{8(t+1)}} \right) \psi (y) |v_y(y,t)| |p(t)|
\nonumber \\ \nonumber & &
+ C_1 \left( \rme^{-\frac{(x + ct)^2}{8(t+1)}} + \rme^{-\frac{(x - ct)^2}{8(t+1)}} \right) \psi(y)|p(t)\dot{p}(t)|
\\ \nonumber & \leq &
C_1 (1+t)^{-1/2} \rme^{-\frac{c}{2}|y| - \frac{2c^2}{M} t} h_1(t) (\epsilon + h_2(t)) + C_1 \rme^{-\frac{c}{2}|y| - \frac{2c^2}{M} t}(\epsilon + h(t)^2) \\ \label{est-phiN3} & \leq &
C_1\rme^{-\frac{c}{2}|y| - \frac{2c^2}{M} t} (\epsilon + h(t)^2).
\end{eqnarray}
Using these estimates in (\ref{est-qdot1}), we arrive readily at (\ref{c:1}), thus proving the estimate for $h_2(t)$ stated in Proposition~\ref{P:claim-zeta}.
\subsection{Estimates for $v(x,t)$ and $v_x(x,t)$}\label{sec-estv}
In this section, we will establish the pointwise bounds
\begin{eqnarray}
|v(x,t)| & \leq & C_1 (\epsilon + h_2(t) + h(t)^2) \theta_1(x,t)
\label{key-est-phi} \\ \label{key-est-phix}
|v_x(x,t)| & \leq & C_1 (\epsilon + h_2(t) + h(t)^2) \theta_2(x,t)
\end{eqnarray}
for $v(x,t)$ and $v_x(x,t)$, respectively, which taken together prove the inequality for $h_1(t)$ stated in Proposition~\ref{P:claim-zeta}. In particular, the proof of Proposition~\ref{P:claim-zeta} is complete once the two estimates above are established. We denote by $C_1$ possibly different constants that depend only on the choice of $\gamma,M$ so that $C_1=C_1(\gamma,M)$.
We focus first on $v(x,t)$. From the integral formulation (\ref{E:v_p}) for $v(x,t)$, we find that
\begin{eqnarray}\label{phi-estimate}
|v(x,t)| & \leq &
C_1\left(|p_0| + |p(t)|\right)\tilde{\mathcal{G}}(x,0,t+1) + \int_\mathbb{R} \tilde{\mathcal{G}}(x,y,t)|v_0(y)|\rmd y
\\ \nonumber & &
+ \int_0^t \int_\mathbb{R} \Big| \tilde{\mathcal{G}}(x,y,t-s) \left[ v_y^2 + \mathcal{N}(\cdot,\cdot,p,\dot{p},v_y)\right](y,s) \Big| \rmd y \rmd s
\\ \nonumber & &
+ \int_0^t \int_\mathbb{R} \Big| [e(x,t-s+1) - e(x,t+1)] \psi(y) \left[ v_y^2 + \mathcal{N}(\cdot,\cdot,p,\dot{p},v_y)\right](y,s) \Big| \rmd y \rmd s.
\end{eqnarray}
In the remainder of this section, we will estimate the right-hand side of (\ref{phi-estimate}) term by term.
First, we note that (\ref{e:tildeg}) implies that there is a constant $C_1=C_1(M)$ so that $\tilde{\mathcal{G}}(x,0,t+1)\leq C_1\theta_1(x,t)$. Using (\ref{p-estimate}) and the fact that $|p(0)|\leq C_1\epsilon$, we therefore obtain
\[
(|p(0)| + |p(t)|) \tilde{\mathcal{G}}(x,0,t+1) \leq C_1 (\epsilon + h_2(t)) \theta_1(x,t),
\]
which is the desired estimate for the first term on the right-hand side of (\ref{phi-estimate}).
Next, we consider the integral term in (\ref{phi-estimate}) that involves the initial data $v_0$. Using (\ref{e:tildeg}) together with our assumption (\ref{e:phi0}) on $\phi_0$, and hence on $v_0$, we see that
\begin{equation}\label{est-initial}
\int_\mathbb{R} |\tilde{\mathcal{G}}(x,y,t) v_0(y)| \rmd y \leq C \epsilon \int_{\mathbb{R}} t^{-1/2} \left(\rme^{-\frac{(x-y+ct)^2}{4t}} + \rme^{-\frac{(x-y-ct)^2}{4t}} \right) \rme^{-\frac{y^2}{M}} \rmd y,
\end{equation}
which is clearly bounded by $\epsilon C_1\theta_1(x,t)$ for $t\geq1$ upon using
\[
\rme^{-\frac{(x-y\pm ct)^2}{4t}} \rme^{-\frac{y^2}{M}} \leq C_1 \rme^{-\frac{(x\pm ct)^2}{Mt}} \rme^{-\frac{y^2}{2M}}.
\]
For $t\leq1$, we can use the estimates
\begin{eqnarray*}
\rme^{-\frac{(x-y\pm ct)^2}{8t}} \rme^{-\frac{y^2}{M}} \;\leq\; 2\rme^{-\frac{(x-y)^2}{8t}}\rme^{-\frac{y^2}{M}} & \leq & C_1 \rme^{-\frac{x^2}{2M}} \\
\int_{\mathbb{R}} t^{-1/2} \left(\rme^{-\frac{(x-y+ct)^2}{8t}} + \rme^{-\frac{(x-y-ct)^2}{8t}} \right) \rmd y & \leq & C_1
\end{eqnarray*}
to conclude that the integral in (\ref{est-initial}) is again bounded by $\epsilon C_1 \theta_1(x,t)$.
We now consider the remaining two integrals in (\ref{phi-estimate}). Note that the definition of $h_1$ gives
\[
|v_y(y,s)| \leq \theta_2(y,s) h_1(s).
\]
Using this fact together with the estimates (\ref{p-estimate}) and (\ref{E:p_est}) and the bound (\ref{bound-N}), we obtain
\begin{eqnarray*}
|\mathcal{N}(y,s,p(s),\dot{p}(s),v_y(y,s))| & \leq & C_1 (1+s)^{-1/2} \left(\rme^{-\frac{(y + cs)^2}{8(s+1)}} + \rme^{-\frac{(y - cs)^2}{8(s+1)}} \right) |v_y(y,s)| |p(s)| \\ & &
+ C_1 \left(\rme^{-\frac{(y + cs)^2}{8(s+1)}} + \rme^{-\frac{(y - cs)^2}{8(s+1)}} \right) |p(s)\dot{p}(s)|
\\ & \leq &
C_1 (\epsilon + h_2(t)) h_1(t) \theta_2(y,s) (1+s)^{-1/2} \left(\rme^{-\frac{(y + cs)^2}{8(s+1)}} + \rme^{-\frac{(y - cs)^2}{8(s+1)}} \right) \\ &&
+ C_1 (\epsilon + h(t)^2) \rme^{-\frac{c^2 s}{M}} \left(\rme^{-\frac{(y + cs)^2}{8(s+1)}} + \rme^{-\frac{(y - cs)^2}{8(s+1)}} \right) \\ & \leq &
C_1 (\epsilon + h(t)^2) \left[ (1+s)^{\gamma-1/2} \theta_1(y,s) \theta_2(y,s) + (1+s)^{\gamma} \theta_1(y,s) \rme^{-\frac{c^2 s}{M}} \right].
\end{eqnarray*}
In \S\ref{S:lemmas}, we will prove the following result for our spatio-temporal template functions.
\begin{Lemma}\label{lem-est-tG}
For each sufficiently large $M$, there is a constant $C_1$ so that
\begin{eqnarray*}
\int_0^t \int_\mathbb{R} |\tilde{\mathcal{G}}(x,y,t-s)| \left[ \theta_2^2 + (1+s)^{\gamma-1/2}\theta_1\theta_2+(1+s)^{\gamma} \theta_1\rme^{-c^2 s/M}\right](y,s) \rmd y \rmd s & \leq & C_1 \theta_1(x,t) \\
\int_0^t \int_\mathbb{R} |\tilde{\mathcal{G}}_x(x,y,t-s)| \left[\theta_2^2 + (1+s)^{\gamma-1/2}\theta_1\theta_2+(1+s)^{\gamma} \theta_1\rme^{-c^2 s/M}\right](y,s) \rmd y \rmd s & \leq & C_1\theta_2(x,t).
\end{eqnarray*}
\end{Lemma}
Using this lemma and the above estimates for $v_y^2+\mathcal{N}$, we obtain the desired estimate
\[
\int_0^t \int_\mathbb{R} \tilde{\mathcal{G}}(x,y,t-s)\left[ v_y^2 +\mathcal{N}(\cdot,\cdot,p,\dot{p},v_y)\right](y,s) \rmd y \rmd s \leq C_1 (\epsilon + h(t)^2) \theta_1(x,t).
\]
Finally, we have the following lemma, whose proof is again given in \S\ref{S:lemmas}, which provides the desired estimate for the last integral in (\ref{phi-estimate}).
\begin{Lemma}\label{lem-est-ediff}
For each sufficiently large $M$, there is a constant $C_1$ so that
\begin{eqnarray*}
\int_0^t \int_\mathbb{R} |e(x,t-s+1) - e(x,t+1)|\psi(y) \left[v_y^2 + \mathcal{N}(\cdot,\cdot,p,\dot{p},v_y) \right](y,s) \rmd y \rmd s & \leq & C_1 (\epsilon + h(t)^2)\theta_1(x,t) \\
\int_0^t \int_\mathbb{R} |e_x(x,t-s+1) - e_x(x,t+1)|\psi(y)\left[v_y^2 +\mathcal{N}(\cdot,\cdot,p,\dot{p},v_y)\right](y,s) \rmd y \rmd s & \leq & C_1(\epsilon + h(t)^2)\theta_2(x,t).
\end{eqnarray*}
\end{Lemma}
In summary, combining the estimates obtained above, we have established the claimed estimate (\ref{key-est-phi}), and it remains to derive the estimate (\ref{key-est-phix}) to complete the proof of Proposition~\ref{P:claim-zeta}. Taking the $x$-derivative of equation (\ref{E:v_p}), we see that
\begin{eqnarray}\label{phix-estimate}
|v_x(x,t)| & \leq & \left(|p(0)| + |p(t)| \right) \tilde{\mathcal{G}}_x(x,0,t) + \int_\mathbb{R}\tilde{\mathcal{G}}_x(x,y,t) v_0(y)\rmd y \\ \nonumber &&
+ \int_0^t \int_\mathbb{R} \tilde{\mathcal{G}}_x(x,y,t-s) \left[ v_y^2 + \mathcal{N}(\cdot,\cdot,p,\dot{p},v_y) \right](y,s) \rmd y \rmd s
\\ \nonumber & &
+ \int_0^t \int_\mathbb{R} [e_x(x,t-s+1) - e_x(x,t+1)]\psi(y) \left[ v_y^2 + \mathcal{N}(\cdot,\cdot,p,\dot{p},v_y) \right](y,s) \rmd y \rmd s.
\end{eqnarray}
Applying the second estimate in Lemmas~\ref{lem-est-tG} and~\ref{lem-est-ediff} to (\ref{phix-estimate}) and using that $\tilde{\mathcal{G}}_x(x,0,t)\leq C\theta_2(x,t)$, we immediately obtain (\ref{key-est-phix}).
\subsection{Proofs of Lemmas~\ref{lem-est-tG} and~\ref{lem-est-ediff}}\label{S:lemmas}
It remains to prove the lemmas that we used in the preceding section.
\begin{proof}[\textbf{Proof of Lemma~\ref{lem-est-tG}}]
We need to show that for each large $M$ there is a constant $C_1$ so that
\[
\int_0^t \int_\mathbb{R} |\tilde{\mathcal{G}}(x,y,t-s)| \left[\theta_2^2 + (1+s)^{\gamma-1/2} \theta_1\theta_2 + (1+s)^{\gamma} \theta_1 \rme^{-c^2 s/M} \right](y,s) \rmd y \rmd s \leq C_1 \theta_1(x,t)
\]
for all $t\geq0$. First, we note that there are constants $C_1,\tilde{C}_1>0$ such that
\[
\tilde{C}_1 \rme^{-y^2/M} \leq |\theta_1(y,s)| + |\theta_2(y,s)| \leq C_1 \rme^{-y^2/M}
\]
for all $0\leq s\leq t\leq1$. Thus, for some constant $C_1$ that may change from line to line, we have
\begin{eqnarray*}
\lefteqn{\int_0^t \int_\mathbb{R} |\tilde{\mathcal{G}}(x,y,t-s)|\left[\theta_2^2 + (1+s)^{\gamma-1/2}\theta_1\theta_2+ (1+s)^{\gamma} \theta_1\rme^{-\frac{c^2 s}{M}}\right](y,s) \rmd y \rmd s}
\\ & \leq &
C_1 \int_0^t \int_\mathbb{R} (t-s)^{-1/2} \rme^{-\frac{(x-y)^2}{4(t-s)}} \rme^{-\frac{y^2}{M}} \rmd y\rmd s
\\ & \leq &
C_1 \int_0^t \left[\int_{\{|y|\ge 2|x|\}} (t-s)^{-1/2} \rme^{-\frac{(x-y)^2}{8(t-s)}} \rme^{-\frac{x^2}{8(t-s)}} \rmd y+ \int_{\{|y|\le 2|x|\}} (t-s)^{-1/2} \rme^{-\frac{(x-y)^2}{4(t-s)}} \rme^{-\frac{4x^2}{M}} \rmd y\right]\rmd s
\\ & \leq &
C_1 \int_0^t \Big[ \rme^{-\frac{x^2}{8(t-s)}}+ \rme^{-\frac{4x^2}{M}} \Big]\rmd s
\\ & \leq &
C_1 \rme^{-\frac{4x^2}{M}}
\\ & \leq &
\frac{C_1}{\tilde{C}_1} \theta_1(x,t)
\end{eqnarray*}
for all $0\leq t\leq1$. An analogous computation can be carried out for the $x$-derivative since $\int_0^t (t-s)^{-1/2}\rmd s$ is bounded uniformly in $0\leq t\leq1$.
Thus, it remains to estimate the expression
\[
\theta_1(x,t)^{-1} \int_0^t \int_\mathbb{R} |\tilde{\mathcal{G}}(x,y,t-s)| \left[ \theta_2^2 + (1+s)^{\gamma-1/2}\theta_1\theta_2 + (1+s)^{\gamma} \theta_1\rme^{-c^2 s/M}\right](y,s) \rmd y \rmd s
\]
for $t\geq1$. Combining only the exponentials in this expression, we obtain terms that can be bounded by
\begin{equation}\label{exp-form}
\exp\left(\frac{(x+\alpha_3 ct)^2}{M(1+t)} - \frac{(x-y+\alpha_1 c(t-s))^2}{4(t-s)} - \frac{(y+\alpha_2 cs)^2}{M(1+s)} \right)
\end{equation}
with $\alpha_j=\pm c$. To estimate this expression, we proceed as in \cite[Proof of Lemma~7]{HowardZumbrun06} and complete the square of the last two exponents in (\ref{exp-form}). Written in a slightly more general form, we obtain
\begin{eqnarray*}
\lefteqn{\frac{(x-y-\alpha_1(t-s))^2}{M_1(t-s)} + \frac{(y-\alpha_2 s)^2}{M_2(1+s)} \;=\;
\frac{(x-\alpha_1(t-s)-\alpha_2 s)^2}{M_1(t-s)+M_2(1+s)} } \\ &&
+ \frac{M_1(t-s)+M_2(1+s)}{M_1M_2(1+s)(t-s)}\left( y - \frac{xM_2(1+s) - (\alpha_1M_2(1+s) + \alpha_2M_1s)(t-s)}{M_1(t-s)+M_2(1+s)}\right)^2
\end{eqnarray*}
and conclude that the exponent in (\ref{exp-form}) is of the form
\begin{eqnarray}\label{exp-form-2}
\lefteqn{ \frac{(x+\alpha_3t)^2}{M(1+t)} - \frac{(x-\alpha_1(t-s)-\alpha_2s)^2}{4(t-s)+M(1+s)} } \\ \nonumber &&
- \frac{4(t-s)+M(1+s)}{4M(1+s)(t-s)}\left( y - \frac{xM(1+s) - (\alpha_1M(1+s) + 4\alpha_2s)(t-s)}{4(t-s)+M(1+s)}\right)^2,
\end{eqnarray}
with $\alpha_j=\pm c$. Using that the maximum of the quadratic polynomial $\alpha x^2+\beta x+\gamma$ is $-\beta^2/(4\alpha)+\gamma$, it is easy to see that the sum of the first two terms in (\ref{exp-form-2}), which involve only $x$ and not $y$, is less than or equal to zero. Omitting this term, we therefore obtain the estimate
\begin{eqnarray}\label{est-on-exp}
\lefteqn{ \exp\left( \frac{(x\pm ct)^2}{M(1+t)} - \frac{(x-y\delta_1c(t-s))^2}{4(t-s)} - \frac{(y-\delta_2cs)^2}{M(1+s)} \right) } \\ \nonumber & \leq &
\exp \left( - \frac{4(t-s)+Ms}{4M(1+s)(t-s)}\left( y - \frac{xM(1+s)+c(\delta_1M(1+s)+ 4\delta_2s)(t-s)}{4(t-s)+M(1+s)}\right)^2 \right)
\end{eqnarray}
for $\delta_j=\pm1$. Using this result, we can now estimate the integral (\ref{exp-form}) term by term using the key assumption that $0<\gamma<\frac12$. The term involving $\theta_2^2$ can be estimated as follows using (\ref{est-on-exp}):
\begin{eqnarray*}
\lefteqn{ \theta_1(x,t)^{-1} \int_0^t \int_\mathbb{R} |\tilde{\mathcal{G}}(x,y,t-s)|\theta^2_2(y,s) \rmd y \rmd s } \\ & \leq &
C_1 (1+t)^{\gamma}\int_0^t \frac{1}{\sqrt{t-s}(1+s)^{1+2\gamma}} \\ && \times
\int_\R \exp\left(- \frac{4(t-s)+M(1+s)}{4M(1+s)(t-s)}\left( y - \frac{[xM(1+s) \pm c(M(1+s) + 4s)(t-s)]}{4(t-s)+M(1+s)}\right)^2 \right) \rmd y\rmd s \\ & \leq &
C_1(1+t)^{\gamma}\int_0^t \frac{1}{\sqrt{t-s}(1+s)^{1+2\gamma}} \sqrt{\frac{4M(1+s)(t-s)}{4(t-s)+M(1+s)}} \rmd s \\ & \leq &
C_1(1+t)^{\gamma}\int_0^{t/2} \frac{1}{(1+s)^{1/2+2\gamma}}\frac{1}{(1+t)^{1/2}} \rmd s + C_1(1+t)^{\gamma}\int_{t/2}^t \frac{1}{(1+s)^{1+2\gamma}}\rmd s \\ & \leq &
C_1(1+t)^{\gamma-1/2}+ C_1(1+t)^{-\gamma},
\end{eqnarray*}
which is clearly bounded since $\gamma<\frac12$. Similarly, we have
\begin{eqnarray*}
\lefteqn{\theta_1(x,t)^{-1} \int_0^t \int_\mathbb{R} |\tilde{\mathcal{G}}(x,y,t-s)|(1+s)^{\gamma-1/2}\theta_1\theta_2 (y,s) \rmd y \rmd s } \\ & \leq &
C_1(1+t)^{\gamma}\int_0^t \frac{1}{\sqrt{t-s}(1+s)^{\gamma+1}} \sqrt{\frac{4M(1+s)(t-s)}{4(t-s)+M(1+s)}} \rmd s \\ & \leq &
C_1(1+t)^{\gamma}\int_0^{t/2} \frac{1}{(1+s)^{\gamma+1/2}}\frac{1}{(1+t)^{1/2}} \rmd s + C_1(1+t)^{\gamma}\int_{t/2}^t \frac{1}{(1+t)^{\gamma+1/2}} \frac1{(1+s)^{1/2}}\rmd s
\\ & \leq &
C_1(1+t)^{\gamma-1/2} + C_1,
\end{eqnarray*}
which is again bounded due to $\gamma<\frac12$. Finally, we estimate
\begin{eqnarray*}
\lefteqn{\theta_1(x,t)^{-1} \int_0^t \int_\mathbb{R} |\tilde{\mathcal{G}}(x,y,t-s)| (1+s)^{\gamma} \rme^{-\frac{c^2 s}{M}}\theta_1(y,s) \rmd y \rmd s} \\ \nonumber & \leq &
C_1(1+t)^{\gamma}\int_0^t \frac{\rme^{-\frac{c^2 s}{M}}}{\sqrt{t-s}} \sqrt{\frac{4M(1+s)(t-s)}{4(t-s)+M(1+s)}} \rmd s \\ \nonumber & \leq &
C_1(1+t)^{\gamma}\int_0^{t/2}\rme^{-\frac{c^2 s}{M}}\frac{1}{(1+t)^{1/2}} \rmd s + C_1(1+t)^{\gamma} \rme^{-\frac{c^2 t}{2M}} \int_{t/2}^t\rmd s \\ \nonumber & \leq &
C_1(1+t)^{\gamma-1/2} + C_1(1+t)^{\gamma+1} \rme^{-\frac{c^2 t}{2M}},
\end{eqnarray*}
which is bounded, again due to $\gamma<\frac12$.
It remains to verify the second inequality in Lemma~\ref{lem-est-tG} which involves $\tilde{\mathcal{G}}_x$. We shall check only the term involving $(1+s)^{\gamma-1/2}\theta_1\theta_2$ as the other cases are similar and, in fact, easier. We have shown above that the resulting integrals are bounded for $0\leq t\leq1$ and therefore focus on the case $t\geq1$. Using that $|\tilde{\mathcal{G}}_x|\leq Ct^{-1/2}|\tilde{\mathcal{G}}|$, which follows by inspection, and employing again (\ref{est-on-exp}), we obtain
\begin{eqnarray*}
\lefteqn{\theta_2(x,t)^{-1} \int_0^t \int_\mathbb{R} |\tilde{\mathcal{G}}_x(x,y,t-s)|(1+s)^{\gamma-1/2}\theta_1\theta_2(y,s) \rmd y \rmd s} \\ & \leq &
C_1(1+t)^{\gamma+1/2}\int_0^t \frac{1}{(t-s)(1+s)^{\gamma+1}} \sqrt{\frac{4M(1+s)(t-s)}{4(t-s)+M(1+s)}} \rmd s \\ & \leq &
C_1(1+t)^{\gamma+1/2}\int_0^{t/2} \frac{1}{t^{1/2}(1+s)^{\gamma+1/2}}\frac{1}{(1+t)^{1/2}}\rmd s + C_1\int_{t/2}^t\frac{1}{(t-s)^{1/2}} \frac{1}{(1+t)^{1/2}}\rmd s
\\ & \leq &
C_1(1+t)^{\gamma}t^{-1/2} + C_1,
\end{eqnarray*}
which is bounded for $t\geq1$. This completes the proof of Lemma~\ref{lem-est-tG}.
\end{proof}
\begin{proof}[\textbf{Proof of Lemma~\ref{lem-est-ediff}}]
We need to show that
\[
\int_0^t \int_\mathbb{R} |e(x,t-s+1) - e(x,t+1)|\psi(y) [v_y^2 +\mathcal{N}(\cdot,\cdot,p,\dot{p},v_y)](y,s) \rmd y \rmd s \leq C_1 (\epsilon + h(t)^2)\theta_1(x,t).
\]
Intuitively, this integral should be small for the following reason. The difference $e(x,t-s)-e(x,t+1)$ converges to zero as long as $s$ is not too large, say on the interval $s\in[0,t/2]$. For $s\in[t/2,t]$, on the other hand, we will get exponential decay in $s$ from the localization of $\psi(y)$ in combination with the propagating Gaussians that appear in the nonlinearity and forcing terms. To make this precise, we write
\begin{equation}\label{e-difference}
e(x,t-s) - e(x,t+1) = \underbrace{e(x,t-s) - e(x,t-s+1)}_{\mathrm{term~I}} + \underbrace{e(x,t-s+1) - e(x,t+1)}_{\mathrm{term~II}}.
\end{equation}
We focus first on the term I and consider the cases $t\geq1$ and $0\leq t\leq1$ separately. First, let $t\geq1$. For $0\leq s\leq t-1$, we have $|e(x,t-s)-e(x,t-s+1)|\leq C\tilde{\mathcal{G}}(x,0,t-s)$, and we can estimate the resulting integral above in the same way as in the proof of Lemma~\ref{lem-est-tG}; we omit the details. For $t-1\leq s\leq t\leq1$, on the other hand, the definition of $e(x,t-s)$ yields
\begin{equation}\label{est-t-small}
|e(x,t-s) - e(x,t-s+1)| \leq C \int_{\frac{x^2}{(1+t-s)}}^{\frac{x^2}{(t-s)}} \rme^{-z^2} \rmd z \leq C \rme^{-x^2/2}.
\end{equation}
Using (\ref{est-phiN})-(\ref{est-phiN3}), namely
\begin{equation}\label{est-N13}
|\psi(y)\mathcal{N}(y,s,p,\dot{p},v_y)| \leq C\rme^{-\frac{c}{2}|y| - \frac{2c^2}{M}s} (\epsilon + h(t)^2)
\end{equation}
for all $s\geq0$, we obtain
\begin{eqnarray*}
\lefteqn{\int_{t-1}^t \int_\mathbb{R} [e(x,t-s) - e(x,t-s+1)]\psi(y)|\mathcal{N}(y,s,p,\dot{p},v_y)|(y,s)\rmd y \rmd s} \\
& \leq & C_1 (\epsilon + h(t)^2) \int_{t-1}^t \rme^{-\frac{x^2}{2}}\rme^{-\frac{2c^2 s}{M}}\rmd s
\;\leq\; C_1 (\epsilon + h(t)^2) \rme^{-\frac{x^2}{2}}\rme^{-\frac{2c^2 t}{M}},
\end{eqnarray*}
which is clearly bounded by $C_1\theta_1(x,t)$ since $\rme^{-c^2 t/M}\leq C_1(1+t)^{-\gamma}$ and
\[
\frac {(x+ ct)^2}{M(1+t)} \leq \frac{2x^2}{M} + \frac{4c^2}{M} + \frac{c^2t}{M}
\]
for arbitrary $M\geq4$. In summary, we have established the desired estimates for the term I in (\ref{e-difference}) for $t\geq1$. For $t\leq1$, the estimate (\ref{est-t-small}) remains true since $t-s$ is small, and proceeding as above yields
\[
\int_{0}^t \int_\mathbb{R} [e(x,t-s) - e(x,t-s+1)]\psi(y) |\mathcal{N}(y,s,p,\dot{p},v_y)| \rmd y \rmd s \leq C(\epsilon + h(t)^2) \rme^{-\frac{x^2}{2}}\rme^{-\frac{2c^2 t}{M}},
\]
which is again bounded by $C_1\theta_1(x,t)$.
It remains to discuss the term II which involves the difference $e(x,t-s+1)-e(x,t+1)$. We have
\begin{eqnarray*}
\lefteqn{|e(x,t-s+1) - e(x,t+1)|} \\ & = &
\left|\int_{t+1}^{t-s+1} e_\tau(x,\tau)\rmd \tau\right| \\ & \leq &
\int_{t-s+1}^{t+1} \left| \frac{c}{\sqrt{4\pi \tau}} \left( \rme^{-\frac{(x-c\tau)^2}{4\tau}} + \rme^{-\frac{(x+c\tau)^2}{4\tau}}\right) + \frac{1}{\tau\sqrt{4\pi}} \left( \frac{(x-c\tau)}{\sqrt{4\tau}}\rme^{-\frac{(x-c\tau)^2}{4\tau}} - \frac{(x+c\tau)}{\sqrt{4\tau}}\rme^{-\frac{(x+c\tau)^2}{4\tau}}\right) \right|\rmd \tau
\\ & \leq &
C\int_{t-s+1}^{t+1} \left( \frac{1}{\sqrt{\tau}} + \frac{1}{\tau} \right) \left( \rme^{-\frac{(x-c\tau)^2}{8\tau}} + \rme^{-\frac{(x+c\tau)^2}{8\tau}}\right) \rmd \tau,
\end{eqnarray*}
where we used in the last inequality that $z\rme^{-z^2}$ is uniformly bounded in $z$. We now use the preceding expression to estimate $\theta_1^{-1}(x,t)(e(x,t-s+1)-e(x,t+1))$ and focus first on the single exponential term
\[
\rme^{\frac{(x-ct)^2}{M(1+t)}}\rme^{-\frac{(x-c\tau)^2}{8\tau}}.
\]
Combining these exponentials and completing the square in $x$ in the resulting exponent, the latter becomes
\[
-\frac{[M(t-\tau+1) + (M-8)\tau]}{8M(t+1)\tau}\left[ x + \frac{c(8-M)\tau(t+1)}{M(t-\tau) + (M-8)\tau}\right]^2 + \frac{c^2(t-\tau+1)^2}{M(t-\tau+1) + (M-8)\tau}.
\]
Using that $\tau\leq t$ and picking $M\geq8$, we can neglect the exponent resulting from the first expression that involves in $x$ and conclude that
\[
\rme^{\frac{(x-ct)^2}{M(1+t)}}\rme^{-\frac{(x-c\tau)^2}{8\tau}} \leq C_1 \rme^{\frac{c^2(t-\tau)}{M}}.
\]
The remaining exponentials can be estimated similarly, and we obtain
\begin{eqnarray*}
\theta_1^{-1}(x,t) |e(x,t-s+1) - e(x,t+1)| & \leq &
C_1 (1+t)^{\gamma}\int_{t-s+1}^{t+1} \left( \frac{1}{\sqrt{\tau}} + \frac{1}{\tau} \right) \rme^{\frac{c^2(t-\tau)}{M}} \rmd \tau \\ \nonumber & \leq &
C_1 (1+t)^{\gamma} (1+t-s)^{-1/2}\rme^{\frac{c^2s}{M}}.
\end{eqnarray*}
Using this inequality together with (\ref{est-N13}) finally gives
\begin{eqnarray*}
\lefteqn{\theta_1(x,t)^{-1}\int_0^t \int_\mathbb{R} [e(x,t-s+1) - e(x,t+1)]\psi(y)|\mathcal{N}(y,s,p,\dot{p},v_y)|(y,s) \rmd y \rmd s} \\ & \leq &
C_1(1+t)^{\gamma} (\epsilon + h(t)^2)\int_0^t (1+t-s)^{-1/2}\rme^{\frac{c^2s}{M}} \rme^{-\frac{2c^2 s}{M}}\rmd s \\ & \leq &
C_1(1+t)^{\gamma} (\epsilon + h(t)^2)\left[(1+t)^{-1/2}\int_0^{t/2} \rme^{-\frac{c^2 s}{M}}\rmd s + \rme^{-\frac{c^2 t}{2M}}\int_{t/2}^t (1+t-s)^{-1/2}\rmd s\right] \\ & \leq &
C_1(\epsilon + h(t)^2).
\end{eqnarray*}
for $M$ sufficiently large, which proves the first estimate in Lemma~\ref{lem-est-ediff}.
It remains to prove the estimate
\begin{eqnarray*}
&& \int_0^t \int_\mathbb{R} |e_x(x,t-s+1) - e_x(x,t+1)|\psi(y)\left[v_y^2 +\mathcal{N}(\cdot,\cdot,p,\dot{p},v_y)\right](y,s) \rmd y \rmd s \\ && \qquad\qquad \leq C_1(\epsilon + h(t)^2)\theta_2(x,t)
\end{eqnarray*}
for the derivative in $x$. Since the derivative of $e(x,t-s+1)-e(x,t+1)$ with respect to $x$ generates an extra decay term $(1+t)^{-1/2}$, we have
\begin{eqnarray*}
\lefteqn{ \theta_2(x,t)^{-1}\int_0^t \int_\mathbb{R} [e_x(x,t-s+1) - e_x(x,t+1)]\psi(y) |\mathcal{N}(y,s,p,\dot{p},v_y)|(y,s) \rmd y \rmd s } \\ & \leq &
C_1 (\epsilon ^2 + h(t)^2)(1+t)^{\gamma+1/2}\int_0^t (1+t-s)^{-1}\rme^{\frac{c^2s}{M}} \rme^{-\frac{2c^2 s}{M}} \rmd s \\ & \leq & C_1 (\epsilon ^2 + h(t)^2)(1+t)^{\gamma+1/2}\left[(1+t)^{-1}\int_0^{t/2}\rme^{-\frac{c^2 s}{M}} \rmd s + \rme^{-\frac{c^2 t}{2M}}\int_{t/2}^t(1+t-s)^{-1}\rmd s\right] \\ & \leq & C_1 (\epsilon ^2 + h(t)^2),
\end{eqnarray*}
which completes the proof of the lemma.
\end{proof}
\paragraph{Acknowledgments.}
Beck, Nguyen, Sandstede, and Zumbrun were supported partially by the NSF through grants DMS-1007450, DMS-1108821, DMS-0907904, and DMS-0300487, respectively.
\bibliographystyle{sandstede}
|
\section{Introduction}
In $\Pp\Pp$ collisions, $\PW$ bosons are produced primarily via the
processes
$\mathrm{u}\mathrm{\bar{d}}\rightarrow\PWp$ and $\mathrm{d}\mathrm{\bar{u}}\rightarrow\PWm$.
The first quark is a valence quark from one of the protons, and the second one is a sea antiquark from
the other proton. Due to the presence of two valence $\mathrm{u}$ quarks in the proton,
there is an overall excess of $\PWp$ over $\PWm$ bosons.
The inclusive ratio of cross sections for $\PWp$ and $\PWm$ bosons production at the Large
Hadron Collider~(LHC)
was measured
to be $1.43 \pm 0.05$ by the Compact Muon Solenoid~(CMS) experiment~\cite{CMS:WZ}
and is in agreement with predictions of the Standard Model~(SM) based on various
parton distribution functions~({\sc PDF}s)~\cite{MSTW:0901, CTEQ:1007}.
Measurement of this production asymmetry
between $\PWp$ and $\PWm$ bosons as a function of boson rapidity
can provide new insights on the $\mathrm{u}/\mathrm{d}$ ratio and the sea antiquark densities in
the ranges of the Bj\"{o}rken parameter $x$~\cite{BJORKEN} probed in $\Pp\Pp$
collisions at $\sqrt{s}=7\TeV$.
However, due to the presence of neutrinos in leptonic $\PW$ decays the boson rapidity is not directly accessible.
The experimentally accessible quantity
is the
lepton charge asymmetry, defined to b
\def\ensuremath{\mathrm{d}\sigma/\mathrm{d}\eta}{\ensuremath{\mathrm{d}\sigma/\mathrm{d}\eta}}
$$
\mathcal{A}(\eta) = \frac{\ensuremath{\mathrm{d}\sigma/\mathrm{d}\eta}(\PWp\rightarrow\ell^+\nu)-\ensuremath{\mathrm{d}\sigma/\mathrm{d}\eta}(\PWm\rightarrow\ell^-\bar{\nu})}
{\ensuremath{\mathrm{d}\sigma/\mathrm{d}\eta}(\PWp\rightarrow\ell^+\nu)+\ensuremath{\mathrm{d}\sigma/\mathrm{d}\eta}(\PWm\rightarrow\ell^-\bar{\nu})},
$$
where $\ell$ is the daughter charged lepton, $\eta$ is the charged lepton
pseudorapidity in the
CMS lab frame ($\eta= -\ln{[\tan{(\frac{\theta}{2})}]}$ where $\theta$ is the polar angle), and
$\mathrm{d}\sigma/\mathrm{d}\eta$ is the differential cross section for charged
leptons from $\PW$ boson decays.
The lepton charge
asymmetry can be used to test SM predictions with high
precision.
Due to the $\mathrm{V}-\mathrm{A}$ structure of the $\PW$ boson couplings to fermions,
theoretical predictions of the charge asymmetry
depend on the transverse momentum~($\pt$) threshold applied on the daughter leptons.
For this reason, we measure $\mathcal{A}(\eta)$ for two different charged lepton $\pt$~($\pt^{\ell}$) thresholds,
25\GeVc and 30\GeVc.
The lepton charge asymmetry and the $\PW$ charge
asymmetry have been studied in $\mathrm{p}\bar{\mathrm{p}}$ collisions by both
the CDF and D0 experiments at the Fermilab Tevatron Collider~\cite{CDF:wasym, D0:asym}.
Current predictions for the lepton charge
asymmetry at the LHC based on different {\sc PDF} models do not
agree well with each other.
A high precision measurement of this asymmetry
at the LHC
can contribute to the improvement of the knowledge of {\sc PDF}s.
The
measurement of the muon charge asymmetry at the LHC with a dataset
corresponding to an integrated luminosity of about 31 $\mathrm{pb}^{-1}$ was
reported recently by the ATLAS experiment~\cite{ATLAS:asym}. In this Letter,
we present a measurement of the lepton charge asymmetry in both
$\mathrm{W} \rightarrow \mathrm{e}\nu$
and $\mathrm{W} \rightarrow \mu\nu$ final states using a
dataset corresponding to an integrated luminosity of 36
$\mathrm{pb}^{-1}$
collected by the CMS detector in March--November 2010.
\section{The CMS Detector}
A detailed description of the CMS experiment can be found
elsewhere~\cite{refCMS}.
The central feature of the CMS apparatus is a
superconducting solenoid, of 6~m internal diameter, 13~m in length,
providing an axial field of
3.8~T. Within the field volume are the silicon pixel and strip tracker,
the crystal electromagnetic calorimeter (ECAL) and the brass/scintillator
hadron calorimeter (HCAL). Muons are measured in gas-ionization detectors
embedded in the steel return yoke of the solenoid.
The most relevant sub-detectors for this measurement are the ECAL,
the muon system, and the tracking system. The electromagnetic calorimeter consists
of nearly 76\,000 lead tungstate crystals which provide coverage in pseudorapidity $|\eta| <$ 1.479 in the
barrel region and 1.479 $< |\eta| <$ 3.0 in two endcap regions.
A preshower detector consisting of two planes of silicon sensors interleaved with a total of 3$\,X_0$ of lead is located in front of the ECAL endcaps.
The ECAL has an ultimate energy resolution of better than 0.5\% for
unconverted photons with transverse energies above 100\GeV.
The electron energy
resolution is 3\% or better for the range of electron energies relevant for
this analysis.
Muons are measured in the pseudorapidity range $|\eta| <$ 2.4,
with detection planes made of three technologies: drift tubes, cathode strip chambers, and resistive plate chambers. Matching the muons to the tracks measured in the silicon tracker results in a transverse momentum resolution of about
2\% in the relevant muon $\pt$ range.
CMS uses a right-handed coordinate system, with the origin at the nominal interaction point, the $x$-axis pointing to the center of the LHC, the $y$-axis pointing up (perpendicular to the LHC plane), and the $z$-axis along the anticlockwise-beam direction. The polar angle, $\theta$, is measured from the positive $z$-axis and the azimuthal angle, $\phi$, is measured in the $x$-$y$ plane.
\section{Analysis Method and Simulation}
The $\mathrm{W} \rightarrow \ell\nu$ candidates are characterized by a
high-$\pt$
lepton accompanied by missing transverse energy $\Em_{\mathrm{T}}$, due
to the escaping neutrino.
Experimentally, $\Em_{\mathrm{T}}$ is determined as
the negative vector sum of the
transverse momenta of
all particles reconstructed using a particle flow algorithm~\cite{pfmet}.
The $\mathrm{W} \rightarrow \mathrm{e}\nu$ and
$\mathrm{W} \rightarrow \mu\nu$ candidates used in this analysis were
collected using a set of inclusive
single-electron and single-muon triggers which did not include $\Em_{\mathrm{T}}$
requirements.
Other physics processes, such as multijet and photon+jet
production~(QCD background),
Drell--Yan~($ \mathrm{Z}/\gamma^*\rightarrow \ell^+ \ell^-$) production, $\mathrm{W} \rightarrow \tau\nu$ production~(EWK background),
and top quark pair ($ \mathrm{t}\bar{\mathrm{t}}$) production can produce
high-$\pt$ electron/muon candidates
and mimic
$\PW$ candidates. Furthermore,
cosmic ray muons can produce fake
$\mathrm{W} \rightarrow \mu\nu$ candidates.
The muon measurement relies on the muon detector, inner tracking and
calorimeters
to form a
robust isolation
variable
to separate signal from background.
This method is not applicable to the electron measurement because electron candidates are accompanied by significant electromagnetic radiation, reducing the power of the isolation in extracting signal from background.
Instead
the electron selection relies on
the calorimeter system and uses the $\Em_{\mathrm{T}}$ to
separate signal from background. These
two measurements are largely independent and cross-check each other.
Monte Carlo~(MC) simulation samples have been
used to develop analysis techniques and estimate some of the background contributions.
The next-to-leading order~(NLO) MC
simulations based on the {\sc POWHEG} event generator~\cite{POWHEG}
interfaced with the CT10 PDF model~\cite{CTEQ:1007} are primarily used.
The QCD multijet background
is generated with the {\sc PYTHIA} event generator~\cite{PYTHIA6} interfaced with
the CTEQ6L PDF model~\cite{CTEQ6L}.
The $ \mathrm{t}\bar{\mathrm{t}}$ background is generated using both {\sc PYTHIA} and
{\sc MC@NLO}~\cite{MC@NLO} event generators.
Other {\sc PYTHIA}-based MC simulations
are used to cross-check MC predictions and help assigning experimental
systematic uncertainties. The {\sc PYTHIA}-based MC samples are normalized using NLO cross sections except for the QCD background samples.
All generated events are passed
through the CMS detector simulation using {\sc GEANT4}~\cite{GEANT4}
and then processed using a reconstruction sequence identical to that used for
collision data.
Pile-up, which consists of the presence
of secondary minimum bias interactions in addition to the primary hard
interaction in an event,
is significant in the data, where
an average of 2--3 vertices are found.
The analysis methods used in both electron and muon channels are
insensitive to the effect of pile-up.
The selection criteria for electron/muon reconstruction and identification
are almost identical to those used in the $\PW$ and
$\mathrm{Z}$ cross section measurements~\cite{CMS:WZ}. A brief summary is given here
for completeness.
\section{Electron Reconstruction and $\mathrm{W} \rightarrow \mathrm{e}\nu$ Signal Extraction}
Electrons are identified as clusters of energy deposited in the ECAL fiducial
volume matched to tracks from the inner silicon tracker~({\it silicon tracks}).
The silicon tracks are
reconstructed using a Gaussian-Sum-Filter~(GSF)
algorithm~\cite{GSF} that takes into account possible energy loss
due to bremsstrahlung in the tracker layers. Particles misidentified as
electrons are suppressed by requiring that the shower shape of the
ECAL cluster
be consistent with an electron candidate,
and that the $\eta$ and $\phi$ coordinates
of the track trajectory extrapolated to the ECAL match the $\eta$ and $\phi$
coordinates of the ECAL cluster.
Furthermore, electrons from $\PW$ decay are isolated from other activity
in the event.
We therefore require that little transverse energy be observed in the
ECAL, HCAL, and silicon tracking system within a cone
$\Delta R <$ 0.3 around the electron direction, where $\Delta R=\sqrt{(\Delta\phi)^2+(\Delta\eta)^2}$ and where calorimeter energy deposits and the track
associated with the electron candidate are excluded.
Due to
the substantial amount of material in front of the ECAL detector,
a large fraction of electrons radiate photons. The resulting photons
may convert close to the original electron trajectory, leading to
a sizable charge misidentification rate~($w$).
The true charge asymmetry, $\mathcal{A}^{\textrm{true}}$, is diluted
due to charge misidentification resulting in an observed asymmetry, $\mathcal{A}^{\textrm{obs}}= \mathcal{A}^{\textrm{true}}(1-2w).$
Three different methods are used to
determine the electron charge. First, the electron charge
is determined by the signed curvature of the
associated GSF track. Second, the charge is determined from the
associated trajectory reconstructed in the silicon tracker
using a Kalman Filter algorithm~\cite{KF}. Third,
the electron charge is determined based on the azimuthal angle between
the vector joining
the nominal interaction point and the ECAL cluster position and the vector
joining the nominal interaction point and innermost hit of the GSF track. It is required that all
three charge determinations from these methods agree.
This procedure significantly reduces the charge misidentification rate
to 0.1\% in the ECAL barrel and to 0.4\%
in the ECAL endcaps.
The $\mathrm{W} \rightarrow \mathrm{e}\nu$ candidates are selected by further
requiring electrons to have $\pt >$ 25\GeVc, $|\eta| <$ 2.4, and
to be associated with one of the electron
trigger candidates used to select the electron dataset. The Drell--Yan
and $ \mathrm{t}\bar{\mathrm{t}}$ backgrounds are suppressed
by rejecting events that contain a second isolated
electron or muon with $\pt >$ 15\GeVc and $|\eta| <$ 2.4.
According to MC simulations,
the data sample of selected electrons
consists of
about 28\% QCD background events,
about
6.5\% EWK background
events,
and about 0.2\% $ \mathrm{t}\bar{\mathrm{t}}$ background events.
The events passing the above selection criteria are divided into six bins of electron
pseudorapidity ($|\eta^\textrm{e}|$): [0.0,\ 0.4],\ [0.4,\ 0.8],\ [0.8,\ 1.2],\
[1.2,\ 1.4],\ [1.6,\ 2.0],\ and\ [2.0,\ 2.4],
for the measurement of the electron charge asymmetry.~(The fourth bin is
reduced to a width of 0.2 in order to
exclude the transition region between the ECAL barrel and endcaps.)
A binned extended maximum
likelihood fit is performed
over the $\Em_{\mathrm{T}}$ distribution to estimate the $\mathrm{W} \rightarrow \mathrm{e}\nu$ signal
yield for electrons~($N^-$) and positrons~($N^+$) in each pseudorapidity bin.
The signal $\Em_{\mathrm{T}}$ shape is derived
from MC simulations with an event-by-event correction to account for energy
scale and resolution differences between data and MC based on
the hadronic recoil energy distributions in $ \mathrm{Z}/\gamma^*\rightarrow \mathrm{e}^+ \mathrm{e}^- $ events selected
from data~\cite{RECOIL}.
The shape of the QCD background is
determined, for each charge, from a signal-free control sample obtained by
inverting a subset of the electron identification criteria.
The $\Em_{\mathrm{T}}$ shapes for
other backgrounds such as the Drell--Yan process,
$ \mathrm{t}\bar{\mathrm{t}}$,
and $\mathrm{W} \rightarrow \tau\nu$ are taken from MC simulations with
a fixed normalization relative to the $\mathrm{W} \rightarrow \mathrm{e}\nu$
yields. The normalization factors are from
the predicted values of the cross sections at NLO.
The yield of the QCD background and the
yield of the $\mathrm{W} \rightarrow \mathrm{e}\nu$ signal are free parameters
in the fit. The results of the fits to the data
are shown for the first pseudorapidity bin in
Figs.~\ref{fig1}~(a) and~\ref{fig1}~(b).
The charge asymmetry is
obtained from
$(N^+-N^-)/(N^++N^-)$.
\begin{figure}[tb]
\begin{center}
\includegraphics[width=0.95\columnwidth]{fig1.pdf}
\hspace{1cm}
\caption{Signal fit to data $\Em_{\mathrm{T}}$ distributions for electrons,
a) $\PWp \rightarrow \mathrm{e}^+\nu$,
b) $\PWm\rightarrow \mathrm{e}^-\bar{\nu}$, and fit to
$\xi$ distributions
for muons, c) $\PWp \rightarrow \mu^+\nu$, and d) $\PWm \rightarrow \mu^-\bar{\nu}$.
Only the results for
the first pseudorapidity bin ($|\eta| <$ 0.4) are shown. In Figs.~c) and d),
only events on the right of the dashed vertical line are included in the
fit. The QCD (multijet and photon+jet production) shape is determined directly from data.}
\label{fig1}
\end{center}
\end{figure}
\section{Muon Reconstruction and $\mathrm{W} \rightarrow \mu\nu$ Signal Extraction}
Muon candidates are reconstructed using two different algorithms: one starts
from inner silicon tracks
and
requires a minimum number of matching hits in the muon chambers,
and
the other finds tracks in the muon system and matches them to
silicon tracks.
A global track fit including both the silicon track hits
and muon chamber hits
is performed to improve the quality of the
reconstructed muon candidates.
The muon candidate is not required
in this analysis to be isolated from other event activity, a
difference in muon selection with respect to~\cite{CMS:WZ}, in order to avoid
bias in the signal extraction fit described below.
The muon charge is identified from the signed curvature of
the associated
silicon track.
A selection on
the silicon track distance of closest approach to the beam spot,
$|d_{xy}|<$ 0.2 $ \mathrm{cm}$, is applied to reduce the cosmic ray background.
The remaining cosmic ray background yield is estimated by
normalizing the $|d_{xy}|$ distribution derived from cosmic ray muon data to the
large $|d_{xy}|$ region~(1.0 $< |d_{xy}| <$ 5.0 $\mathrm{cm}$) in the data sample.
The
estimated
cosmic ray background contamination is
about $10^{-5}$
of the expected $W\rightarrow \mu\nu $ signal
yield
and
is neglected.
The
$\mathrm{W} \rightarrow \mu\nu$ candidates are selected by further
requiring the muon $\pt$ to be greater than 25\GeVc,
$|\eta| <$ 2.1, and that the candidate matches
one of the muon trigger candidates.
The Drell--Yan background
is suppressed by rejecting events which contain a
second isolated muon
with $\pt >$ 15\GeVc, $|\eta| <$ 2.4, and
passing the above muon quality selections.
The events which passed the above selection criteria are divided into six bins of muon
pseudorapidity ($|\eta^{\mu}|$):
[0.0,\ 0.4],\ [0.4,\ 0.8],\ [0.8,\ 1.2],\
[1.2,\ 1.5],\ [1.5,\ 1.8],\ and\ [1.8,\ 2.1].
The $\mathrm{W} \rightarrow \mu\nu$ signal estimation is done by fitting
the distribution of an isolation variable
$\xi=\Sigma(E_T)$
defined as the scalar sum of the transverse momenta
of silicon tracks (excluding the muon candidate)
and energy deposits in both ECAL and HCAL in a cone $\Delta R <$ 0.3 around the muon direction.
The calorimeter energy deposit associated with the muon candidate is rescaled
to ensure a uniform transverse energy response as a function of polar angle.
With this correction the shape of the
signal $\xi$ distribution
becomes independent of pseudorapidity, peaking at a constant $E_{\mathrm{T}}$
value of about 2.5\GeV.
The correction was obtained from muons in the $ \mathrm{Z}/\gamma^*\rightarrow \mu^+ \mu^- $ MC sample and checked with real muons from the $ \mathrm{Z}/\gamma^*\rightarrow \mu^+ \mu^- $ data sample.
The shape of the $\xi$ distribution for muons from $\mathrm{W} \rightarrow \mu\nu$
is parametrized as a Landau distribution
convolved with a Gaussian resolution function.
The tail of the Landau distribution is modified to be an exponential
function.
The $\mathrm{W} \rightarrow \mu\nu$ signal, the Drell--Yan background,
the $ \mathrm{t}\bar{\mathrm{t}}$ background, and the $\mathrm{W} \rightarrow \tau\nu$ background
have been shown in MC simulations to have $\xi$ distributions that are
identical and
lepton-charge independent.
The charge independence has been confirmed in studies of $ \mathrm{Z}/\gamma^*\rightarrow \mu^+ \mu^- $ data.
In the final fit the shape parameters for the signal $\xi$
distribution are fixed using
$ \mathrm{Z}/\gamma^*\rightarrow \mu^+ \mu^- $ data, except for the exponential
tail parameters which are fixed using $ \mathrm{Z}/\gamma^*\rightarrow \mu^+ \mu^- $ MC
simulations due to limited sample of $ \mathrm{Z}/\gamma^*\rightarrow \mu^+ \mu^- $ in data.
The QCD background is parametrized by an empirical function
$\xi^{\alpha}\cdot e^{\beta \sqrt{\xi}}$.
The QCD parametrization and the value of background parameter $\alpha$
is determined directly from data using a QCD background control sample obtained
by selecting events with
large impact parameter significance and little $\Em_{\mathrm{T}}$
in the event.
The background parameter $\beta$ is allowed to float.
The QCD yield is
determined separately
for negative and positive charges.
The region used for signal estimation is chosen to be
0 $< \xi <$ 25\GeV. From MC studies,
the expected
QCD multijet, EWK, and $ \mathrm{t}\bar{\mathrm{t}}$ backgrounds are
about 13.0\%, 6.9\%, and 0.3\%, respectively
within the interval
0 $< \xi <$ 10\GeV, in which most of $\mathrm{W} \rightarrow \mu\nu$ signal candidates are found.
An unbinned extended maximum likelihood
fit to the $\xi$ distribution
is performed simultaneously on the $\PWp \rightarrow \mu^+\nu$ and
$\PWm \rightarrow \mu^-\bar{\nu}$ candidates to determine
the total $\mathrm{W} \rightarrow \mu\nu$ signal yield and the charge asymmetry in
each pseudorapidity bin.
Background events from $ \mathrm{Z}/\gamma^*\rightarrow \mu^+ \mu^- $ are selected
if one of the daughter muons is outside of the detector acceptance.
This part of the $ \mathrm{Z}/\gamma^*\rightarrow \mu^+ \mu^- $ background is
normalized
to the observed $ \mathrm{Z}/\gamma^*\rightarrow \mu^+ \mu^- $
events in data obtained by inverting the Drell--Yan veto selection.
The acceptance ratio for one versus two muons from $ \mathrm{Z}/\gamma^*\rightarrow \mu^+ \mu^- $ events
is estimated with MC simulations.
There are small contributions to the
$ \mathrm{Z}/\gamma^*\rightarrow \mu^+ \mu^- $ background with both muons
staying
within the detector coverage but one of them
failing reconstruction, muon identification, or
isolation requirement. This background contribution
is estimated directly from data. The
$ \mathrm{Z}/\gamma^*\rightarrow \tau^+ \tau^- $ background is normalized to the
estimated $ \mathrm{Z}/\gamma^*\rightarrow \mu^+ \mu^- $
background with ratios determined from MC simulations.
The $\mathrm{W} \rightarrow \tau\nu$ background is
normalized to the $\mathrm{W} \rightarrow \mu\nu$ signal yield in the data using the
ratio between $\mathrm{W} \rightarrow \tau\nu$ background and $\mathrm{W} \rightarrow \mu\nu$ signal
yield determined from MC simulations. The $ \mathrm{t}\bar{\mathrm{t}}$ background is estimated
directly from
MC simulations with cross sections normalized to the predicted value at NLO.
The EWK and $ \mathrm{t}\bar{\mathrm{t}}$ backgrounds
are estimated
for the $\PWp \rightarrow \mu^+\nu$ and
$\PWm \rightarrow \mu^-\bar{\nu}$ candidates separately.
The results of the fits to the data
are shown for the first pseudorapidity bin in
Figs.~\ref{fig1}~(c) and~\ref{fig1}~(d).
Only the region $\xi >$ 0.8\GeV is included in the fit because
the small region $\xi <$ 0.8\GeV exhibits a complex shape for both
signal and QCD background due to the zero-suppressed readout of the CMS
calorimeters. However,
events in this region are included to determine the
$\mathrm{W} \rightarrow \mu\nu$ signal
yield and charge asymmetry, as the number of QCD background events in this
region is
negligible, confirmed by a Monte Carlo simulation study.
\section{Systematic Uncertainties}
The systematic uncertainties considered for both the electron and muon channels are mainly due to the
lepton charge misidentification rate, possible efficiency differences between the $\ell^+$
and $\ell^-$,
lepton momentum~(energy) scale and
resolution,
and signal estimation.
The electron charge misidentification rate is measured in data
using the $ \mathrm{Z}/\gamma^*\rightarrow \mathrm{e}^+ \mathrm{e}^- $ data sample to be within 0.1--0.4\%, increasing with electron pseudorapidity.
The measured electron charge asymmetry is corrected for the charge
misidentification rate.
The statistical error on the electron
charge misidentification rate is taken as the systematic
uncertainty.
The muon charge misidentification rate is studied using $\mathrm{W} \rightarrow \mu\nu$
MC simulations and is estimated to be at the level of
$10^{-5}$.
The muon charge misidentification rate is further studied with a cosmic ray muon data
sample in which one cosmic ray muon passes through the center of the CMS detector
and
is reconstructed as two muon candidates with opposite charge~\cite{COSMICS}.
A total of $16\,422$ cosmic ray muon events with at least two muon candidates
are selected,
and no event is found to
have a same sign muon pair. This constrains the muon
charge misidentification rate to be less than $10^{-4}$. Charge misidentification therefore
has a negligible
effect on the measured muon charge asymmetry.
The efficiency difference between $\ell^+$ and $\ell^-$
can result in a bias on the measured charge asymmetry.
The total lepton efficiency (including lepton reconstruction, identification,
and trigger efficiencies) in each pseudorapidity bin is measured
using the $ \mathrm{Z}/\gamma^*\rightarrow \ell^+ \ell^-$ data for $\ell^+$ and
$\ell^-$, respectively. The efficiency ratio is calculated and
found to be
consistent with unity within the statistical uncertainty.
No correction is made to the observed charge asymmetry. However,
the statistical errors on the efficiency ratios are treated
as systematic uncertainties.
The inclusive
electron efficiency ratio is found to be $1.007\pm0.014$, and the error on this
ratio is used to determine the systematic uncertainty for the electron
channel.
Within the statistical uncertainty the muon efficiency ratio is also constant across the
pseudorapidity coverage. However,
due to a small known
charge/pseudorapidity-dependent muon momentum scale bias, the statistical uncertainty on the
bin-by-bin efficiency ratio is used.
This is
the dominant systematic uncertainty for
both the electron and muon channels.
In order to compare our results directly
to theoretical predictions, the measured charge asymmetry in the electron (muon) channel is corrected
for lepton energy~(momentum)
bias and resolution
effects.
The lepton scale and resolution are determined directly from the $ \mathrm{Z}/\gamma^*\rightarrow \ell^+ \ell^-$
data and are used to smear MC lepton $\pt^{\ell}$ at the generator level. The correction to
the measured
charge asymmetry is estimated in each pseudorapidity bin
by comparing the charge asymmetry as determined in the MC
with the resulting
asymmetry after smearing.
The
uncertainties on the energy~(momentum) scale and
resolutions are taken as sources for
systematic uncertainties. The electron energy scale bias is studied using $ \mathrm{Z}/\gamma^*\rightarrow \mathrm{e}^+ \mathrm{e}^- $
data and determined to be
within 1\%. The charge-dependent muon momentum
scale bias is also studied using $ \mathrm{Z}/\gamma^*\rightarrow \mu^+ \mu^- $ data sample and
determined to be within 1\%.
This energy~(momentum) scale bias dominates the systematic uncertainty due to
lepton energy~(momentum) scale and resolution.
The impact of the QED final-state
radiation~(FSR)
on the lepton charge asymmetry is also studied using {\sc POWHEG} MC samples and a
reduction of the asymmetry at the level of 0.1--0.2\% is found.
The charge asymmetry is corrected for the FSR effect and
the full correction is taken as additional
systematic uncertainty which is summed in quadrature with
the lepton energy~(momentum) scale systematic uncertainty.
For the electron channel,
the dominant systematic uncertainty on the signal and background estimation is from the
modeling of the $\Em_{\mathrm{T}}$ shape for the signal.
Others, such as the uncertainties on the
background cross sections and QCD background shape modeling,
also contribute. The systematic uncertainty due to the QCD
background shape is studied by using different QCD background control regions
to derive the QCD background $\Em_{\mathrm{T}}$ shape.
For the muon channel,
the systematic sources for signal and background estimation are due to
signal and background parametrization, uncertainty in estimating
Drell--Yan and $ \mathrm{t}\bar{\mathrm{t}}$ background yields, and the
$\mathrm{W} \rightarrow \tau\nu$ background to $\mathrm{W} \rightarrow \mu\nu$ signal ratio.
The largest contribution is from
the estimation of Drell--Yan
background
which is dominated
by the limited number
of Drell--Yan dimuon events and variations
of the acceptance ratio for one versus two muons from $ \mathrm{Z}/\gamma^*\rightarrow \mu^+ \mu^- $ events.
This acceptance ratio depends on both PDF models and the $\mathrm{Z}$ boson
$\pt$ spectrum. The {\sc PYTHIA} and {\sc POWHEG} MC samples are used
to estimate the ratios, and differences between these two MC simulations~(at the level of 5--10\%) are treated as systematic uncertainties.
The $ \mathrm{t}\bar{\mathrm{t}}$
background yield is varied by 18.6\% to reflect
the theoretical uncertainty on the $ \mathrm{t}\bar{\mathrm{t}}$ cross section.
The ratio of
the $\mathrm{W} \rightarrow \tau\nu$
background to the $\mathrm{W} \rightarrow \mu\nu$ signal and ratio of
the $ \mathrm{Z}/\gamma^*\rightarrow \tau^+ \tau^- $
background to the $ \mathrm{Z}/\gamma^*\rightarrow \mu^+ \mu^- $ background
are estimated using both {\sc POWHEG} and {\sc PYTHIA} MC simulations.
The differences between these two MC simulations are taken as sources of
systematic uncertainties.
The robustness of the signal parametrization is studied using pseudo-experiments,
whose pseudo-data are
obtained by sampling fully simulated MC events according to Poisson distributions with means set to
the measured signal and background yields in data.
A small bias
on the measured charge asymmetry at the level of 0.1--0.2\% is
found and taken as additional systematic uncertainty
on the signal estimation.
\begin{table*}[tb]
\caption{Summary of the systematic uncertainties. All values are in percent.
}
\label{tab1}
\begin{center}
{\footnotesize
\begin{tabular*}{\textwidth}{@{\extracolsep{\fill}} l|rrrrrr|rrrrrr} \hline\hline
\multicolumn{13}{c}{\rule[-1.8mm]{0mm}{4mm} $p_{\mathrm{T}}^{\ell} >$ 25\GeVc \rule[-1.8mm]{0mm}{6.0mm}}\\
\cline{1-13}
&\multicolumn{6}{c|}{Electron Channel}& \multicolumn{6}{c}{Muon Channel}\\
$|\eta|$ bin & [0.0, & [0.4, & [0.8, & [1.2, & [1.6, & [2.0, &
[0.0, & [0.4, & [0.8, & [1.2, & [1.5, & [1.8, \\
& 0.4]& 0.8]& 1.2]& 1.4]& 2.0]& 2.4]&
0.4]& 0.8]& 1.2]& 1.5]& 1.8]& 2.1]\\\hline
Charge Misident. & 0.02& 0.03& 0.03& 0.08& 0.09& 0.10& 0& 0& 0& 0& 0& 0\\
Eff. Ratio & 0.70& 0.70& 0.70& 0.70& 0.70& 0.70& 0.59& 0.39& 0.92& 0.72& 0.81& 1.17\\
$ e/\mu$ Scale & 0.11& 0.09& 0.19& 0.47& 0.40& 0.45& 0.50& 0.48& 0.50& 0.48& 0.50& 0.42\\
Sig. \& Bkg. Estim.& 0.16& 0.19& 0.26& 0.33& 0.25& 0.25& 0.23& 0.29& 0.34& 0.40& 0.53& 0.58\\\hline
Total & 0.73& 0.73& 0.77& 0.90& 0.85 &0.87 & 0.80& 0.68& 1.10& 0.95& 1.08& 1.37\\\hline\hline
\multicolumn{13}{c}{ \rule[-1.8mm]{0mm}{5.5mm} $p_{\mathrm{T}}^{\ell} >$ 30\GeVc }\\\cline{1-13}
&\multicolumn{6}{c|}{Electron Channel}& \multicolumn{6}{c}{Muon Channel}\\
$|\eta|$ bin & [0.0, & [0.4, & [0.8, & [1.2, & [1.6, & [2.0, &
[0.0, & [0.4, & [0.8, & [1.2, & [1.5, & [1.8, \\
& 0.4]& 0.8]& 1.2]& 1.4]& 2.0]& 2.4]&
0.4]& 0.8]& 1.2]& 1.5]& 1.8]& 2.1]\\\hline
Charge Misident. & 0.02& 0.02& 0.03& 0.07& 0.08& 0.10& 0& 0& 0& 0& 0& 0\\
Eff. Ratio & 0.70& 0.70& 0.70& 0.70& 0.70& 0.70& 0.59& 0.39& 0.93& 0.72& 0.82& 1.18\\
$ e/ \mu$ Scale & 0.07& 0.17& 0.26& 0.46& 0.53& 0.55& 0.80& 0.78& 0.83& 0.81& 0.73& 0.77\\
Sig. \& Bkg. Estim. & 0.16& 0.19& 0.26& 0.33& 0.25& 0.25& 0.20& 0.20& 0.27& 0.35& 0.51& 0.56\\\hline
Total & 0.72& 0.75& 0.79& 0.91& 0.92& 0.93& 1.01& 0.90& 1.27& 1.14& 1.21& 1.52\\\hline\hline
\end{tabular*}
}
\end{center}
\end{table*}
\section{Results and Conclusions}
Table~\ref{tab1} summarizes systematic uncertainties for both the
electron
and muon channels.
The measured charge asymmetry results are summarized in
Table~\ref{tab2}
with both statistical
and systematic uncertainties shown.
The measurements have been repeated with a lepton $\pt^{\ell} >$ 30\GeVc. This
requirement
selects a subset of events with lepton pseudorapidity closer to the
$\PW$ boson rapidity and enables us to test PDF predictions in a more
constrained region of phase space.
The electron and muon measurements are in agreement with each other.
The experimental results are compared to theoretical predictions
obtained using {\sc RESBOS}~\cite{RES1,RES2,RES3} and {\sc MCFM}~\cite{MCFM} generators interfaced with
CT10W PDF model~\cite{CTEQ:1007}.
The {\sc RESBOS} generator performs a resummation at the next-to-next-to-leading
logarithmic order and gives a
more realistic
description of the $\PW$ boson $\pt$ spectrum
than a fixed-order calculation.
Figure~\ref{fig2} shows a comparison of these asymmetries to
predictions from the MSTW2008NLO PDF model~\cite{MSTW:0901} and the CT10W PDF model~\cite{CTEQ:1007}. The central values of both predictions are
obtained using the
{\sc MCFM} MC~\cite{MCFM} and the PDF error bands are estimated using
the PDF reweighting technique~\cite{PDFERROR}.
Our data suggest a flatter pseudorapidity
dependence of the
asymmetry than the PDF
models studied.
\begin{table*}[tb]
\caption{Summary of charge asymmetry~($\mathcal{A}$) results.
The first uncertainty is
statistical and the second is systematic. The theoretical predictions are
obtained using {\sc RESBOS}~($\mathcal{A}^{\mathrm{R}}$) and {\sc MCFM}~($\mathcal{A}^{\mathrm{M}}$) interfaced with CT10W PDF model.
The PDF uncertainties~($\Delta (+/-)$) are estimated using the PDF reweighting
technique.
The charge asymmetries and PDF errors are given in percent.
For each pseudorapidity bin the
theoretical prediction is calculated using the
averaged differential cross sections for positively
and negatively charged leptons respectively. The statistical
uncertainty on the theoretical prediction is about 0.1\%. }
\label{tab2}
\begin{center}
{\footnotesize
\begin{tabular*}{\textwidth}{@{\extracolsep{\fill}} l|cccc|cccc} \hline\hline
& \multicolumn{4}{c|}{\rule[-1.8mm]{0mm}{6.0mm} $p^{\ell}_{\mathrm{T}}>$ 25\GeVc} &
\multicolumn{4}{c}{\rule[-1.8mm]{0mm}{6.0mm} $p^{\ell}_{\mathrm{T}} >$ 30\GeVc}\\ \hline\hline
\rule[-1.8mm]{0mm}{5.5mm}$|\eta^\textrm{e}|$&\rule[-1.8mm]{0mm}{5.5mm} $\mathcal{A}(\textrm{e})$ ($\pm \mathrm{stat}\pm \mathrm{sys}$)&\rule[-1.8mm]{0mm}{5.5mm} $\mathcal{A}^{\mathrm{R}}$& \rule[-1.8mm]{0mm}{5.5mm} $\mathcal{A}^{\mathrm{M}}$& \rule[-1.8mm]{0mm}{5.5mm} $\Delta (+/-)$ &\rule[-1.8mm]{0mm}{5.5mm}
$\mathcal{A}(\textrm{e})$ ($\pm \mathrm{stat}\pm \mathrm{sys}$)&\rule[-1.8mm]{0mm}{5.5mm} $\mathcal{A}^{\mathrm{R}}$& \rule[-1.8mm]{0mm}{5.5mm} $\mathcal{A}^{\mathrm{M}}$& \rule[-1.8mm]{0mm}{5.5mm} $\Delta (+/-)$ \\\hline
\mbox{[0.0, 0.4]} & 15.5 $\pm$ 0.6 $\pm$ 0.7 & 15.7& 15.3& ${+0.8}/{- 1.0} $&
13.4 $\pm$ 0.7 $\pm$ 0.7& 13.4& 13.1&${+0.7}/{-0.9} $\\
\mbox{[0.4, 0.8]} & 16.7 $\pm$ 0.6 $\pm$ 0.7 & 16.9& 16.7&${+0.9}/{-1.0} $&
15.1 $\pm$ 0.7 $\pm$ 0.8 & 14.6& 14.5&${+0.8}/{-0.8} $\\
\mbox{[0.8, 1.2]} & 17.5 $ \pm$ 0.7 $\pm$ 0.8 & 19.3& 19.2&${+0.8}/{- 1.1} $&
15.2 $\pm$ 0.7 $\pm$ 0.8 & 16.9& 16.8&${+0.8}/{-1.0} $\\
\mbox{[1.2, 1.4]} & 19.4 $\pm$ 1.0 $\pm$ 0.9 & 21.6& 21.7&${+0.8}/{- 1.1} $&
16.9 $\pm$ 1.1 $\pm$ 0.9 & 19.1& 18.9&${+0.8}/{-1.0} $\\
\mbox{[1.6, 2.0]} & 23.6 $\pm$ 0.8 $\pm$ 0.9 & 25.6& 25.4&${+0.8}/{- 1.1} $&
21.3 $\pm$ 0.9 $\pm$ 0.9 & 23.4& 23.7&${+0.8}/{-1.1} $\\
\mbox{[2.0, 2.4]} & 27.1 $\pm$ 0.8 $\pm$ 0.9 & 27.1& 26.9&${+0.8}/{- 1.1} $&
25.0 $\pm$ 0.9 $\pm$ 0.9 & 25.7& 25.4&${+0.8}/{-1.1} $\\\hline\hline
\rule[-1.8mm]{0mm}{5.5mm} $|\eta^{\mu}|$& \rule[-1.8mm]{0mm}{5.5mm} $\mathcal{A}(\mu)$($\pm \mathrm{stat}\pm \mathrm{sys}$) &\rule[-1.8mm]{0mm}{5.5mm} $\mathcal{A}^{\mathrm{R}}$& \rule[-1.8mm]{0mm}{5.5mm} $\mathcal{A}^{\mathrm{M}}$&\rule[-1.8mm]{0mm}{5.5mm} $\Delta (+/-)$ & \rule[-1.8mm]{0mm}{5.5mm}
$\mathcal{A}(\mu)$($\pm \mathrm{stat}\pm \mathrm{sys}$) & \rule[-1.8mm]{0mm}{5.5mm} $\mathcal{A}^{\mathrm{R}}$& \rule[-1.8mm]{0mm}{5.5mm} $\mathcal{A}^{\mathrm{M}}$& \rule[-1.8mm]{0mm}{5.5mm} $\Delta (+/-)$ \\\hline
\mbox{[0.0, 0.4]} & 14.7 $\pm$ 0.6 $\pm$ 0.8 & 15.7& 15.3&${+0.8}/{-1.0} $&
13.1 $\pm$ 0.7 $\pm$ 1.0& 13.4& 13.1&${+0.7}/{-0.9} $\\
\mbox{[0.4, 0.8]} & 15.9 $\pm$ 0.6 $\pm$ 0.7 & 16.9& 16.7&${+0.9}/{-1.0} $&
13.9 $\pm$ 0.7 $\pm$ 0.9 & 14.6& 14.5&${+0.8}/{-0.8} $ \\
\mbox{[0.8, 1.2]} & 18.4 $ \pm$ 0.6 $\pm$ 1.1 & 19.3& 19.2&${+0.8}/{-1.1} $&
15.8 $\pm$ 0.7 $\pm$ 1.3 & 16.9& 16.8&${+0.8}/{-1.0} $\\
\mbox{[1.2, 1.5]} & 20.7 $\pm$ 0.7 $\pm$ 1.0 & 22.0& 22.0&${+0.8}/{-1.1} $&
18.5 $\pm$ 0.8 $\pm$ 1.1 & 19.6& 19.4&${+0.8}/{-1.0} $\\
\mbox{[1.5, 1.8]} & 23.1 $\pm$ 0.8 $\pm$ 1.1 & 24.6& 24.5&${+0.8}/{-1.1} $&
20.2 $\pm$ 0.8 $\pm$ 1.2 & 22.2& 21.9&${+0.8}/{-1.1} $\\
\mbox{[1.8, 2.1]} & 25.3 $\pm$ 0.8 $\pm$ 1.4 & 26.5& 26.3&${+0.8}/{-1.0} $&
23.1 $\pm$ 0.9 $\pm$ 1.5 & 24.5& 24.1&${+0.8}/{-1.1} $\\
\hline\hline
\end{tabular*}
}
\end{center}
\end{table*}
\begin{figure}[tb]
\begin{center}
\includegraphics[width=0.75\columnwidth]{fig2.pdf}
\hspace{1cm}
\caption{Comparison of the measured lepton charge
asymmetry to
different PDF models for a) lepton $\pt^{\ell} >$ 25\GeVc and b) lepton $p^{\ell}_T >$ 30\GeVc.
The error bars include both statistical and systematic uncertainties.
The PDF uncertainty band is corresponding to the 90\% confidence interval~(C.I.).
The bin width for each data point is shown by the filled bars in fig. b). The data points are placed at the centers of pseudorapidity bins,
except that for display purposes the first three data points are
shifted $+0.025$~($-0.025$) for
electron (muon). }
\label{fig2}
\end{center}
\end{figure}
In summary, we have measured the lepton charge asymmetry in both the $\mathrm{W} \rightarrow \mathrm{e}\nu$ and
$\mathrm{W} \rightarrow \mu\nu$ channels using a data sample corresponding to
an integrated luminosity
of $36\ \mathrm{pb}^{-1}$ collected with the CMS detector at the LHC.
In each pseudorapidity bin the precision of the most inclusive measurements
is less than 1.6\% for both channels.
This
high precision measurement of the $\PW$ lepton charge asymmetry at the LHC
provides new inputs to
the PDF global fits.
\section{Acknowledgements}
We wish to thank James~W. Stirling for very useful discussions on PDF models and John Campbell for providing theoretical
predictions based on MCFM.
We wish to congratulate our colleagues in the CERN accelerator departments for the excellent performance of the LHC machine. We thank the technical and administrative staff at CERN and other CMS institutes, and acknowledge support from: FMSR (Austria); FNRS and FWO (Belgium); CNPq, CAPES, FAPERJ, and FAPESP (Brazil); MES (Bulgaria); CERN; CAS, MoST, and NSFC (China); COLCIENCIAS (Colombia); MSES (Croatia); RPF (Cyprus); Academy of Sciences and NICPB (Estonia); Academy of Finland, ME, and HIP (Finland); CEA and CNRS/IN2P3 (France); BMBF, DFG, and HGF (Germany); GSRT (Greece); OTKA and NKTH (Hungary); DAE and DST (India); IPM (Iran); SFI (Ireland); INFN (Italy); NRF and WCU (Korea); LAS (Lithuania); CINVESTAV, CONACYT, SEP, and UASLP-FAI (Mexico); PAEC (Pakistan); SCSR (Poland); FCT (Portugal); JINR (Armenia, Belarus, Georgia, Ukraine, Uzbekistan); MST and MAE (Russia); MSTD (Serbia); MICINN and CPAN (Spain); Swiss Funding Agencies (Switzerland); NSC (Taipei); TUBITAK and TAEK (Turkey); STFC (United Kingdom); DOE and NSF (USA).
|
\section*{References}
|
\section{Introduction}
\label{intro}
During the past 20 years vast progress has been made in unveiling the
properties of neutrinos. For decades neutrinos were thought to be massless,
which no longer holds true: flavour oscillations found in the leptonic sector, studying neutrinos
coming from the sun \cite{sno,superKsolar}, the atmosphere \cite{superKatmos}, high energy
accelerators beams \cite{minos10,opera10} and nuclear power plants \cite{kamland03}, are explained by neutrino oscillations requiring a non zero neutrino mass.
However, no absolute mass scale can be fixed with experiments studying the oscillatory behaviour.
To achieve this, one has to investigate weak decays, such as beta decays or neutrinoless double beta decays
\begin{equation}
(Z,A) \rightarrow (Z+2,A) + 2 e^- \quad (0\mbox{$\nu\beta\beta$-decay}) \, .
\end{equation}
The latter violates total lepton number by two units and thus is not allowed in the Standard Model. The 0\mbox{$\nu\beta\beta$-decay}\ is the gold plated process to distinguish whether neutrinos are Majorana or Dirac particles. Furthermore, a match of helicities of the intermediate neutrino states is necessary which is done in the easiest way by introducing a neutrino mass. This mass is linked with the experimentally observable half-life via
\begin{equation}
\label{eq:1}
\left(T_{1/2}^{0 \nu}\right)^{-1} = G^{0 \nu}(Q, Z) \left| M_{GT}^{0\nu} - M_{F}^{0\nu} \right|^2 \left(\frac{\ensuremath{\langle m_{\nu_e} \rangle}~}{m_e}\right)^2 \, ,
\end{equation}
where \ensuremath{\langle m_{\nu_e} \rangle}~ is the effective Majorana neutrino mass, given by $\ensuremath{\langle m_{\nu_e} \rangle}~ = \left| \sum_{i}U_{ei}^2m_i\right|$ and $U_{ei}$ is the corresponding element in the
leptonic PMNS mixing matrix, $G^{0 \nu}(Q, Z) $ is a phase space factor and $M_{GT}^{0\nu} - M_{F}^{0\nu}$ describes the nuclear transition matrix element.
The experimental signature is the emission of two electrons with a sum energy
corresponding to the Q-value of the nuclear transition.
A potential evidence has been claimed in the 0\mbox{$\nu\beta\beta$-decay}\ of $^{76}$Ge with $T_{1/2}^{0 \nu} = \unit[2.23^{+0.44}_{-0.31} \times 10^{25}]{yr}$ at \unit[90]{\%} CL \cite{kla06}. In addition, the SM process of neutrino accompanied double beta decay,
\begin{equation}
(Z,A) \rightarrow (Z+2,A) + 2 e^- + 2 \mbox{$\nu_e$} \quad (2\mbox{$\nu\beta\beta$-decay}) \,
\end{equation}
can be investigated, which is expected with half-lives arround \unit[$10^{20}$]{yr}. For recent reviews see \cite{avi08}.
Additional information is provided by the alternative process of positron decay in combination with electron capture (EC).
Three different decay modes can be considered:
\begin{align}
(Z,A) \rightarrow& (Z-2,A) + 2 e^+ + (2 \mbox{$\nu_e$}) &\mbox{(\ensuremath{\beta^+\beta^+}~{})}\\
e^- + (Z,A) \rightarrow& (Z-2,A) + e^+ + (2 \mbox{$\nu_e$}) & \mbox{(\ensuremath{\beta^+/{\rm EC}}{})}\\
2 e^- + (Z,A) \rightarrow& (Z-2,A) + (2 \mbox{$\nu_e$}) &\mbox{(\ensuremath{{\rm EC/EC}}~{})}
\end{align}
Decay modes containing a positron have a reduced Q-value as each generated positron accounts for a reduction
of 2 $m_ec^2$. Thus, the full energy is only available in the \ensuremath{{\rm EC/EC}}~ mode and makes it the most probable one.
However, it is also the most difficult to detect, only producing X-rays instead of \unit[511]{keV} gammas.
Furthermore, it has been shown that \ensuremath{\beta^+/{\rm EC}}\ transitions have an enhanced sensitivity to right-handed weak
currents (V+A interactions) \cite{hir94} and thus would help to disentangle the physics mechanism of 0\mbox{$\nu\beta\beta$-decay}.
In the last years, also neutrinoless \ensuremath{{\rm EC/EC}}~ modes have been discussed with renewed interest, because of a potential
resonance enhancement up to a factor of $10^6$ in the decay if the initial and final excited state are degenerate
\cite{suj04}. Recently, a series of isotopes with extreme low Q-value was explored for enhancement in the ECEC mode
and with $^{152}$Gd a very promising candidate was found \cite{eli11}.
Another branch of search is linked to excited state transitions. The signal in this case is extended by looking at
the corresponding de-excitation gammas. However, in the approach of a passive sample on a Ge-detector
it will not allow to distinguish between the 2\mbox{$\nu\beta\beta$-decay}\ and 0\mbox{$\nu\beta\beta$-decay}\ mode. Thus, the deduced half-live limits are valid for both.
The investigation of 2\mbox{$\nu\beta\beta$-decay}\ modes into excited states will add information on nuclear structure, valuable for matrix
element calculations. Furthermore, a potential observation of 0\mbox{$\nu\beta\beta$-decay}\ into an excited $2^+$-state would likely point to other
contributions besides neutrino masses. The searches described in this paper are based on the search for excited
state transitions.
An element getting little attention in the past is palladium with the isotopes of interest \mbox{$^{110}$Pd}~ and \mbox{$^{102}$Pd}.
Among the eleven 0\mbox{$\nu\beta\beta$-decay}\ candidates with a Q-value larger than \unit[2]{MeV}, \mbox{$^{110}$Pd}~ has several advantages: it has the second-highest natural abundance (\unit[11.72]{\%}) and, in addition, it is an excellent candidate to probe the single-state dominance hypothesis for 2\mbox{$\nu\beta\beta$-decay}, i.e. that only the lowest lying intermediate 1$^+$-state will contribute to the nuclear transition matrix element describing its 2\mbox{$\nu\beta\beta$-decay}. Only one rather weak experimental limit in the order of \unit[$10^{17}$]{yr} exists for 0\mbox{$\nu\beta\beta$-decay}\ ground state transitions in \mbox{$^{110}$Pd}~ \cite{win52}. Theoretical predictions for the 2\mbox{$\nu\beta\beta$-decay}\ ground state mode are in the range of \unit[0.12 - 29.96 $\times$ 10$^{20}$]{yr} \cite{chandra05,semenov2000,civitarese1998,stoica1994,hirsch1994,staudt1990,domin2005,suhonen2011}. Theoretical predictions for the excited state transitions are \mbox{\unit[4.4 $\times$ 10$^{25}$]{yr}} \cite{domin2005}, \mbox{\unit[8.37 $\times$ 10$^{25}$]{yr}} \cite{stoica1994}, \mbox{\unit[1.5 $\times$ 10$^{25}$]{yr}} \cite{raduta2007} and \mbox{\unit[0.62 - 1.3 $\times$ 10$^{25}$]{yr}} \cite{suhonen2011} for the $2^+_1$ state and \mbox{\unit[2.4 $\times$ 10$^{26}$]{yr}} \cite{domin2005} and \mbox{\unit[4.2 - 9.1 $\times$ 10$^{23}$]{yr}} \cite{suhonen2011} for the $0^+_1$ state.
The second isotope \mbox{$^{102}$Pd}\ has a Q-value of \unit[1172]{keV}, a natural abundance of \unit[1.02]{\%} and is able to decay via \ensuremath{{\rm EC/EC}}~ and
\ensuremath{\beta^+/{\rm EC}}. These decay modes have never been studied for \mbox{$^{102}$Pd}\ experimentally and no theoretical predictions exist. The level schemes
of both isotopes are shown in Fig.~\ref{fig:levels}.
\begin{figure}[h]
\centering
\includegraphics[width=16cm]{Pd_LevelScheme.eps}
\caption{Level schemes of \mbox{$^{110}$Pd}~ (left) and \mbox{$^{102}$Pd}\ (right) decays.}
\label{fig:levels}
\end{figure}
\section{Experimental Setup}
\label{setup}
The measurement was performed in the Felsenkeller Underground Laboratory in Dresden with a shielding depth of 120 mwe.
A sample of \unit[802.35]{g} of Pd was used, which was purified before by C. HAFNER GmbH + Co. KG.
It was placed in a standard Marinelli baker (D6) of \unit[70]{cm}
diameter and \unit[21]{cm} height. The sample was positioned on a HPGe detector with an efficiency of \unit[90]{\%} routinely used for $\gamma$-spectroscopic measurements. The detector has an \unit[1]{mm} thick Al-window towards the sample.
It is surrounded by a \unit[5]{cm} copper shielding embedded in another shielding of \unit[15]{cm} of clean
lead. The inner \unit[5]{cm} of the lead shielding have a low contamination of $^{210}$Pb of only \unit[$2.7 \pm 0.6$]{Bq/kg} while
the outer \unit[10]{cm} have an activity of \unit[$33 \pm 0.4$]{Bq/kg}.
The detector is placed inside a special building which acts as a Faraday cage and a Rn shield; additionally, the setup is flushed with nitrogen in order to reduce Rn contamination. More details can be found in \cite{deg09,degering08} and a schematic drawing is shown in Fig.~\ref{fig:schema}. The Pd was stored underground for more than one year prior to the measurement except for \unit[18]{d} of purification. The data were collected with a 8192 channel MCA from Ortec and were converted into the ROOT format for analysis.
\begin{figure}[h]
\begin{center}
\includegraphics[width=10cm]{PdDetector.eps}
\caption{\small Schematic drawing of the used setup. Composed from pictures in \cite{degering08}}
\label{fig:schema}
\end{center}
\end{figure}
An extensive calibration was performed using 17 $\gamma$-lines from 8 different nuclides resulting in a linear energy calibration curve of
\begin{equation}
E [\unit{keV}] = \unit[0.342746]{\frac{\unit{keV}}{channel}\times channel} - \unit[4.337734]{keV}\, .
\end{equation}
The measurement range of the spectrum goes up to \unit[2.8]{MeV}.
The resolution was calibrated using the calibration lines and fitted with a second order polynomial.
The actual values at the energies of the lines of interest will be discussed in the corresponding analysis section.
Despite purification, the measured spectrum is dominated by intrinsic contaminations of the Pd. Clear $\gamma$-lines from the $^{238}$U and $^{232}$Th
decay chains as well as $^{40}$K are visible. However, former Americium contributions have been removed completely by
the purification.
The actual background spectrum of the detector system itself without any sample is at least an order of magnitude smaller
in the regions of interest and can be neglected, hence the spectrum is completely dominated by the Pd sample
contaminations.
\section{Analysis and Results}
A total of \unit[16.2]{d} of data were accumulated resulting in \unit[13.00]{$\rm kg \cdot d$}
of exposure.
In the following the two isotopes of interest are discussed separately.
For the analysis, the natural abundances of the latest IUPAC evaluation
have been used which are \unit[11.27]{\%} (\mbox{$^{110}$Pd}~) and \unit[1.02]{\%} (\mbox{$^{102}$Pd}) respectively \cite{bie93}.
As the searches are purely based on gamma detection, the obtained results
apply for both, 0\mbox{$\nu\beta\beta$-decay}\ and 2\mbox{$\nu\beta\beta$-decay}\ modes. The decays into the first excited $0_1^+$-state
de-excite via an intermediate $2^+_1$- state. Thus, there will be an angular correlation
among the gammas, with an observational probability W($\theta$) that the second
gamma is emitted with an angle $\theta$ with respect to the first one given by
\begin{equation}
W(\theta)= \frac{5}{4} \left(1 - 3 \cos^2 \theta +4 \cos^4 \theta\right)\, .
\end{equation}
It can be seen that the probability of both gammas beeing emitted in the same direction is larger than an uncorrelated emission. However, the detection efficiency for a single gamma is small which results in a low probability to observe a summation peak and thus the searches are based on the individual gamma energies only.
The efficiency for full energy detection was determined and cross checked in several ways.
The most important one was replacing the actual used volume by a SiO$_2$ sample
of exactly the same geometry. The intrinsic contaminations of the natural decay chains of $^{238}$U
and $^{232}$Th as well as $^{40}$K produced various $\gamma$-lines and acted as an extended calibration source. The 17 $\gamma$-lines used for the energy calibration were also used for the efficiency determination
in the region from \unit[238.6]{keV} (from $^{212}$Pb decay) up to \unit[2614.3]{keV} (from $^{208}$Tl decay).
The efficiency in the region above \unit[200]{keV} can be well fitted by two exponential functions.
It varies between \unit[5.77]{\%} at \unit[468]{keV} and \unit[3.89]{\%} at \unit[815]{keV}, being the lowest and highest
energy lines of interest for the search described in this paper.
To account for the difference in self-absorption of Pd
and SiO$_2$, extensive Monte Carlo simulations were performed using the AMOS code
\cite{amos}. The amount of Monte Carlo was chosen in a way that the statistical error for the
full energy peak for both, Pd and SiO$_2$, was less than \unit[0.1]{\%}.
The simulations agree within an error of less than \unit[15]{\%} with the measurements and tend to
be slightly higher. This can easily be explained by small geometric differences in the simulation
and the experiment. However, independent from that is the ratio of both self-absorption simulations for Pd and SiO$_2$.
Hence, the ratio was used to scale the experimentally well determined efficiency curve of SiO$_2$ to the one of Pd (Fig.~\ref{fig:efficency}).
A validation of the procedure was performed within the vicinity of the lines of interest
by using prominent background lines apparent in the spectrum,
namely \unit[295.21]{keV} and \unit[351.92]{keV} (from $^{214}$Pb),
\unit[233.63]{keV} (from $^{212}$Pb), \unit[583.19]{keV} (from $^{208}$Tl) and \unit[609.32]{keV} (from $^{214}$Bi),
which results in good agreement.
\begin{figure}[h]
\begin{center}
\includegraphics[width=10cm]{EffCal_fit.eps}
\caption{\small Measured efficiencies and the fitting function.}
\label{fig:efficency}
\end{center}
\end{figure}
The analysis is based on the extraction of upper count limits with the Feldman Cousins method described in \cite{fc98}. A simple constant was chosen as a background model after concluding that the area around the ROI is sufficiently flat. The background was determined with a likelihood fit in a selected region around the peak position taken to be \unit[$\pm 30$]{keV} excluding the peak range which is considered to be \unit[$\pm 5$]{keV}.
Likelihood fits with the Gaussian peak shapes and the background constants resulted in statistical downward fluctuations for all four peaks. This is a clear sign that no significant signal is observed (see Fig.~\ref{fig:pd102} and \ref{fig:pd110}). In this case, an upper limit for the count rate can be calculated using the background only hypothesis and the maximal statistical fluctuation of a Gaussian distributed background for a certain confidence level. This is commonly referred to as sensitivity. However, the Feldman Cousins approach also considers the observed downward fluctuations and results in more appropriate results for low count rates.
\begin{figure}[H]
\begin{center}
\includegraphics[width=10cm]{ROI_469.eps}
\caption{\small Peak region of \mbox{$^{102}$Pd}\ for the $0^+ \rightarrow 2^+$ transition with the illustration of the background model, the best fit and the peak with the upper count limit at \unit[95]{\%} CL.}
\label{fig:pd102}
\end{center}
\end{figure}
\begin{figure}[H]
\begin{center}
\includegraphics[width=10cm]{ROI_815.eps}
\caption{\small Peak region of \mbox{$^{110}$Pd}~ for the $0^+ \rightarrow 2^+$ transition with the illustration of the background model, the best fit and the peak with the upper count limit at \unit[95]{\%} CL.}
\label{fig:pd110}
\end{center}
\end{figure}
In order to obtain a numerical value for the upper limit, all bins within the FWHM of an expected peak are combined into a single analysis bin. The resulting fraction of the peak that is covered by the analysis bin is dependent on the binning of the data but always larger than \unit[76]{\%}. The background expectation and the measured count rate are used to evaluate an upper count limit for this bin with the ROOT implementation of the Feldman and Cousins method. The result is then scaled to the full peak area.
From the background point of view, exactly the same $\gamma$-lines in the performed search could be produced from the beta decays of the
intermediate nuclide of the investigated double beta system, which will be discussed in more detail in the next
section. However, their contribution can be rejected by the non-observation of other,
more prominent $\gamma$-lines at different locations in the spectrum.
The only prominent background line to be expected within \unit[$\pm 5$]{keV} of any of the four
lines under investigation is from $^{137}$Cs at \unit[661.66]{keV} potentially influencing the $2^+_1$-limit in the \mbox{$^{110}$Pd}~ system.
However, no indication of this line is observed.
\subsection{The \mbox{$^{110}$Pd}~ system}
\label{data}
Two lines were investigated for \mbox{$^{110}$Pd}~ at energies \unit[657.76]{keV} (corresponding to the 2$^+_1$ transition)
and \unit[815.35]{keV} (additionally emitted in the $0^+_1$ decay) respectively.
The corresponding energy resolutions at these energies are 1.51 and \unit[1.61]{keV} (FWHM).
Potential $\gamma$-lines mimicking the signal would come from $^{110}$Ag and $^{110m}$Ag
decays. $^{110}$Ag has a half-life of 24.6 s only and thus has to be produced in-situ. With the
given shielding this can be excluded. Potentially more dangerous is the long-living $^{110m}$Ag
(half-live of 249.79 d). This isotope has two prominent lines at \unit[1384.3]{keV} and \unit[1505.04]{keV}. They are not visible in the spectrum and thus this contribution can be excluded for this search.
No lines were visible at both peak positions of interest and thus an upper limit (\unit[95]{\%} CL)
of signal events of 10.53 and 7.34 could be extracted for \unit[657.76]{keV} and \unit[815.35]{keV} respectively.
Using the known Pd mass and efficiencies, this can be converted into lower half-live limits of
\begin{align}
T_{1/2}^{(0\nu + 2\nu)}\, \mbox{$^{110}$Pd}~ \rightarrow \rm {^{110}Cd} (0^+_1,\unit[815.3]{keV}) > \unit[5.89 \times 10^{19}]{yr}\ (\unit[95]{\%} CL)& \\
T_{1/2}^{(0\nu + 2\nu)}\, \mbox{$^{110}$Pd}~ \rightarrow \rm {^{110}Cd} (2^+_1,\unit[657.8]{keV}) > \unit[4.40 \times 10^{19}]{yr}\ (\unit[95]{\%} CL)& \, .
\end{align}
These are the first experimental limits for excited state transitions in the \mbox{$^{110}$Pd}~ system.
\subsection{The \mbox{$^{102}$Pd}\ system}
Two lines were investigated for \mbox{$^{102}$Pd}\ at energies \unit[468.59]{keV} (only emitted in the $0^+_1$ decay) and
\unit[475.05]{keV} (corresponding to the 2$^+_1$ transition) respectively.
The energy resolution at these energies is \unit[1.39]{keV} (FWHM) for both lines.
Potential $\gamma$-lines mimicking the signal would come from $^{102}$Rh and $^{102m}$Rh
decays. $^{102}$Rh with a half-life of 207 d has no reasonable line to check. The strongest one is the \unit[475.05]{keV} line.
As there is no signal in this region it can be concluded that it does not contribute to the \unit[468.59]{keV}
region. On the other hand $^{102m}$Rh (half-live of \unit[2.9]{yr}) has multiple lines to explore, the most restricting ones
are a line at \unit[631.28]{keV} with \unit[56]{\%} emission probability and at \unit[697.49]{keV} with \unit[44]{\%}. Both of them
are not observed in the spectrum and thus can exclude such a contribution.
No lines were visible at both peak positions of interest and thus an upper limit (\unit[95]{\%} CL)
of signal events of 17.64 and 6.24 could be extracted for \unit[475.05]{keV} and \unit[468.59]{keV} respectively.
The obtained half-live limits are
\begin{align}
T_{1/2}^{(0\nu + 2\nu)}\, \mbox{$^{102}$Pd}\ \rightarrow \rm {^{102}Ru} (0^+_1,\unit[468.6]{keV}) > \unit[7.64 \times 10^{18}]{yr}\ (\unit[95]{\%} CL)& \\
T_{1/2}^{(0\nu + 2\nu)}\, \mbox{$^{102}$Pd}\ \rightarrow \rm {^{102}Ru} (2^+_1,\unit[475.1]{keV}) > \unit[2.68 \times 10^{18}]{yr}\ (\unit[95]{\%} CL)& \, .
\end{align}
These are the first experimental limits on \mbox{$^{102}$Pd}\ double beta decays.
\section{Summary}
\label{conclusion}
Double beta decay transitions into excited states for the two Pd-isotopes \mbox{$^{102}$Pd}\ and \mbox{$^{110}$Pd}~ have been
investigated for the first time. These transitions contain valuable informations about the physics mechanism of double beta
decay and the involved nuclear physics. However, no signal into the first excited 0$^+$ and 2$^+$ states have
been observed.
\section*{Acknowledgement}
The authors would like to thank D. Degering (VKTA Dresden) for his help with the
underground measurements and D. Sommer for her help with the Monte Carlo
simulations for efficiency determinations.
|
\section{Introduction}
\label{intro}
The current development of nanomaterials for molecular electronics
has triggerd the study of heat conduction in low-dimensional systems from new perspectives.
It is well known that the derivation of the thermal properties in one and two dimensional systems
is a controversial topic which has been deeply studied in the last two decades.~\cite{Lepri03,Dhar08}
In the case of perfect harmonic lattices, the lack of phonon-phonon interactions
leads to no thermal resistivity, which gives rise to an infinite thermal conductivity in the thermodynamic limit.
This means, Fourier's law is not valid in these systems and heat conduction exhibits anomalous behavior.
Nonintegrability and an external substrate potential constitute the sufficient conditions to completely change
this scenario.~\cite{Hu98}
In such a case, the conservation law of total momentum is violated and
Fourier's law reveals as valid. Recently, the most studied proposals are
the Frenkel-Kontorova (FK),~\cite{Gillan85,Tsironis99,Li06_fi4,Zhong09} and the $\phi^4$ potentials.~\cite{Aoki00,Hu99,Li07,Piazza09}
Several authors have
revisited the unusual thermal properties of one dimensional systems
in order to propose different applications such as thermal rectifiers or thermal transistors.~\cite{Terraneo02,Hu06,Li06,Zhang10}
Some of these studies have been extended to two and three dimensional structures~\cite{Lee05, Saito10} and complex networks~\cite{Liu07,Liu10}, where it has been
addressed the relevant influence of the coupling among several one dimensional lattices.
More interestingly, the analysis of the mechanisms mediating energy flow in low dimensional biomolecules
is a fundamental issue for the understanding of many biologically relevant
functions. The electronic and vibrational degrees of freedom of biomolecular systems, specially those containing helix structures
, i.e. double-helix DNA or $\alpha$-helices in proteins, can be modeled as ladder structures
of coupled one-dimensional lattices.~\cite{Iguchi97,Diaz07,Scott92, Henning02}
Recently, the thermal conductivity of double-stranded molecules has been studied in Ref.~\onlinecite{Zhong10} where
the interchain interaction between two identical chains has revealed as a positive or negative effect on heat conduction,
depending on the strength of the nonlinearity present in the system.
In this paper we address some interesting issues about heat transport along double-stranded molecular systems
which remain open. In particular, we consider a ladder-model where two different lattices, a harmonic and an anharmonic one, are
coupled by harmonic forces. In view of the clearly different thermal properties of these two subsystems
when they are isolated, our work analyzes how the coupled system behavior is affected depending on the coupling strength.
In addition, due to the asymmetry present in the system, we address the possibility of heat rectification.
\section{Model}
\label{model}
In this section we will present the theoretical formalism we will be dealing with
in our study on heat transport along double-stranded molecules.
In particular, we consider a system conformed by a harmonic one dimensional lattice
which is coupled by harmonic forces to an anharmonic one.
The dimensionless Hamiltonian of such a system is written as follows:
\begin{eqnarray}
{\cal H}&=& {\cal H}_{H}+{\cal H}_{A}+{\cal H}_{int} \ ,\nonumber \\
{\cal H}_{A}&=& \sum_{n=1}^N \frac{1}{2} \dot{x}^2_n + W(x_{n},x_{n-1})+V(x_n)\ ,\nonumber \\
{\cal H}_{H}&=& \sum_{n=1}^N \frac{1}{2} \dot{y}^2_n + W(y_{n},y_{n-1})\ ,\nonumber \\
{\cal H}_{int}&=& \sum_{n=1}^N k_{int} W(x_{n},y_{n})\ .
\label{Hamiltonian}
\end{eqnarray}
${\cal H}_{A}$ and ${\cal H}_{H}$ describe the dynamics of the subchains, the harmonic (H) and the anharmonic (A) one respectively,
each one consisting of $n=1\ldots N$ sites.
Both are affected by a harmonic potential $W(x_n,x_{n-1})=\frac{1}{2}(x_n-x_{n-1})^2$ and chain (A) is considered within the Frenkel-Kontorova model.
Thus, chain (A) is affected by an on-site potential $V(x_n)=\frac{-V_0}{4\pi^2}\cos 2\pi x_n$, whose strength is defined by $V_0$.
For the sake of simplicity, we will consider the same elastic constant for both chains, $k$, in terms of which the harmonic
interchain coupling $k_{int}$ is expressed. The latter interaction is considered within the Hamiltonian ${\cal H}_{int}$,
whose characteristic oscillator frequency is $\omega_0=k/m$ .
Notice that hereafter the period of the external potential, $b$, is taken as the length scale of the system and the magnitude
$kb^2$ will be the energy unit.~\cite{Hu98}
Fig.~\ref{sketch} presents a schematic view of the system under study.
\begin{figure}[ht]
\centerline{\includegraphics[width=70mm,clip]{contactos2.eps}}
\caption{Schematic view of a fragment of the ladder-model under consideration.
The harmonic lattice is connected at left and right edges to two independent heat reservoirs at temperatures $T_C$ and $T_H$ respectively.}
\label{sketch}
\end{figure}
Our aim is to study the formation of a thermal gradient along this double-stranded structure
by connecting two independent heat reservoirs at both ends of the harmonic lattice.
Heat baths effects will be included in the dynamics of the edge sites of the system, $y_1$ and $y_N$,
by using stochastic Langevin equations.~\cite{Lepri03}
This will modify the Newton's equations of motion of the affected sites, $n=1$ and $n=N$, as follows:
\begin{equation}
\frac{d^2y_{n}}{dt^2}=-W^{\prime}(y_n,y_{n-1})-W^{\prime}(y_{n},y_{n+1})-\gamma\frac{dy_n}{dt} + f_n(t) \ ,
\label{NewtonLang}
\end{equation}
where $\gamma$ is the coupling between the system and the stochastic bath and the prime indicates derivative
with respect to $y_n$. Random forces, which include temperature effects, are defined as:
\begin{equation}
\langle f_{n}(t) f_{n}(t')\rangle= 2 T \gamma \delta(t-t')\ .
\label{StocForces}
\end{equation}
The relationship between this dimensionless temperature $T$ and the real one $T_r$
is given by $T=K_B T^r/k b^2$, where $K_B$ refers to the Boltzmann constant.~\cite{Hu98}
The temperature of the baths $T$ will be different for the left cold ($T_C$) and right hot ($T_H$) sides of the chain.
Exploiting the equations of motion, the local heat flux is defined by the continuity equation as:
\begin{eqnarray}
J_n&=&\dot{x}_n\Big(\frac{\partial W^{\prime}(x_n,x_{n-1})}{\partial x_n}+k_{int}\frac{\partial W^{\prime}(x_n,y_n)}{\partial x_n}\Big)\nonumber \\
&+&\dot{y}_n\Big(\frac{\partial W^{\prime}(y_n,y_{n-1})}{\partial y_n}+k_{int}\frac{\partial W^{\prime}(x_n,y_n)}{\partial y_n}\Big)\ .
\label{localflux}
\end{eqnarray}
Molecular dynamics simulations are performed by the numerical method proposed by Greenside and Helfand as a correction
to the Runge-Kutta method for stochastic equations (3o4s2g),~\cite{Helfand79,Greenside81,Casado03} by considering
a long enough integration time such that the stationary state is established.
The time step is $\delta t=10^{-5}$ and the friction constant of the baths is set to $\gamma=0.5$ in all simulations.
In the final state the time averaged heat flux reaches a constant value along the system such that $J =\langle J_1\rangle=...=\langle J_N\rangle$.
Thus, the thermal conductivity for a finite system can be calculated as $\kappa =J N/(T_C-T_H)$.
Notice that $\kappa$ will be size-independent in the case of normal heat transport.
Similarly, in the steady state the time averaged temperature calculated as
$T_n=\langle \dot{x}^2_n + \dot{y}^2_n \rangle$ will reach the stationary thermal profile.
\section{Single anharmonic chain}
\label{single}
To better understand the influence of the coupling to nonlinear degrees of freedom, first
we would like to perform a preliminary study of a single Frenkel-Kontorova chain connected to
two independent heat reservoirs at temperatures $T_C=T_{M}-\delta T$ and right $T_H=T_{M}+\delta T$ at the left and right
edges of the chain. Thus, we integrate the Newton's equations derived from ${\cal H}_{A}$ and heat baths effects are included
in the dynamics of sites $x_1$ and $x_N$ according to Eq.~\ref{NewtonLang}.
We have performed molecular dynamics simulations for different values of $T_{M}$, $\delta T=0.05$ and $V_0=8$.
Once the steady state is reached, the stationary local temperature profile is calculated as $T_n=\langle \dot{x}^2_n\rangle$
which is shown in left panels of Fig.~\ref{singleU8}.
\begin{figure}[ht]
\centerline{\includegraphics[width=80mm,clip]{SingleU8new3.eps}}
\caption{Left panels show the stationary temperature profile calculated by numerical simulations in a FK lattice of $N=50$ sites connected
to two independent heat reservoirs at different temperatures, $T_{H,C}=T_{M}\pm 0.05$.
Right panels present the comparison between the numerical DOS calculated in equilibrium with the theoretical
frequency band obtained according to the SCPT (shadowed region). The solid line accounts for the weight function $F(T_{M})$.
The temperatures considered in the simulations are a) $T_{M}=0.10$,
b) $T_{M}=0.15$ and c) $T_{M}=0.20$. The inset shows the frequency band of states calculated by SCPT as a function
of temperature and $V_0=8$ in the shadowed region.}
\label{singleU8}
\end{figure}
As it is already known, our results in Fig.~\ref{singleU8} (b) and (c)
show that if the FK-potential strength is large enough respect to the temperature $T_{M}$,
a well-defined thermal gradient is formed. Furthermore, in such a case, the Fourier heat law is valid and the total heat flux along the system
does not diverge in the thermodynamic limit contrary to the situation occurring in harmonic systems.
Fig.~\ref{singleU8}(a) demonstrates on the other hand that $T_{M}$ has to be high enough to be able to activate
the vibrational states of the anharmonic band. This can be understood by invoking a Landauer-like equation to calculate heat
flux in the ballistic regime.~\cite{Zhong09}
\begin{equation}
J\sim\int \hbar \omega [n_C(\omega)-n_H(\omega)] \tau(\omega) d\omega.
\label{LandauerFlux}
\end{equation}
Here $\tau(\omega)$ refers to the transmission coefficient, being $\omega$ the frequency of every vibrational state,
and the mode distribution of the heat baths are considered classical within the Maxwell-Boltzmann statistics
$n_{R,L}(\omega)=\exp(-\hbar \omega/K_B T_{R,L}^r)$. In the limit of linear response, $\delta T<<T_{M}$,
the thermal conductivity can be written as:~\cite{Zimmermann08}
\begin{equation}
\kappa=\frac{J}{2 \delta T}\sim \int x^2 \exp(-x) \tau(x) dx.
\label{LandauerCond}
\end{equation}
The integral in Eq.~\ref{LandauerCond}, being $x=\hbar \omega/K_B T_{M}^r$, is given by the product between the transmission coefficient and
a weight function $F(T_{M})=x^2 \exp(-x)$,
which accounts for bath thermal effects. This means
that $F(T_{M})$ defines which vibrational system modes can be excited by the heat reservoirs and contribute to the heat flux.
Right panels of Fig.~\ref{singleU8} show the numerical density of vibrational states (DOS), calculated by Fourier transforming the velocity autocorrelation
function of equilibrium molecular dynamics simulations, namely, $T_C=T_H=T_{M}$. We also plot the
weight function $F(T_{M})$ for the $T_{M}$ considered in the simulations.
It is clear that for low $T_M$ there is very few states under the curve
$F(T_{M})$, and therefore, only those states are involved in heat transfer along the system, see left panel of Fig.~\ref{singleU8}(a).
The lack of heat carrying modes gives rise to a stationary thermal profile deviated from the linear thermal gradient and thus,
Fourier's law is not expected to be valid.
To get further insight into the behavior of the nonlinear system the self-consistent phonon theory (SCPT) can be considered.
This approach consists in replacing the anharmonic potential for a harmonic approximation such that the new effective harmonic
strength is temperature-dependent as follows:~\cite{Dauxois93,Shao08}
\begin{equation}
V(x_n)=\frac{-V_0}{4\pi^2}\cos (2\pi x_n)\rightarrow \frac{U(T)}{2}x_n^2\ ,
\label{SCPTpot}
\end{equation}
By performing a variational study,~\cite{Dauxois93} the effective interaction
$U(T)$ for the Hamiltonian ${\cal H}_{A}$ can be obtained by solving the following transcendental equation:
\begin{equation}
U(T)=V_0 \exp\Big(\frac{-2 T\pi^2}{ \sqrt{U(T)(U(T)+4)}}\Big)
\label{SCPT}
\end{equation}
Once $U(T)$ is known, the dispersion relation of the system, $\bar \omega_A^2(p,T)= U(T)+ 2 (1-cos(p))$, being $p\in[0,\pi]$,
and the DOS for the effective harmonic system can be defined analytically. Hereafter $\bar \omega$ refers to frequencies expressed in units of the fundamental frequency $\omega_0=k/m$
of the system.
The frequency band of the SCPT for $V_0=8$ as a function of the temperature is
presented in the shadowed region of the inset of Fig.~\ref{singleU8}.
The comparison between the frequency band defined by $\bar \omega_A^2(p,T)$ and the numerical DOS calculated in equilibrium,
is also shown in the right panel of Fig.~\ref{singleU8} by considering
the solution of Eq.~\ref{SCPT} at $T=T_{M}$.
Figure~\ref{singleU8} shows from top to bottom that the anharmonicity shifts the
DOS to lower frequencies when $T_M$ is increased, as predicted by the SCPT approach.
However, it is well known that the SCPT approximation is not able to predict the broadening of the frequency
band when thermal effects become more relevant at higher temperatures, see Fig.~\ref{singleU8}(c).
In view of these arguments, we want to stress two conclusions which will be taken into account hereafter.
On one hand it is clear that $T_M$ in the system should be high enough to activate
the anharmonic modes in order to create a well formed thermal gradient along the lattice in presence of the thermal baths.
However, in order to have an accurate description of the thermal properties based on the SCPT, $T_M$ should be low enough
so that thermal effects are kept within this approach validity.
For the sake of clarity we will focus on temperature regimes which fulfilled these two requirements hereafter.
\section{Ladder-model}
\label{double}
Now that we have reviewed some of the characteristics of an anharmonic one dimensional lattice,
we focus on the main system of interest in this work, the ladder-model presented in Fig.~\ref{sketch},
whose Hamiltonian is described by Eq.~\ref{Hamiltonian}.
Notice that the heat reservoirs at temperatures $T_C$ and $T_H$ are connected to both edges
of the harmonic chain. Therefore, the dynamics of sites $y_1$ and $y_N$ will be described
by Eqs.~\ref{NewtonLang} accordingly.
As previously mentioned, a well formed thermal gradient cannot arise along an isolated harmonic chain connected
to two heat baths at different temperatures. However, its thermal properties are expected to change
by switching on its interaction with the anharmonic chain in our system. Thus, by
increasing $k_{int}$ the effect of the anharmonicity will be intensified in the whole ladder-system.
Our results demonstrate that for a large enough interchain interaction, the vibrational spectra of both chains become mixed,
and a linear thermal profile will arise in the system.
This transition is shown in left panels of Fig.~\ref{double} for a system of $N=100$~sites, where the anharmonicity
is $V_0=5$ and two interchain interactions, $k_{int}=0.05$ and $k_{int}=1.00$ are considered.
Figure~\ref{double}(c) also shows that the thermal conductivity along the lattice decays as a function of $k_{int}$.
This is a result of the fact that the contribution from the anharmonic chain to the double-lattice dynamics is stronger for a larger coupling
and therefore, the thermal resistivity is higher.
\begin{figure}[ht]
\centerline{\includegraphics[width=90mm,clip]{Figgrad3.eps}}
\caption{Left panels show the stationary temperature profile calculated by numerical simulations in a ladder-model
(Fig.~\ref{sketch}) of $N=100$ sites connected to two independent heat reservoirs at different temperatures, $T_{C}=0.1$ and $T_{H}=0.2$.
The temperature profile is plotted for the harmonic $T_n(H)$ and the anharmonic $T_n(A)$ chain, and the coupled system $T_n$.
The considered parameters are $V_0=5$ and a) $k_{int}=0.05$ and b) $k_{int}=1.00$.
Right panels show the finite-size thermal conductivity as a function of c) the coupling $k_{int}$ and d) the system size $N$ for $k_{int}=0.05$ and $k_{int}=1.00$.}
\label{double}
\end{figure}
More interestingly, the presence of the thermal gradient
for a large enough coupling is associated to a finite-size thermal conductivity which does not depend on the system size, and
therefore, Fourier's law is expected to be valid, see Fig.~\ref{double}(d).
In order to give an estimation of the threshold interaction $k_{int}^*$ necessary
to create a linear thermal gradient in the ladder-model, we will consider a
fully harmonic ladder-system, whose vibrational spectrum consists of two bands, an acoustic and an optical one.
In particular the optical band will be at $\bar \omega_H^2(p,k_{int})= 2k_{int} + 2 (1-cos(p))$, namely,
it will be shifted to higher frequencies by increasing the $k_{int}$.
Note that in the real system under consideration (Fig.~\ref{sketch}) for small couplings the anharmonic band will play the
role of the optical band in the harmonic ladder-model.
Once the interaction is large enough so that the anharmonic lattice
can be affected by the heat reservoirs and thus, it can contribute to establish the thermal gradient
it will present a band of anharmonic states at frequencies $\bar \omega_A^2(p,T_{M})= U(T_{M})+ 2 (1-cos(p))$.
This means that, the higher the temperature affecting the anharmonic lattice
is, the lower the frequencies of the activated anharmonic modes are,
see Sec.~\ref{single}.
Therefore, we will define the threshold coupling such that the bands defined by
$\bar \omega_H^2(p,k_{int})$ and $\bar \omega_A^2(p,T_{M})$ will have spectral overlap.
\begin{figure}[ht]
\centerline{\includegraphics[width=80mm,clip]{U8new2.eps}}
\caption{Left panels show the stationary temperature profile for the harmonic $T_n(H)$ and the anharmonic chain $T_n(A)$
calculated by numerical simulations in a ladder-model (Fig.~\ref{sketch})
of $N=100$ sites connected to two heat reservoirs at $T_{C}=0.1$ and $T_{H}=0.2$.
Right panels present the theoretical DOS of a single FK chain affected by a mean temperature $T_A$ according to the SCPT,
as well as the DOS of the optical band of a fully harmonic ladder-model for several $k_{int}$.
The considered parameters are $V_0=8$ and a) $k_{int}=0.1$, b) $k_{int}=0.25$ and c) $k_{int}=1.75$.}
\label{evol}
\end{figure}
According to several simulations, we have concluded that the overlap condition
between the bands $\bar \omega_H^2(p,k_{int})$ and $\bar \omega_A^2(p,T_{M})$
depends on the temperature present in the system.
This is related to the thermal broadening effects mentioned in Sec.~\ref{single}.
When these effects, not included in SCPT, are negligible, we have numerically tested
that the thermal gradient arises if approximately half of the states are common to the two considered bands.
Thus, the interaction threshold can be estimated by:
\begin{equation}
k_{int}^*=\frac{U(T_{M})-2}{2}\ .
\label{threshold}
\end{equation}
Figure~\ref{evol} shows the formation of the thermal gradient along the ladder-model by increasing $k_{int}$
for a system of $N=100$~sites and $V_0=8$. The right panels present the theoretical DOS corresponding to the band $\bar \omega_A^2(p,T_{A})$
of a single anharmonic lattice affected by an averaged temperature $T_{A}=\sum_nT_n(A)/N$
obtained in the simulations for each coupling. Additionally, the DOS of the theoretical optical
band for the fully harmonic ladder-model $\bar \omega_H^2(p,k_{int})$ for the considered couplings is plotted.
In Fig.~\ref{double}(c), it is shown a well-formed thermal gradient along the system for the coupling $k_{int}^*=1.75$,
for which the DOS of $\bar \omega_H^2(p,k_{int}^*)$ and $\bar \omega_A^2(p,T_{A})$ overlap in 50\% of their area.
In Fig.~\ref{critic} we show a good agreement between our prediction of $k_{int}^*$ according to Eq.~\ref{threshold} and the threshold
interaction obtained by numerical simulations for a system size of $N=100$~sites and different values of $V_0$.
\begin{figure}[ht]
\centerline{\includegraphics[width=60mm,clip]{Fig3.eps}}
\caption{Critical coupling $k_{int}^*$ necessary to create a thermal gradient in the ladder-system
(Fig.~\ref{sketch}) of $N=100$ sites as a function of the anharmonicity $V_0$. Solid line represents the theoretical prediction Eq.~\ref{threshold} in comparison
with the values for $k_{int}^*$ numerically obtained, solid dots. In all cases $T_{C}=0.1$ and $T_{H}=0.2$ have been considered}
\label{critic}
\end{figure}
\section{Heat rectification}
\label{rect}
In previous works it has been well established that symmetry breaking and nonlinearity
are the necessary ingredients in a system to support heat rectification.~\cite{Terraneo02,Hu06,Li06,Zhang10} In this regard,
we propose a new thermal rectifying mechanism present in the ladder-model under consideration.
On the one hand, it is clear that nonlinearity comes from the anharmonicity of one of the chains.
On the other hand, the symmetry breaking occurs by considering different connections for the heat reservoirs,
in a similar way which has been considered in the case of charge transport in DNA.~\cite{RafaelRect}
Figure~\ref{figsrect} shows the two considered baths configuration where $T_H$ and $T_C$ refer to
the temperature of the hot and the cold heat reservoirs as previously.
\begin{figure}[ht]
\centerline{\includegraphics[width=60mm,clip]{LURD2.eps}}
\centerline{\includegraphics[width=60mm,clip]{LDRU2.eps}}
\caption{Schematic view of a fragment of the ladder-model under consideration connected to two
independent heat reservoirs at $T_H$ and $T_C$ in two different configurations. Notice that in a)
the hot bath is connected to the anharmonic chain and in b) this connection is made to the harmonic one.
Accordingly heat current flows in opposite directions, $J_+$ and $J_-$, respectively.}
\label{figsrect}
\end{figure}
Notice that in Fig.~\ref{figsrect}(a)
the hot reservoir is connected to the left edge of the anharmonic chain while in Fig.~\ref{figsrect}(b) the position of the
baths connection is reversed. This means that the heat current flows in opposite directions for both cases
but more interestingly, our results show that these currents reach different values in the stationary state, and therefore the system behaves as a thermal rectifier.
We perform molecular dynamics simulations by connecting the ladder-model described by the Hamiltonian Eq.~\ref{Hamiltonian} to two
heat reservoirs. In Fig.~\ref{figsrect}(a) the dynamics of sites $x_1$ and $y_N$ will be described
by Eqs.~\ref{NewtonLang} under Langevin baths at temperatures $T_H=0.2$ and $T_C=0.1$ respectively. In Fig.~\ref{figsrect}(b)
the connections are reversed and sites $y_1$ and $x_N$ are the one affected by Eqs.~\ref{NewtonLang} accordingly.
In order to analysis the main features of the heat rectification, we will numerically calculate the thermal conductivity,
a) $\kappa_+ =J_+ N/(T_H-T_C)$ and b) $\kappa_- = J_- N/(T_C-T_H)$, for several couplings $k_{int}$. Notice that the interchain interaction has revealed as
the key parameter to change the thermal properties of the ladder-model in Sec.~\ref{double}. These results are presented in Figure~\ref{FigRect}(a)
for $V_0=5$ and two different system sizes $N=100$ and $N=200$.
\begin{figure}[ht]
\centerline{\includegraphics[width=90mm,clip]{Figrect.eps}}
\caption{a) Thermal conductivity for the two cases presented in Fig.~\ref{figsrect}, $\kappa_+$ and $\kappa_-$,
as a function of the interchain coupling with $V_0=5$ and for two different system sizes $N=100$ and $N=200$.
b) System size dependence of the thermal conductivity difference $\Delta_{max}$ for $V_0=5$.
c) Normalized thermal conductivity difference $\Delta/\Delta_{max}$ as a function of the coupling $k_{int}$ for $N=100$
and two different anharmonicities $V_0=5$ and $V_0=8$.
}
\label{FigRect}
\end{figure}
Our results establish the existence of heat rectification, $\kappa_+>\kappa_-$, for a certain range of the interchain interaction $k_{int}$.
The difference between both conductivities $\Delta=\kappa_+-\kappa_-$ reaches a maximum $\Delta_{max}$ for a particular coupling
$k_{int}^{0}\sim 0.25$, where we find a strong rectification effect of 35\%. Qualitatively we can understand this rectifying behavior
by the following handwaving argument. It is clear that if there is no coupling between both chains, the system cannot reach any stationary state,
since the heat current is not able to flow from one bath to the other. However, when the coupling is switched on,
new channels connecting both chains allow for the heat flow from the hot reservoir to the cold one and the thermal conductivity increases.
Due to the coupling in a) the heat current flows from a bad conductor, the anharmonic chain at $T_H$,
to a good one, the harmonic chain at $T_C$, while in b) the situation is the opposite one, see Fig.~\ref{figsrect}.
Therefore, in both cases
$\kappa_+$ and $\kappa_-$ increases with $k_{int}$ but since the heat current is deviated
to a more efficient heat conductor in a), $\kappa_+$ is larger than $\kappa_-$. In order to support this argument Fig.~\ref{figsrect}(b) shows that $\Delta_{max}$ increases with
the system size $N$, this means with the number of coupling channels, as expected.
Once a good connection between both baths is established, if $k_{int}>k_{int}^{0}$
according to Sec.~\ref{double}, the effect of the anharmonicity will be intensified
in the ladder-system, and therefore the conductivity is expected to decrease, see Fig.~\ref{double}(c).
When the coupling is so large, that the vibrational spectra of both chains are completely mixed, the rectifying effect
is expected to disappear and thus $\kappa_+=\kappa_-$. In Fig.~\ref{figsrect}(a) we show that this situation occurs for a $k_{int}^{\infty}\sim1.0$.
Last, we would like to address, how the anharmonicity $V_0$ affects the main features of heat rectification in the system.
For the sake of comparison Fig.~\ref{figsrect}(c) shows the magnitude $\Delta/\Delta_{max}$ as a function of $k_{int}$ for
two different anharmonicities, $V_0=5$ and $V_0=8$. Our results show that the qualitative rectification behavior does not depend on $V_0$
but its main features are shifted to larger couplings $k_{int}$. In a similar way, in Sec.~\ref{double} it was already demonstrated that the threshold interaction
$k_{int}^*$ for which the vibrational spectra of the subchains are mixed in the system, increases with the anharmonicity $V_0$, see Fig.~\ref{critic}.
It is to be mentioned that it is not trivial to establish a direct relationship between $k_{int}^{\infty}$ and $k_{int}^{*}$ since the baths connections and
thus, their thermal effects are different in Fig.~\ref{sketch} and Fig.~\ref{figsrect}.
However, according to our simulations it turns out that $k_{int}^{\infty}>k_{int}^{*}$, and therefore, we can predict that the main heat rectification effects will appear
for $k_{int}<k_{int}^{*}$.
\section{Conclusions}
\label{conclusions}
In this paper, we have studied heat conduction along a ladder-model
consisting of a harmonic and an anharmonic one dimensional chain coupled by harmonic interactions.
We have analyzed how the thermal properties of the coupled ladder-system depends on the strength of the coupling.
In particular, by connecting two independent heat reservoirs to the harmonic counterpart of the system,
we have demonstrated that for a large enough interchain interaction $k_{int}$,
a thermal gradient can be formed along the system and the thermal conductivity will remain constant for increasing size systems.
This means that the ladder-system will present normal heat transport contrary to the case of an isolated harmonic system.
We have estimated the threshold interaction of this transition by considering the self-consistent phonon theory for a single anharmonic chain
in comparison to a fully harmonic ladder-model. Our estimations has been proven to be in good agreement with numerical results
based on molecular dynamics simulations.
Taking advantage of the nonlinear effects present in our system, it was shown that
heat current along the ladder-model behaves differently if the heat baths connections are interchanged between the harmonic and the anharmonic chain as
in Fig.~\ref{figsrect}. This mechanism reveals strong heat rectification effects of more than 30\%.
The main qualitative features of the rectifying device do not depend
on the system size nor on the anharmonicity present in the system. However, the maximum rectification rates increases with the system size
and shifts to larger couplings in the case of stronger anharmonic effects.
\acknowledgments
This work has been supported by DFG-Projekt CU 44/20-1, MCINN-Project MAT2010-17180, and by the South Korea
Ministry of Education, Science and Technology Program "World Class University" under contract R31-2008-000-10100-0.
E. D. acknowledges financial support by Ministerio de Eduacion y Ciencia and the program Flores Valles-UCM.
We acknowledge fruitful discussions with H. Sevincli and S. Avdoshenko.
|
\section{Introduction}
The distance of only $\sim$50 kpc to SN 1987A makes a number of unique
observations possible. In particular, the supernova can
after more than 20 years still be observed as resolved ejecta, glowing from
radioactive input.
About four years after exploding, the supernova entered the $^{44}$Ti-dominated phase. Since then, its emission has been that of a cool ($\sim$$10^2$ K) gas powered by radioactive decays, complemented by freeze-out emission from the hydrogen envelope. Owing to the long life-time of $^{44}$Ti ($\sim$85 years), the character of the spectrum has changed little since this transition
To understand and interpret the late-time emission from the ejecta, detailed spectral synthesis modeling is needed.
The spectral formation process in supernovae is complex, especially at late times when all important processes are non-thermal, but radiative transfer effects may still be important. High-energy gamma-rays and positrons from the radioactive decays deposit their energy into free and bound electrons, producing a population of fast electrons which heat, excite and ionize the gas. The ionizations produce secondary electrons that additionally contribute to these processes. The degradation process is quantified by solving the Boltzmann equation for each of the zones present in the supernova \citep[e.g.][KF92 hereafter]{Xu1991ED, Kozma1992}. Once the deposition in the various channels have been determined, the temperature, ionization structure and NLTE level populations can be computed by solving the equations for thermal and statistical equilibrium or their time-dependent variants \citep[e.g.][KF98 a, b hereafter]{deKool1998, Kozma1998I, Kozma1998II}. Because the ionization balance affects the solution for the radioactive deposition, the equation systems are coupled and a solution is found by iteration.
The equilibrium also depends on the radiation field, which should simultaneously be solved for. The strong velocity gradients in supernovae imply unique radiative transfer effects.
As they travel through the ejecta, photons are continuously red-shifting with respect to the comoving frame, and are therefore exposed to absorption in lines over a broad wavelength range. Because especially iron-group elements have a large number of lines at UV/blue wavelengths, radiation here experiences a complex transfer and a significant fraction of it emerges by fluorescence as a quasi-continuum at longer wavelengths \citep{Li1996}. The efficient blocking by atomic lines is equivalent to an effective continuum opacity known as line blocking. The emerging UV emission may often be due to 'holes' in this line blocking where radiation can escape \citep{Mazzali1993}. Although the optical depths decrease over time, the multi-line transfer is important for modeling supernova spectra for many years or even decades after explosion.
The observational determination of the masses of the three main
radioactive isotopes \iso{56}Ni, \iso{57}Ni, and \iso{44}Ti constitutes
one of the main constraints on explosion models of core collapse
supernovae. The production of these isotopes is sensitive to both density and temperature, and thus
to the explosion dynamics \citep[e.g.][]{Woosley1991}. During the first $\sim$500 days the decay of \iso{56}Co (which is the product of rapid \iso{56}Ni decay) dominated the
bolometric light curve of SN 1987A. From the light curves, the \iso{56}Ni mass could
be determined to be $0.069\ \hbox{$~\rm M_{\odot}$}$
\citep{Bouchet1991a}. The determination of the
\iso{57}Ni mass was complicated by the fact that when this isotope became important, the bolometric light curve
was already affected by the delayed release of ionization energy
(freeze-out). A time-dependent modeling that took this into account resulted in an estimated \iso{57}Ni mass of $\sim$$3.3\e{-3}
\hbox{$~\rm M_{\odot}$}$ \citep[][FK93 hereafter]{Fransson1993}, which agrees well with the mass derived from
observations of the infrared Co II/Fe II lines and the gamma-rays emitted in the decay \citep{Varani1990,Kurfess1992}.
The determination of the \iso{44}Ti-mass is
complicated by a number of factors, the first being the freeze-out contribution just as for \iso{57}Ni. Second, at late times the low temperature of the gas and the presence of dust lead to most of the deposited energy being re-radiated in the largely unobservable far-infrared,
which prohibits a determination of the bolometric luminosity. The \iso{44}Ti-mass must therefore be determined
from detailed modeling of the fraction of the energy that emerges in the (observable) UV/optical/NIR bands, and must consider the time-dependent freeze-out. From this type of modeling, \citet[][FK02 hereafter]{Fransson2002} determined the \iso{44}Ti-mass to $(0.5-2)\e{-4}$ M$_\odot$, which agrees with the $(1-2)\e{-4}$ M$_\odot$~estimated from a nebular analysis of the eight-year spectrum by \citet[][C97 hereafter]{Chugai1997}. \citet{Lundqvist2001} derived an upper limit of $1.1\e{-4}$ M$_\odot$~, based on the non-detection of the [Fe II] 26 $\mu$m line by ISO, which rested on the assumption of local positron deposition and the absence of any dust cooling.
Another possibility for determining the \iso{44}Ti-mass is by direct detection of the gamma-rays produced in the decay. Only in one case, Cas A, has such a detection been made \citep{Iyudin1994,Vink2001}, with the mass estimated to
$1.6^{+0.6}_{-0.3} \times 10^{-4} \hbox{$~\rm M_{\odot}$}$ \citep{Renaud2006}. For Cas A we do not
know the masses of the \iso{56}Ni and \iso{57}Ni isotopes though.
For SN 1987A, the detection limits by INTEGRAL
resulted in only a weak upper limit of $1\e{-3}\ \hbox{$~\rm M_{\odot}$}$ for the \iso{44}Ti-mass
\citep{Shtykovskiy2005}.
In this paper we present a detailed model of the UV/optical/NIR spectrum of SN 1987A at an age of eight years, using a 19 M$_\odot$~explosion model as input. Using a new code with a detailed radiative transfer treatment, our objectives are
to see if the spectral model can reproduce the main features in the observed spectrum, to understand the contributions by various zones and elements to the spectral formation process, to quantify the importance of the (non-local) line transfer, and to determine the \iso{44}Ti-mass as accurately as possible. Previous discussions of the late-time ejecta spectra can be found, e.g., in \citet{Wang1996} and C97. Our paper also serves to describe the code we developed for reference in future modeling.
\section{Modeling}
\label{sec:modeling}
In KF98 a, b, a self-consistent model for the nebular phase spectrum of SN 1987A was presented. This included
a detailed calculation of the gamma-ray/positron thermalization and a
determination of the time-dependent ionization, excitation, and temperature
structure of the ejecta. Comparing the resulting spectra from two different explosion models, the evolution of all major emission lines
were analyzed. \citet{Fransson2002} contains an update of this work, where
the modeling of some of the most important lines and
the broad band photometry were discussed.
A shortcoming in these models was that (non-local) radiative transfer in atomic lines was not included, which introduced an
uncertainty for the internal radiation field and thereby for the ionization balance, and also for the emerging spectrum.
Here, we calculate in detail the line scattering and fluorescence throughout the spectrum (Sect. \ref{sec:RT}), which significantly improves the accuracy of the model. In addition, we updated the atomic
physics by adding new atomic data and including more and larger model atoms.
While we do a more accurate radiative transfer calculation, this
comes at the expense of time-dependent effects as our model assumes steady-state. In
FK93 it was shown that in the envelope the recombination time scale became comparable to the radioactive
decay time scale at $\sim$800 days. From that time on, much of the
ionization energy was not radiated instantaneously but on a longer time scale, and a time-dependent calculation was
therefore necessary. In the
\iso{44}Ti-dominated phase the radioactive time scale becomes much
longer, $\sim$85 years \citep{Ahmad2006}, and is less
relevant. Instead, it is the adiabatic expansion time scale, i.e. the age of
the supernova, that is the most relevant, and should be compared
to the recombination and cooling time scales. In zones where a time-dependent treatment is necessary, we use the solutions obtained by an updated version of the time-dependent code used in KF98 a, b for
the ionization balance. As we show below (Sect. \ref{Res:Tandion}), this is only necessary in the hydrogen envelope.
\subsection{Ejecta model}
\label{sec:em}
We base our ejecta model on the 19 M$_\odot$~explosion model computed by \citet[][WH07 hereafter]{Woosley2007}. From the distribution of elements in this model, we construct seven types of zones; Fe/He, Si/S, O/Si/S, O/Ne/Mg, O/C, He, and H, where the zones are named after their dominant elements. The mass cuts for the zones are taken to be where the most common or second most common element changes.
The mass of the Fe/He zone is adjusted to give a total \iso{56}Ni mass of 0.072 M$_\odot$~in the ejecta (the value of 0.069 M$_\odot$~from \cite{Bouchet1991a} corrected for the slighly larger distance we use here), and the mass fractions of \iso{57}Co and \iso{44}Ti in this zone are then adjusted to give a total \iso{57}Co mass of of $3.3\e{-3}$ (FK93) and a total \iso{44}Ti mass of $1.5\e{-4}$\Mo, which we later show to be our favored value for this isotope. The H zone mass is adjusted to match the observed estimate of $\sim$8 M$_\odot$~(KF98 b). Finally, we replace the He/C zone in the explosion model with a He/N zone, that has not experienced any outward mixing of carbon. This is also based on modeling in KF98 b, where it was found that a significant He/C zone would result in [C~I] lines much stronger than the observed ones.
The masses and compositions of the zones in the model are given in Table \ref{table:chem}. Abundances are taken as the value in the middle of each zone. While this does not exactly conserve the total element masses, the differences are small compared to using zone-averaged values, and the middle values should give a more consistent description for the typical composition.
Because the WH07 models are based on solar metallicity progenitors, we replace the abundances of all elements with mass fractions differing by less than a factor 1.25 from the solar values with the LMC abundances taken from \citet{Russell1992}. A comparison of models with different metallicities in \citet[][WW95 hereafter]{WW95} shows that most nucleosynthetic yields only weakly depend on metallicity, so the abundances for the processed elements should be very similar for solar and LMC metallicity progenitors. Observations of the circumstellar material indicate that CNO-burning products have been mixed out from the helium core into the hydrogen envelope \citep{Fransson1989CNO}. We use the abundances of He, N and O for the envelope derived by \citet{Mattila2010} and of C by \citet{Lundqvist1996} (see Table \ref{table:chem})
We model the supernova as a spherically symmetric, homologously expanding nebula, divided into a core
and an envelope. The core contains the heavy element zones (Fe/He, Si/S, O/Si/S, O/Ne/Mg, O/C), and
fractions of the He and H zones, which are mixed inwards during the explosion. These fractions were chosen to be 25\% (see Sect. \ref{sec:thecore}). The envelope contains the remaining 75\% of the He and H zones.
The extent of the core can be estimated from the widths of
the lines that are emitted mainly from the heavy element zones, for example the iron lines and [O~I] $\lambda \lambda$ 6300, 6364. The oxygen lines have maximum expansion velocities of $\sim$1700 \hbox{km\,s$^{-1}$}, whereas the iron, cobalt, and nickel lines extend to $\sim$2500 \hbox{km\,s$^{-1}$}, with some contribution even out to $\sim$ 3500 \hbox{km\,s$^{-1}$} \citep[see][for a review]{McCray1993}.
Our single-core model forces a compromise value,
which we choose as $V_{\rm core}=2000~$\hbox{km\,s$^{-1}$}, the same as in KF98 a, b. A higher core velocity means less gamma-ray deposition, but because the positrons dominate the energy input at late times, this choice is not critical for the energy budget here.
\subsubsection{The core}
\label{sec:thecore}
\begin{table*}
\caption{Chemical compositions (mass fractions) of the zones used in the model.}
\label{table:chem}
\centering
\begin{tabular}{r l l l l l l l l}
\hline\hline
Zone & Fe/He & Si/S & O/Si/S & O/Ne/Mg & O/C & He & H\\
\hline
Mass (M$_\odot$): & 0.092 & 0.14 & 0.16 & 1.9 & 0.58 & 1.3 & 8.0\\
\hline
Mass fractions:\\
\hline
$^{56}$Ni & 0.75 & 0.022 & 0 & 0 & 0 & 0 & 0\\
$^{57}$Co & 0.034 & $9.0\e{-4}$ & $1.0\e{-5}$ & 0 & 0 & 0 & 0 \\
$^{44}$Ti & $1.4\e{-3}$ & $5.2\e{-5}$ & $1.6\e{-5}$ & 0 & 0 & 0 & 0\\
\hline
H & $2.3\e{-6}$ & $4.2\e{-8}$ & $8.7\e{-9}$ & $4.6\e{-10}$ & $2.1\e{-10}$ & $2.5\e{-7}$ & 0.60\\
He & 0.14 & $8.3\e{-6}$ & $4.3\e{-6}$ & $2.0\e{-6}$ & 0.078 & 0.99 & 0.40 \\
C & $3.6\e{-7}$ & $1.4\e{-6}$ & $8.5\e{-5}$ & 0.031 & 0.28 & $3.0\e{-4}$ & $2.4\e{-4}$** \\
N & $1.4\e{-6}$ & $1.1\e{-9}$ & $1.8\e{-5}$ & $3.2\e{-5}$ & $4.7\e{-6}$ & $9.0\e{-3}$ & $2.2\e{-3}$** \\
O & $9.7\e{-6}$ & $1.6\e{-5}$ & 0.46 & 0.70 & 0.62 & $1.6\e{-4}$ & $1.8\e{-3}$**\\
Ne & $1.0\e{-5}$ & $1.6\e{-6}$ & $1.4\e{-4}$ & 0.22 & 0.018 & $1.1\e{-3}$ & $6.1\e{-4}$*\\
Na & $6.0\e{-7}$ & $1.0\e{-6}$ & $1.2\e{-6}$ & $7.0\e{-3}$ & $1.9\e{-4}$ & $1.8\e{-4}$ & $1.6\e{-5}$*\\
Mg & $1.2\e{-5}$ & $1.6\e{-4}$ & $7.4\e{-4}$ & 0.042 & $4.9\e{-3}$ & $5.4\e{-4}$* & $5.4\e{-4}$*\\
Al & $2.1\e{-5}$ & $2.6\e{-4}$ & $1.9\e{-4}$ & $4.0\e{-3}$ & $3.0\e{-5}$* & $3.0\e{-5}$* & $3.0\e{-5}$*\\
Si & $1.6\e{-4}$ & 0.44 & 0.27 & $3.1\e{-3}$ & $2.8\e{-4}$* & $2.8\e{-4}$* & $2.8\e{-4}$*\\
S & $1.2\e{-4}$ & 0.39 & 0.21 & $2.6\e{-4}$ & $1.2\e{-4}$* & $1.2\e{-4}$* & $1.2\e{-4}$*\\
Ar & $1.2\e{-4}$ & 0.051 & 0.043 & $8.4\e{-5}$ & $5.4\e{-5}$* & $5.4\e{-5}$* & $5.4\e{-5}$*\\
Ca & $1.3\e{-3}$ & 0.031 & 0.012 & $3.1\e{-5}$ & $2.3\e{-5}$* & $2.3\e{-5}$* & $2.3\e{-5}$*\\
Sc & $2.3\e{-7}$ & $2.3\e{-7}$ & $1.2\e{-6}$ & $1.6\e{-6}$ & $5.6\e{-7}$ & $1.4\e{-8}$* & $1.4\e{-8}$*\\
Ti (stable) & $8.6\e{-4}$ & $2.9\e{-4}$ & $8.4\e{-5}$ & $6.2\e{-6}$ & $5.3\e{-6}$ & $2.3\e{-6}$* & $2.3\e{-6}$* \\
V & $2.1\e{-5}$ & $1.3\e{-4}$ & $3.0\e{-6}$ & $6.4\e{-7}$ & $4.5\e{-7}$* & $4.5\e{-7}$* & $4.5\e{-7}$*\\
Cr & $1.3\e{-3}$ & $3.2\e{-3}$ & $4.1\e{-5}$ & $1.3\e{-5}$ & $1.2\e{-5}$* & $1.2\e{-5}$* & $1.2\e{-5}$*\\
Mn & $1.0\e{-5}$ & $6.6\e{-4}$ & $1.9\e{-6}$ & $3.8\e{-6}$ & $6.6\e{-6}$* & $6.6\e{-6}$* & $6.6\e{-6}$*\\
Fe (stable) & $2.3\e{-3}$ & 0.050 & $4.3\e{-4}$ & $6.9\e{-4}$ & $7.2\e{-4}$* & $7.2\e{-4}$* & $7.2\e{-4}$*\\
Co (stable) & $3.2\e{-8}$ & $5.5\e{-9}$ & $1.4\e{-6}$ & $1.9\e{-4}$ & $1.3\e{-4}$ & $2.4\e{-6}$* & $2.4\e{-6}$*\\
Ni (stable) & 0.037 & $2.4\e{-3}$ & $1.2\e{-3}$ & $5.6\e{-4}$ & $2.3\e{-4}$ & $4.8\e{-5}$* & $4.8\e{-5}$*\\
\hline
Mean atomic weight & 19.4 & 31.6 & 21.5 & 16.9 & 12.1 & 4.0 & 1.4 \\
\hline
\end{tabular}
\tablefoot{
Primordial LMC abundances are marked with an asterisk, and CNO abundances derived from observations of the inner circumstellar ring are marked with two asterisks. See also Sect. \ref{sec:em}. \\
}
\label{table:zoneabund}
\end{table*}
We assume that the zones in the core experience complete macroscopic mixing, where they fragment into $N_{\rm cl}$ clumps that become completely mixed in velocity space (see Fig. \ref{fig:ejectamodel}). Support for this macroscopic mixing comes from theoretical considerations \citep[e.g.][]{Mueller1991}, multi-dimensional simulations \citep[e.g.][]{Herant1991,Mueller1991,Kifonidis2006,Hammer2010}, the early emergence of X-rays and gamma-rays \citep[][]{Dotani1987, Pinto1988}, and from the similar line profiles of many of the elements \citep[e.g.][]{Spyromilio1990}.
We use a value of $N_{\rm cl}=100$ based on the analysis of the iron clumps in \citet{Li1993iron}. In Sect. \ref{sec:fragmentation} we find that the model spectrum is insensitive to this choice. We further assume that no microscopic mixing occurs \citep[e.g.][]{Fransson1989, Fryxell1991}, so that the zones retain their compositional integrity.
The clumpy structure of the core calls for a special handling of the computational grid. We use a gridless domain for the core region; when a clump emits photons, the position of the clump within the core is selected by a random draw. When we follow photons through the core (Sect. \ref{sec:RT}), exiting one clump leads to a random draw that determines which type of clump to encounter next. We set the probability of encountering any given type of zone to be proportional to the total surface area of that zone type. A random number also selects the impact parameter on the new clump.
The path traveled through zone $i$ is then proportional to $A_i\cdot R_i \propto V_i$, which recovers the desired property that a photon should spend its time in a given zone in proportion to the filling factor of the zone, $\epsilon_i$.
The core structure is a 'virtual' one in the sense
that there is no fixed density grid; we have a set of spherical clumps
which together make up the total volume of the core, but which obviously cannot be
exactly fit together since they are all spheres. There are several advantages to this
virtual grid. The extension of the Monte Carlo philosophy to encompass also
a randomization of the grid is a powerful and appealing way to make computations in clumpy media such as this.
While the radiative transfer can still be done realistically by the use of surface and impact probabilities as described above,
a virtual grid naturally offers a way to capture the situation of macroscopic mixing and to
parametrize the degree of fragmentation of the core. Finally, the line profiles automatically become smooth rather
than jagged, as when using shells.
The filling factors for the oxygen zones are set from the oxygen number density derived from the evolution of the [O~I]\ 6300 \AA/6364 \AA~line ratio, $n_{\rm O}=6.2\e{10}~\mbox{cm}^{-3}$~at 100 days \citep{Spyromilio1991, Li1992}. This allocates $\sim$10\% of the core volume to oxygen clumps.
The iron clumps are believed to expand because of their content of radioactive materials \citep{Woosley1988, Herant1991,Basko1994}, which was verified for SN 1987A by modeling of the cooling lines from the iron clumps \citep{Li1993iron}. The iron zone was found to have a filling factor of $\epsilon_{\rm Fe/He}\gtrsim 0.3$, assuming $V_{\rm core}=2500$ \hbox{km\,s$^{-1}$}. The same density limit using $V_{\rm core}=2000$ \hbox{km\,s$^{-1}$} corresponds to $\epsilon_{\rm Fe/He}\gtrsim 0.6$. We use a value of 0.6 in our modeling here.
In the explosion model, the Si/S zone contains $\sim$5\% of the $^{56}$Ni, and can therefore be expected to expand as well. For this zone, we therefore use a density in between the oxygen zones and the iron zone, chosen to be ten times lower than the oxygen zone density. This gives a filling factor of $\sim$3\%. In Sect. \ref{sec:lineids}, we find support from our modeling for a low density in this zone.
Hydrodynamical simulations by \citet{Fryxell1991} find $\sim$1 M$_\odot$~of hydrogen mixed inside 2000 \hbox{km\,s$^{-1}$}, while \citet{Herant1992} find $\sim$2 M$_\odot$. From modeling of the H$\alpha$ line in SN 1987A, KF98 b estimate the presence of 2.2 M$_\odot$~of H-zone gas inside 2000 \hbox{km\,s$^{-1}$}. Here, we take 2 M$_\odot$~of H-zone material to be mixed inside the core, which corresponds to $\sim$25\% of the total H zone mass. We assume the same fraction (25\%) of the He zone to be mixed into the core as well, giving a He-zone core component of 0.3~M$_\odot$~and a He-envelope component of 1.0~M$_\odot$.
The total mass of the core, including the mixed-in He and H, is then 5.2 M$_\odot$.
The filling factors for the H and He zones in the core are set by allocating the remaining volume so that these zones obtain equal number densities. This gives the H component a filling factor of $\sim$26\% and the He component a filling factor of $\sim$2\%. See also KF98 a for further discussions regarding filling factors.
\subsubsection{The envelope}
Outside the core we attach an envelope consisting of an inner He zone
followed by logarithmically spaced (constant $dR/R$) H-zone shells. As mentioned above, the total He-zone mass is 1.3 M$_\odot$, of which we put $1.0$ M$_\odot$~in the envelope and $0.3$ M$_\odot$~in the core. The total H-zone mass is taken as 8 M$_\odot$~from H$\alpha$ modeling (KF98 b), of which we put $6$ M$_\odot$~in the envelope and $2$ M$_\odot$~in the core (see Sect. \ref{sec:thecore})
We use a density distribution of the H-envelope based on hydrodynamical modeling by \citet{Shigeyama1990} (model 14E1). The validity of this model is confirmed by modeling of the Mg~II \wl2800 and Mg~I \wl2852 features in C97. The density profile can be described as gradually steepening from an initial -2 power law to an asymptotic $\sim$-8 for $V>5000$~\hbox{km\,s$^{-1}$}. We terminate the envelope at a velocity of 10,000 \hbox{km\,s$^{-1}$}.
\begin{figure}[htb]
\centering
\includegraphics[width=0.8\linewidth,clip=true]{./15937_f1.eps}
\caption{Schematic structure of the used ejecta model. The core consists of several chemically distinct zones distributed
as $N_{\rm cl}$ clumps each. The envelope consists of an initial He zone (gray) followed by shells of H-rich gas (white). The densities of the H shells are decreasing outward.}
\label{fig:ejectamodel}
\end{figure}
\begin{table*}
\caption{Properties of the zones in the model.}
\centering
\begin{tabular}{l | l l l l l l}
\hline\hline
Zone & Mass & $V_{\rm in}$ & $V_{\rm out}$ & Filling factor ($\epsilon$) & Density & Rec. time\\
& (M$_\odot$) & (\hbox{km\,s$^{-1}$}) & (\hbox{km\,s$^{-1}$}) & & (cm$^{-3}$) & (years)\\
\hline
Fe/He & 0.092 & 0 & 2000 & 0.6 & $1.8\e{4}$ & 2.5\\
Si/S & 0.14 & 0 & 2000 & $2.9\e{-2}$ & $3.5\e{5}$ & 0.8 \\
O/Si/S & 0.16 & 0 & 2000 & $4.1\e{-3}$ & $4.2\e{6}$ & 0.4\\
O/Ne/Mg & 1.9 & 0 & 2000 & $7.3\e{-2}$ & $3.5\e{6}$ & 0.9 \\
O/C & 0.58 & 0 & 2000 & $2.0\e{-2}$ & $5.6\e{6}$ & 0.9\\
He-core & 0.32 & 0 & 2000 & $1.5\e{-2}$ & $1.3\e{7}$ & 0.5\\
H-core & 2 & 0 & 2000 & 0.26 & $1.3\e{7}$ & 2.1\\
\hline
He-env & 1.0 & 2000 & 2200 & 1 & $1.7\e{6}$ & 1.3\\
H-env-1 & 1.9 & 2200 & 2730 & 1 & $2.5\e{6}$ & 6.9\\
H-env-2 & 2.0 & 2730 & 3390 & 1 & $1.4\e{6}$ & 10 \\
H-env-3 & 1.2 & 3390 & 4210 & 1 & $4.3\e{5}$ & 17 \\
H-env-4 & 0.61 & 4210 & 5230 & 1 & $1.2\e{5}$ & 29 \\
H-env-5 & 0.24 & 5230 & 6490 & 1 & $2.4\e{4}$ & 49\\
H-env-6 & 0.083 & 6490 & 8050 & 1 & $4.3\e{3}$ & 105\\
H-env-7 & 0.028 & 8050 & 10000 & 1 & $7.6\e{2}$ & 332\\
\hline
\end{tabular}
\tablefoot{
The last column shows the recombination time obtained from a steady state solution (see Sect. \ref{Res:Tandion}).}
\label{table:zonemasses}
\end{table*}
\subsection{Energy deposition and degradation}
\label{sec:edep}
We calculate the deposition of gamma-rays emitted by the clumps containing \iso{56}Co, \iso{57}Co and \iso{44}Ti. Based on the simulations by \citet{Colgate1980, Woosley1989} and \citet{Fransson1989} we assume absorption with effective opacities of $\kappa(^{56}\mbox{Co})=0.030\left(\frac{\bar{Z}}{\bar{A}}/0.5\right)$,~$\kappa(^{57}\mbox{Co})=0.072\left(\frac{\bar{Z}}{\bar{A}}/0.5\right)$ and $\kappa(^{44}\mbox{Ti}) = 0.040\left(\frac{\bar{Z}}{\bar{A}}/0.5\right)$~cm$^{2}$g$^{-1}$, where $\bar{Z}$ and $\bar{A}$ are the average nuclear charges and atomic weights in each zone. We do not include other radioactive isotopes such as \iso{60}Co and \iso{22}Na, whose abundances are highly uncertain. We discuss the possible effects of these in Sect. \ref{sec:discussion}.
From each zone containing radioactive material, we emit and track \iso{56}Co, \iso{57}Co, and \iso{44}Ti gamma-ray packets
as described for the UV/optical/NIR radiation in Sect. \ref{sec:RT}, with the difference that these packets travel in straight paths until they reach the edge of the nebula, depositing a fraction $\exp{\left(-\kappa \rho_{\rm i} l\right)}$ of their energy for each traveled distance $l$, where $\rho_{\rm i}$ is the density of zone $i$. We use radioactive luminosities taken from KF98, with the \iso{44}Ti
values updated for a life-time of 85 years instead of 78 years
After $\sim$2000 days, $^{44}\mbox{Ti}$ completely dominates the deposition. The degradation of the deposited gamma-ray and positron energy into heating, ionizations and excitations is obtained by solving the Spencer-Fano equation using the routine
developed by KF92. For this routine, we updated the collisional cross sections for Fe~II \citep{Ramsbottom2005,Ramsbottom2007} and Ca I \citep{Samson2001}. Ionizations are generally assumed to leave the ions in their ground states, except for O II, for which we compute specific rates to the first few excited states.
In our standard model, the positrons are assumed to be absorbed
on the spot. As noted before, e.g. in C97, this requires a magnetic field. The average positron
energy in the \iso{44}Sc decay is 0.6 MeV and the maximum energy of
the positrons is 1.47 MeV \citep{Browne1986}. The energy loss per unit
length, $dE/dx$, is dominated by excitations and
ionizations and is given by the Bethe formula \eq{ {dE\over dx} = -{4
\pi r_0^2 m_{\rm e} c^2 \rho \over m_{\rm u}\beta^2} {Z\over A} \left[
\ln({\beta^2 m_{\rm e} c^2\over I}) - f_z \right]~,
\label{eq:bethe}}
where $r_0$ is the electron radius, $\beta = V/c$, $I$ is the effective ionization potential of the atom, $f_z$ is a relativistic correction factor, and all other symbols have their usual meaning.
In Fig. \ref{fig:stopping_power} we show the stopping range, $\rho E/
\frac{dE}{dx}$ (g cm$^{-2}$), for the most important elements in the core
together with the total range, averaged over the different
zones in the core.
These ranges should be compared to the mass column density $D$ of the various zones in the core and envelope. For the total core, this is given by
\eqn{D =
0.93\left(\frac{M_{\rm core}}{5 \hbox{$~\rm M_{\odot}$}} \right) \left({V_{\rm core} \over 2000
\ \hbox{km\,s$^{-1}$} }\right)^{-2} \left({t \over \ 8 \ \rm yrs }\right)^{-2}
\ \rm g \ cm^{-2}~. }
The column density for a core with $V_{\rm core}=1800$ \hbox{km\,s$^{-1}$} at 7.87 years is shown as
the upper dashed line in Fig. \ref{fig:stopping_power}. For comparison
we also show the column density of the Fe/He zone as the lower
horizontal line.
That the column density of the Fe/He zone is nearly two orders of magnitude lower than the stopping range for Fe and He means that in the absence of a sufficiently strong magnetic field the positrons will not be trapped in this zone, even if it is not strongly fragmented into many small pieces. The figure also shows, however, that the total core has a sufficiently high column density to trap the positrons even without a magnetic field. There should therefore be no positrons entering the envelope at this or earlier epochs.
\begin{figure}[htb]
\centering
\includegraphics[width=0.8\linewidth]{./15937_f2.eps}
\caption{Positron stopping range for different elements as a function of kinetic energy. The solid black curve denoted 'tot' shows the range weighted by the different zones in the core. The lower horizontal dashed line marks the column density of the Fe/He zone (assuming it to exist as a single clump), and the upper that of the full core for $V_{\rm core} = 1800 ~ \hbox{km\,s$^{-1}$}$. The long-dashed curve shows the cumulative positron distribution.}
\label{fig:stopping_power}
\end{figure}
The presence of magnetic fields in supernova ejecta has been suggested many times, see e.g. \citet{Axelrod1980}.
\cite{Colgate1980} argue that any magnetic field present at the time of the explosion becomes radially 'combed' and will therefore not be important for the trapping \citep[see also][]{Ruiz-Lapuente1998}.
While the positrons should be confined to the core at eight years, it is therefore not clear whether they are absorbed by the zones producing them (which is mainly the Fe/He zone, but also the Si/S zone), or if they can propagate into the other core zones before they are absorbed. We will later investigate this question by comparing the spectra produced in the different scenarios (Sect. \ref{sec:pos}).
\subsection{Statistical and thermal equilibrium}
\label{sec:sate}
For each zone we solve the equations for statistical equilibrium with
respect to the ionization and excitation structure, as well as the
equation of thermal equilibrium. These equations are essentially the ones described in KF98 a, with some
modifications and improved rate calculations described below. For each iteration, the algorithm deployed is to first compute the ionization balance, then iterate solutions to excitation structure and thermal equilibrium until the temperature stabilizes, and finally to calculate the radiative transfer (Sect. \ref{sec:RT}). From the radiative transfer, new photoionization rates, photoexcitation rates (although these are set to zero in this work, see Sect. \ref{sec:excstr}), and radiative heating rates are calculated, and used for the new solutions for ionization, excitation and temperature in the next iteration. Iteration continues until the maximum change in any zone temperature, electron fraction or ion abundance is less than 1\%. We then also verify that the change in the emerging radiation field is neglegible.
\subsubsection{Ionization balance}
Steady-state for the ionization balance is appropriate as long as the recombination time-scale $(1/n_{\rm e}\alpha)$ is short compared to the evolutionary and radioactive time scales. For the densities at the epoch studied here, this is satisfied (Sect. \ref{Res:Tandion}) for all core zones as well as for the envelope helium zone. The hydrogen envelope, however, experiences freeze-out after $\sim$800 days (FK93). For the envelope, we therefore use the solutions for the hydrogen ionization fraction from a time-dependent calculation, based on the model in KF98 a, given in Table \ref{table:freeze}. We do, however, use steady-state solutions for the other elements in the envelope, because the time-dependent model lacks photoionizations by line radiation, and may therefore significantly underestimate the ionization degree for some of them. We also take the hydrogen envelope temperatures from the time-dependent solution, because temperature and ionization balance are coupled.
\begin{table}[h!]
\caption{The degree of hydrogen ionization, and the temperature in the hydrogen envelope at day 2875, taken from a time-dependent calculation based on the model in KF98 a, b.}
\centering
\begin{tabular}{r l l l}
\hline\hline
Velocity & Number density & n(H II)/n(H I) & Temperature\\
(\hbox{km\,s$^{-1}$}) & (cm$^{-3}$) &- & (K)\\
\hline
core & $1.7\e{7}$ & $6.5\e{-5}$ & 119\\
2000 & $3.2\e{6}$ & $2.4\e{-4}$ & 114\\
2400 & $2.5\e{6}$ & $2.5\e{-4}$ & 101\\
2800 & $1.5\e{6}$ & $3.3\e{-4}$ & 93\\
3300 & $5.4\e{5}$ & $7.3\e{-4}$ & 89\\
4100 & $2.4\e{5}$ & $1.3\e{-3}$ & 85\\
4440 & $1.2\e{5}$ & $2.2\e{-3}$ & 84\\
5000 & $4.6\e{4}$ & $4.4\e{-3}$ & 84\\
5500 & $2.2\e{4}$ & $7.5\e{-3}$ & 83\\
6000 & $1.1\e{4}$ & $1.2\e{-2}$ & 81\\
6500 & $5.4\e{3}$ & $1.6\e{-2}$ & 76\\
7000 & $3.0\e{3}$ & $2.0\e{-2}$ & 73\\
7500 & $1.6\e{3}$ & $2.4\e{-2}$ & 70\\
8000 & $9.5\e{2}$ & $2.8\e{-2}$ & 67\\
8500 & $5.7\e{2}$ & $3.2\e{-2}$ & 64\\
9000 & $3.5\e{2}$ & $3.5\e{-2}$ & 64\\
9500 & $2.2\e{2}$ & $3.8\e{-2}$ & 62\\
10,000 & $1.4\e{2}$ & $4.2\e{-2}$ & 60\\
\hline
\end{tabular}
\tablefoot{
Note that the zones here are somewhat differently structured than those we use for the calculations in this paper (Table \ref{table:zonemasses}).
}
\label{table:freeze}
\end{table}
For all other zones, the ionization balance is computed from the steady-state variant of Eq. (10) (and its generalized version) in KF98 a. We include the first three ionization stages of all elements listed in Sect. \ref{sec:excstr}. The main improvement in our treatment here is that we now compute detailed photoionization rates from the Monte Carlo simulation (Sect. \ref{sec:RT}). For zone $i$ and atom $k$, the photoionization rate is
\begin{equation}
P_{\rm i,k} = \frac{\sum_j N_{\rm i,k, j}}{V_{\rm i} \sum_j n_{\rm i,k,j}^*}~,
\end{equation}
where $j$ is the level, $N_{\rm i,k,j}$ is the total number of photoionizations for zone $i$, atom $k$, level $j$, in the Monte Carlo simulation, $V_{\rm i}$ is the zone volume and $n_{\rm i,k,j}^*$ is the previous solution to the number density of level $j$.
\subsubsection{Excitation structure}
\label{sec:excstr}
Given the ion abundances, steady-state is appropriate for the atomic level populations since the collisional and radiative time-scales are short. We compute NLTE solutions for the neutral and singly ionized
elements of H, He, C, N, O, Ne, Na, Mg, Al, Si, S, Ar, Ca, Sc, Ti, V, Cr, Mn, Fe, Co, and Ni, as well as for doubly ionized Fe, using model atoms that include fine-structure levels.
With their large number of
resonance lines in the UV and optical, iron-group elements are especially
important for the radiative transfer, which motivated us to include also those with small abundances. Appendix \ref{app:atomicdata} gives references for the atomic data that we use.
For the line optical depths, we use the Sobolev approximation \citep{Sobolev1957}
\begin{equation}
\tau_{\rm ij} = \frac{A_{\rm ji}\lambda_{\rm ji}^3}{8\pi}\frac{g_{\rm j}}{g_{\rm i}}t\left(n_{\rm i}-\frac{g_{\rm i}}{g_{\rm j}}n_{\rm j}\right)~,
\end{equation}
with the corresponding escape probabilities $\beta_{\rm ij}^{\rm S}$
\begin{equation}
\beta_{\rm ij}^{\rm S} = \frac{1-e^{-\tau_{\rm ij}}}{\tau_{\rm ij}}~,
\end{equation}
where $i$ and $j$ are the lower and upper levels, respectively. The Sobolev approximation is valid as long as the region of line interaction is much smaller than the region over which physical conditions change, which is generally fulfilled in supernovae. The errors caused by line-overlap should be small in the nebular phase \citep{Li1996}. For lines with small $\beta_{\rm ij}^{\rm S}$, we also take into account continuum destruction probabilities \citep{Hummer1985,Chugai1987}, so that the effective escape probability is
\begin{equation}
\beta_{\rm ij}^{\rm eff} = \beta_{\rm ij}^{\rm S} + \beta_{\rm ij}^{\rm C}.
\end{equation}
where the $\beta_{\rm ij}^{\rm C}$ probabilities are calculated as in KF98 a. The effective radiative deexcitation rates used in the NLTE equations are then $A_{\rm ji}\beta_{\rm ij}^{\rm eff}$.
The statistical equilibrium is obtained by solving Eq. (15) in KF98 a. Deexcitation of the He I(2$^3$S) state also occurs through Penning ionizations of H I. Our treatment here differs mainly in that we now compute detailed photoionization rates $P_{\rm i,k,j}$ from the Monte Carlo calculation (Sect. \ref{sec:RT})
\begin{equation}
P_{\rm i,k,j} = \frac{N_{\rm i,k,j}}{V_{\rm i} n_{\rm i,k,j}^*}~.
\end{equation}
As in KF98, we include non-thermal ionizations and photoionizations only to the first level of the ionized element. The only exception is for O II, where non-thermal ionizations to the excited states $^2D_{\rm o}$, $^2P_{\rm o}$ and $^4P$ are also included. Non-thermal ionizations are included only from the ground multiplets in the atoms, whereas photoionizations can occur from excited states where cross sections are available
In this work, we set all photoexcitation/deexcitation rates to zero. Although these can in principle be computed from the radiative transfer calculation and be included in the population equations, this would force a slow single-stepping Monte Carlo transfer. We also aim to study the scattering/fluorescence process in detail (Sect. \ref{sec:sprop}), which is easier if we
can track the Monte Carlo packets through all scattering events until they escape. Finally, this approximation also allows us to compute faster (approximate) NLTE solutions including a fewer number of states. At these late times, photoexcitations do not strongly influence the solutions beyond the direct scattering/fluorescence.
For the treatment of charge transfer, see
Sect. \ref{sec:ct}
\subsubsection{Thermal balance}
For the core zones, the temperature is computed by balancing heating with cooling. Steady-state is appropriate as long as adiabatic cooling is neglegible. This may not be fulfilled at late times for the oxygen, helium, and hydrogen zones (KF98 a). However, we compared the steady-state temperatures in these zones with those obtained from the time-dependent model (KF98 a, b), and they are found to be very similar. We therefore use the steady-state approximation for all core zones. For the envelope zones, we use the temperatures calculated in the time-dependent model. These are given in Table \ref{table:freeze}.
Heating has contributions from non-thermal deposition, photoionizations, collisional deexcitations (occuring in the Monte Carlo calculation), charge transfer reactions, free-free absorptions, and Penning ionizations
\begin{equation}
H = H_{\rm nt} + H_{\rm pi} + H_{\rm cd} + H_{\rm ct} + H_{\rm ff} + H_{\rm pe}~.
\end{equation}
The terms $H_{\rm pi}$, $H_{\rm cd}$ and $H_{\rm ff}$ are computed from the Monte Carlo simulation (Sect. \ref{sec:RT}) by summing all contributions from Eqs. (\ref{eq:piheating}), (\ref{eq:ffheating}) and (\ref{eq:collheat}).
The charge transfer heating is \citep{Kingdon1996}
\begin{equation}
H_{\rm ct} = \sum_{\rm k,k'}n_{\rm k} n_{\rm k'}\xi^{\rm CT}_{\rm k,k'}\Delta E_{\rm k,k'}~,
\end{equation}
where $\xi^{\rm CT}_{\rm k,k'}$ is the reaction rate between atoms $k$ and $k'$, and $\Delta E_{\rm k,k'}$ is the energy defect of the reaction, which is negative for endothermic reactions, which act as coolers.
The heating from Penning ionizations is, based on the rate in \citet{Bell1970}
\begin{equation}
H_{\rm pe} = 7.5\e{-21}n_{\rm He I (2^3 S)}~n_{\rm HI}~
\end{equation}
Cooling is computed from Eqs. (3)-(6) in KF98 a. It has contributions from line cooling, recombination cooling, and free-free emission.
\subsection{Charge transfer}
\label{sec:ct}
We use charge transfer rates taken from \citet{Rutherford1971}, \citet{Rutherford1972}, \citet{Arnaud1985}, \citet{Pequignot1986}, \citet{Kimura1993}, \citet{Swartz1994}, \citet{Kingdon1996}, \citet{Stancil1998} and \citet{Zhao2004}.
For reactions where no measured or calculated rates exist, we use the recipe for estimating rates given by \citet{Pequignot1986}.
The many unknown or uncertain rates introduce an uncertainty for the level of ionization for some of the elements. For some zones, this prohibits a definite determination of which elements that re-emit the non-thermal and radiative ionization energy. We discuss this and the potential effects on the spectrum in Sect. \ref{Res:Tandion} and \ref{sec:eoct}.
When the final states are unknown, we assume these to be the ground states of the elements. An exception to this is the important O II + C I $\rightarrow$ O I + C II reaction, where we assume that the oxygen atom is left in the excited 2p($^1$D) state, which is the one closest to resonance with an energy defect of 0.4 eV (Sect. \ref{sec:onemg}). We also assume that the reaction Ca I + O II $\rightarrow$ Ca II + O I occurs to the excited 5p($^2$P) state in Ca II \citep{Rutherford1972}.
\subsection{Radiative transfer}
\label{sec:RT}
Radiative transfer is treated with a Monte Carlo technique. The application of this method for dealing with differentially expanding astrophysical envelopes was described by \citet{Abbott1985}, where it was applied to the case of stellar winds. \citet{Lucy1987} and \citet{Mazzali1993} describe the method for computing photospheric-phase supernova spectra. The main difficulty working with the equation of radiative transfer is the complexity introduced by the line overlaps that are caused by the differential velocity field. This gives a non-local coupling for which a Monte Carlo treatment is better suited. In addition, Monte Carlo radiative transfer is easily generalized to 2D and 3D, and is well suited for parallelization.
The early Monte Carlo codes mentioned above allowed for electron and resonance line scattering, where deexcitations occured in the same transition as the absorption. \citet{Lucy1999} and \citet{Mazzali2000} improved on this treatment by including fluorescence in all transitions directly connected to the (initial) upper level. \citet{Kasen2006} used the same treatment for the fluorescence in computing photospheric-phase, time-dependent spectra for Type Ia supernovae.
To allow for a more exact treatment of the fluorescence, but still retain the properties of indivisibility and indestructibility for the photon packets, a Monte Carlo formalism based on the concept of macro-atoms was developed in a series of papers by \citet{Lucy2002,Lucy2003,Lucy2005}. Applications and further developments of this method are described in e.g. \citet{Sim2007a}, \citet{Sim2007b} and \citet{Kromer2009}.
Here, we implemented a Monte Carlo algorithm with a similarly high level of physical realism by computing the complete deexcitation cascade following each photon packet absorption. A new photon packet is created for each radiative deexcitation in the cascade, with collisional deexcitations being allowed for as well. The (potential) creation of several packets following the absorption of one packet is handled by a recursive technique. We find that this works well and have therefore not enforced the property of indivisibility otherwise commonly used to simplify the bookkeeping \citep[e.g.][]{Lucy2002}. Also, the property of indestructibility devised in those papers has its main advantage as a Lambda accelerator in the photospheric phase, and is not enforced here.
\subsubsection{Packet emission}
Below, quantities are denoted by primed symbols in the comoving frame, and by unprimed symbols in the rest frame (observer frame). We ignore terms of order $(V/c)$ or higher in the transfer, except for the critical Doppler shifts. The letter $z$ denotes a random number between 0 and 1, newly sampled for each computation.
Total emissivities $j_{\lambda'}'(i,j)$ from lines and continua (two-photon, free-bound and free-free) are obtained for each zone $i$ and wavelength bin$j$ from solutions to the ionization balance, temperature and NLTE level populations (Sect. \ref{sec:sate}). Photon packets are then created with energy
\begin{equation}
E_{\rm i,j}' = \frac{4\pi j'_{\rm \lambda'}(i,j) V \Delta \lambda_{\rm j}'}{N_{\rm i,j}}~,
\end{equation}
where $V$ is the volume of the zone, $\Delta \lambda'_{\rm j}$ is the width of the wavelength bin, and $N_{\rm i,j}$ is the number of packets to split the emission into. The wavelength $\lambda'$ is chosen as the maximum wavelength of bin $j$. This choice ensures that no line self-absorptions, which are already included in the Sobolev escape formalism, occur in the transfer. We currently use a uniform weighting for the number of packets emitted per zone and wavelength bin, $N_{\rm i,j} = N_0$. From inspection of the emerging spectrum (which we also smooth, see Sect. \ref{sec:emspec}), we find $N_0\sim$100 to be sufficient for obtaining neglegible Monte Carlo noise, for binning width $\Delta \lambda'/\lambda'=10^{-3}$.
The starting radius of the packet is sampled from
\begin{equation}
r = \left[r_{\rm i,in}^3 + z\left(r_{\rm i,out}^3-r_{\rm i,in}^3\right)\right]^{1/3}~,
\label{eq:rstart}
\end{equation}
which corresponds to uniform sampling in a shell bounded by $r_{\rm i,in}=V_{\rm i,in}t$ and $r_{\rm i,out}=V_{\rm i,out}t$. The shell velocities are listed in Table \ref{table:zonemasses}.
Isotropic emission corresponds to sampling a direction cosine $\mu$ from
\begin{equation}
\mu = -1 + 2z~.
\label{eq:dircos}
\end{equation}
If the zone is a core zone, the internal radius in the (spherical) clump is also determined fro
\begin{equation}
r_{\rm int} = z^{1/3} R_{\rm i,clump}~,
\label{eq:rclump}
\end{equation}
where $R_{\rm i,clump}$ is the radius of the clumps of zone type $i$. An internal direction cosine $\mu_{\rm int}$ is also randomly drawn according to Eq. (\ref{eq:dircos})
\subsubsection{Packet propagation}
For each packet to be emitted, a random number is drawn to determine at which optical depth the packet should be absorbed (if at all)
\begin{equation}
\tau_{\rm abs} = -\ln(1-z)~.
\label{eq:lifetau}
\end{equation}
Next, the distance to the next critical point is computed.
The comoving wavelength of the photon packet is continuously red-shifting as the packet
travels through the differentially expanding ejecta. From the NLTE solutions, we have a wavelength ordered list of all lines with optical depth $\tau_{\rm ij} > 10^{-3}$. The velocity distance to the next line is computed as
\begin{equation}
\Delta V_{\rm line} = c\left(\lambda_{\rm line}^0/\lambda'-1\right)~
\label{eq:linedist}
\end{equation}
where $\lambda_{\rm line}^0$ is the (atom rest frame) wavelength of the the line closest to $\lambda'$ in that list.
The distance to the next ionization edge is computed as
\begin{equation}
\Delta V_{\rm ion.edge} = c\left(\lambda_{\rm ion.edge}^0/\lambda'-1\right)~
\end{equation}
where $\lambda_{\rm ion.edge}^0$ is the (atom rest frame) ionization edge closest in wavelength to $\lambda'$.
Also, the distance to the next shell $\Delta V_{\rm shell}$ is computed,
as is the distance to the clump edge $\Delta V_{\rm clump}$, if the zone is a core zone.
The velocity distance to the next critical point is then chosen as the smallest of these four distances
\begin{equation}
\Delta V = \mbox{min} \left\{\Delta V_{\rm line},~\Delta V_{\rm ion.edge},~\Delta V_{\rm shell},~\Delta V_{\rm clump}\right\}~.
\end{equation}
The continuum optical depth for this distance is computed as
\begin{equation}
\tau_{\rm cont} = \Delta V t \left(\alpha_{\rm i,dust} + \sum_{\rm k,j} \alpha_{\rm pi}^{\rm i,k,j}(\lambda') + \alpha_{\rm i,ff}(\lambda')\right)~,
\end{equation}
where $\alpha_{\rm i,dust}$ is the absorption coefficient for the dust
\begin{eqnarray}
\alpha_{\rm i,dust} &=& \frac{\tau_{\rm d}}{V_{\rm core}t}~~~~~~~V_{\rm out}^i \le V_{\rm core} \nonumber \\
&=& 0~~~~~~~~~~~~~~~\mbox{otherwise},
\end{eqnarray}
$\alpha_{\rm pi}^{\rm i,k,j}(\lambda')$ is the photoionization absorption coefficient,
and $\alpha_{\rm i,ff}(\lambda')$ is the free-free absorption coefficient.
Electron scattering is neglegible at the epoch studied here, but can easily be included as well.
The accumulated optical depth $\tau_{\rm tot}$ is increased by $\tau_{\rm cont}$. If this causes $\tau_{\rm tot}$ to transcend $\tau_{\rm abs}$, the packet is continuum-absorbed, and propagation terminates. The total number of photoionizations for zone $i$, atom $k$, level $j$, is then increased by
\begin{equation}
\Delta N_{\rm i,k,j} = \frac{\Delta V t\alpha_{\rm pi}^{\rm i,k,j}(\lambda')}{\tau_{\rm cont}} N_{\rm packet}~,
\label{eq:deltaNikj}
\end{equation}
where $N_{\rm packet}$ is the number of photons in the packet
\begin{equation}
N_{\rm packet} = \frac{E'}{hc/\lambda'}~.
\end{equation}
The photoelectric heating rate contribution is
\begin{equation}
\Delta H_{\rm pi}^i = \frac{1}{V_{\rm i}}\sum_{\rm k,j} \Delta N_{\rm i,k,j}\left(\frac{hc}{\lambda'}-\chi_{\rm k,j}\right)~,
\label{eq:piheating}
\end{equation}
where $\chi_{\rm k,j}$ is the ionization potential for atom $k,$ level $j$. The free-free heating rate contribution is
\begin{equation}
\Delta H_{\rm ff}^i = \frac{1}{V_{\rm i}} \frac{\Delta V t\alpha_{\rm i, ff}(\lambda')}{\tau_{\rm cont}}E'~.
\label{eq:ffheating}
\end{equation}
If continuum absorption does not happen, the position, flight angle, wavelength and energy of the packet are now updated. The new position is
\begin{equation}
r_{\rm new} = \left[r^2 + (\Delta Vt)^2 +2 r \Delta V t \mu\right]^{1/2}~.
\end{equation}
The new flight angle is
\begin{equation}
\sin{\theta}_{\rm new} = \sin{\theta}\frac{r}{r_{\rm new}}~.
\end{equation}
The new comoving wavelength is
\begin{equation}
\lambda'_{\rm new} = \lambda' \left(1 + \frac{\Delta V}{c}\right)~,
\label{eq:redshift}
\end{equation}
and the new comoving energy i
\begin{equation}
E'_{\rm new} = E'\left(1+\frac{\Delta V}{c}\right)^{-1}~.
\end{equation}
If the new point $r_{\rm new}$ corresponds to the passage of an ionization edge or a shell edge, new velocity distances are computed, starting from Eq. (\ref{eq:linedist}), and the procedure is repeated. If $r_{\rm new}$ corresponds to entering a new core clump, its type is chosen from surface impact probabilities given by
\begin{equation}
p_{\rm i} = \frac{R_{\rm i,clump}^2}{\sum_j R_{\rm j,clump}^2}~,
\end{equation}
which is valid as long as all clumps have the same number of fragments ($N_{\rm cl}$). An impact parameter is also drawn as
\begin{equation}
\mu_{\rm imp} = z~.
\end{equation}
Propagation then continues into the new clump, starting from Eq. (\ref{eq:linedist}).
\subsubsection{Line interaction}
If $r_{\rm new}$ corresponds to reaching a line with $\tau_{\rm ij} \ge 10^{-3}$, $\tau_{\rm tot}$ is increased by $\tau_{\rm line}$. If $\tau_{\rm tot}$ then transcends $\tau_{\rm abs}$, the packet is absorbed in the line.
All downward radiative and collisional transition rates are then computed. If $u$ is the upper level of the line, the radiative deexcitation rates are ($i<u$)
\begin{equation}
R_{\rm u,i} = A_{\rm u,i}\beta_{\rm u,i}^{\rm eff}~.
\label{eq:raddeexc}
\end{equation}
The collisional deexcitation rates are \citep{OB}
\begin{equation}
C_{\rm u,i} = \frac{8.629\e{-6}}{T^{1/2}}\frac{\Upsilon_{\rm u,i}(T)}{g_{\rm u}} n_{\rm e}~,
\label{eq:colldeexc}
\end{equation}
where $\Upsilon_{\rm u,i}(T)$ is the effective collision strength.
The rates are normalized to obtain the relative transition probabilities, and one of the deexcitation channels is selected by a random draw.
If this is a radiative transition (with a wavelength shorter than a pre-selected cut-off value, chosen as 2.1 $\mu$m here) a new photon packet is emitted. The energy of this packet is
\begin{equation}
E_{\rm out}'=E_{\rm new}'\left(\Delta E_{\rm ui}/E_{\rm u}\right)~
\label{eq:energypacket}
\end{equation}
where $\Delta E_{\rm ui}$ is the energy difference between levels $u$ and $i$ and $E_{\rm u}$ is the energy of level $u$. A direction is randomly drawn from Eq. (\ref{eq:dircos}
, and the packet is followed, starting from Eq. (\ref{eq:lifetau}), until it (and its potential offspring packets) are absorbed or escape the ejecta.
If the deexcitation is instead a collisional one, a contribution to the volumetric heating rate
\begin{equation}
\Delta H_{\rm cd}^{\rm i} = \frac{1}{V_{\rm i}}E_{\rm out}'~,
\label{eq:collheat}
\end{equation}
is recorded. We ignore the possibility of collisional excitations in the cascade. This is a good approximation at late times when the low temperatures make upward collisions much rarer than downward ones.
New deexcitation probabilities from level $i$ are now computed from Eqs. (\ref{eq:raddeexc}) and (\ref{eq:colldeexc}), and a channel is again randomly selected. The process repeats until the atom has deexcited to the ground state. The energy in each deexcitation from level $i$ to level $j$ is given by Eq. (\ref{eq:energypacket}) with $\Delta E_{\rm ui}$ replaced by $\Delta E_{\rm ij}$.
This treatment means that
\begin{equation}
\sum E_{\rm out}' = E_{\rm new}'~.
\end{equation}
While this becomes only an approximate treatment for the energy distribution of the fluorescence when the absorption occurs from an excited state, this is rare at late times when most atoms are in their ground states. This treatment will likely be modified for calculations at earlier epochs when this may not be the case. Note that the outgoing energy (assuming only radiative deexcitations) is generally not conserved in the rest frame, although it is in the comoving frame. This corresponds to adiabatic losses of the radiation field.
\subsubsection{Emerging spectrum}
\label{sec:emspec}
When it reaches the edge of the nebula (with expansion velocity $V_{\rm e}$), the observed wavelength and energy of the escaping packet are obtained as
\begin{equation}
\lambda_{\rm obs} = \lambda'\left(1-\frac{V_{\rm e}}{c}\mu_{\rm e}\right)~,
\end{equation}
\begin{equation}
E_{\rm obs} = E'\left(1-\frac{V_{\rm e}}{c}\mu_{\rm e}\right)^{-1}~,
\end{equation}
where $\mu_{\rm e}$ is the direction cosine of the packet upon crossing the outer edge.
The escaping packets are binned according to $\lambda_{\rm obs}$, using a wavelength bin size of $\Delta \lambda/\lambda=0.001$.
The final spectrum is obtained by smoothing this distribution with a box-car filter of width $\Delta \lambda/\lambda=0.002$.
An alternative method for computing the emerging spectrum would be to solve for full-size atoms, including photoexcitation rates in the NLTE level population solutions, and compute the formal integral. However, computing large NLTE atoms is time-consuming and also numerically challenging in terms of convergence and accuracy. The line transfer mainly depends on the populations of the ground states and other low-lying states, which are more reliably computed than high-lying states \citep{Lucy1999}. We therefore choose to compute smaller NLTE solutions without photoexcitation/deexcitation rates and treat the scattering/fluorescence process explicitly in the Monte Carlo transfer, which also allows us to investigate its properties from the emerging packet histories.
Because our model assumes steady state, we checked that the typical photon flight times $t_{\rm p}$ are short compared to the dynamic time scale. Crossing the ejecta takes a time of about $\sim V_{\rm e}t/c$ for the photons, which means that $t_{\rm p}/t \sim V_{\rm e}/c \ll 1$. However, UV photons may experience multiple line scatterings before emerging and their flight times could therefore become long. Here, the presence of dust in the ejecta limits the total flight times to a maximum of a few crossing times, and steady state is therefore justifiable from the radiative transfer perspective as well.
\subsection{Dust}
\label{sec:dust1}
There is strong evidence from blue-shifting line peaks, a developing IR excess, and
a simultaneous decrease in the optical flux that dust formed in the
inner parts of the ejecta, starting around day 530 \citep{Lucy1989,Wooden1993}, or possibly as early as day 350 \citep{Meikle1993}. By day 700 most of the dust formation was probably over \citep{Lucy1989}. The apparently weak wavelength dependence of the dust extinction and the absence of
any dust features in the IR emission suggest that the dust was at least partly formed in optically thick clumps \citep{Lucy1991part2}.
The last far-IR observations are from
day 1731 \citep{Bouchet1996}, when the dust temperature was estimated to be $\sim$155 K.
\citet{Fassia2002} find that essentially all the NIR ejecta lines have blue-shifted peaks even at day 2112, which indicates that the dust was still optically thick at that time. We determined the median value of their measured line-shifts to 500 \hbox{km\,s$^{-1}$}. This agrees well with \citet{Wang1996}, who find blue-shifts of $\sim$400 \hbox{km\,s$^{-1}$}~for several optical lines at 1862 and 2210 days.
Between 2000 and 6000 days there were no instruments capable of detecting the FIR emission by the ejecta.
\citet{Bouchet2004} identify a Gemini $N$ filter detection at 6094 days with ejecta dust.
Observations at 6526 days were dominated by the heated
dust emission from the ejecta -- ring collision, and no ejecta dust was detected
\citep{Bouchet2006}. But the HST imaging shows a clear 'hole' in
the central parts of the ejecta even at these late times \citep{Larsson2011}, suggesting that the dust is still optically thick. This may, however, also be a result of external X-ray heating from the ring collisions \citep{Larsson2011}.
We examined the observed line profiles of Mg~I] \wl4571 and H$\alpha$
in the eight-year spectrum (see Sect. \ref{sec:TOS} for a description of the observed spectrum). Both show distinct blue-shifts of their peaks of $\sim$700 and $\sim$600 \hbox{km\,s$^{-1}$}, respectively,
indicating that the dust is still providing extinction at this epoch. This was also noted from the H$\alpha$ profile by C97.
The constancy of the line shifts compared to earlier epochs suggests that the dust resides in co-expanding clumps with optical depths much higher than unity also at this late epoch.
Assuming emission and
absorption in a homogeneous sphere of velocity $V_{\rm d}$, the dust radial optical depth $\tau_{\rm d}$
can be found from the blue-shift of the peak $\Delta V$ \citep{Lucy1989}
\begin{equation}
\frac{\Delta V}{V_{\rm d}} = -1 + \frac{\ln\left({1+\tau_{\rm d}}\right)}{\tau_{\rm d}}~.
\label{formula:taud}
\end{equation}
Using this equation
for the Mg~I] \wl4571 and H$\alpha$ line shifts
with $V_{\rm d}=2000$ \hbox{km\,s$^{-1}$} gives $\tau_{\rm d} = 1.0-1.2$, similar to its value at earlier epochs \citep{Lucy1989,Wooden1993}. Using $\Delta V = 400 ~\hbox{km\,s$^{-1}$}$, as found from the day 1862 and day 2210 spectra by \citet{Wang1996}, gives a value of $\tau_{\rm d}=0.6$, and using $\Delta V = 500 ~\hbox{km\,s$^{-1}$}$, as determined in \citet{Fassia2002}, gives $\tau_{\rm d}=0.8$.
Because it is unknown exactly in which zone(s) the dust resides, we model the dust as a constant, gray opacity within the core. From the estimates above, $\tau_{\rm d}$ is likely to be in the $0.6-1.2$ range. For our standard model we choose $\tau_{\rm d}=1$.
We also note that dust in the ejecta can have strong effects on the cooling (KF98 a, b), with consequences for the spectrum depending on where the dust formed.
\section{The observed spectrum}
\label{sec:TOS}
For the UV/optical range, we use the post-COSTAR HST Faint Object Spectrograph observations obtained on January 7 1995 (2875 days or 7.87 years after explosion), which covers the wavelength range from 1650 to 9200 \AA. Details on the observations and the data reduction can be found in C97. The flux calibration is stated to be good to $\sim$5\%. The aperture excludes the inner circumstellar ring, but the northern outer loop passes through the same line of sight as the central debris, and therefore contaminates
the spectrum with some narrow lines. This is not a major problem because these lines can be easily identified. The strongest ones are C III] \wl1909, [O~II] $\lambda$ 3727, [O III] $\lambda \lambda$ 4959, 5007, [N~II] \wll6548, 6584, and H$\alpha$ \wl6563 \citep{Panagia1996}. We did not remove these narrow lines from the observed spectra displayed in the paper.
For the near-infrared, we use the AAT spectrum from \citet{Fassia2002} from day
2952.
The IR spectrum contains contributions from both the ejecta and the inner equatorial ring. Because of the limited spectral resolution of $\lambda/\Delta\lambda=120$, narrow lines from the ring can appear almost as broad as the ejecta lines. In addition,
the removal of both continuum and line emission from star 3 and possibly star 2 adds
a significant uncertainty. The accuracy of the absolute flux levels is stated as $\pm40\%$. The NIR spectrum
is therefore useful for line identifications, but only for approximate quantitative comparisons.
To compare the model to the observations, we adopt a distance of 51.4 kpc \citep{Panagia2005}, the extinction curve (for $\lambda<8000~\AA~$) derived from the neighboring star 2 in \citet{Scuderi1996}, which has $E_{\rm B-V}=0.19$. For $\lambda>8000~\AA$ we use the extinction curve in \citet{Cardelli1989}.
A different analysis by \citet{Fitzpatrick1990} find $E_{\rm B-V}=0.16$, and we take $\pm$0.03 to be a reasonable estimate for the extinction uncertainty.
This dereddening gives a total luminosity in the 1600-8500 \AA~range of $3.4\e{35}$~\hbox{erg\,s$^{-1}$}. The value found by
C97 is $3.5\e{35}$~\hbox{erg\,s$^{-1}$}, using a distance of 50 kpc and $E_{\rm B-V}=0.2$. We also adjust the model spectrum for interstellar UV absorption lines based on \citet{Blades1988}.
The distance uncertainty to SN 1987A is stated as 1.2 kpc in \citet{Panagia2005}, which corresponds to a flux uncertainty of $\pm$4.8\%.
Dereddening the spectrum with $E_{\rm B-V}=0.16$ or 0.22 instead of 0.19 implies as change of $\sim\pm$ 13\% in the total luminosity in the 2000-8000 \AA~range (which is the range we later use for estimating the \iso{44}Ti mass). Combined, the observed luminosity then has an estimated uncertainty of $\sim (5^2+4.8^2+13^2)^{1/2}=\pm 15\%$.
\section{Results}
\label{sec:results}
\subsection{Energy deposition and degradation}
At eight years, \iso{44}Ti completely dominates the energy input over \iso{56}Co and \iso{57}Co. Using a $^{44}$Ti mass of $1.5\e{-4}$\Mo, which we later show to be the best value for reproducing the observed spectrum, the total energy deposition in the ejecta is $1.9\e{36}$ \hbox{erg\,s$^{-1}$}. Of this, $\sim$85\% is positron energy; the other $\sim$15\% is gamma-ray energy, which represents $\sim$5\% of the $5.2\e{36}$ \hbox{erg\,s$^{-1}$} gamma-rays emitted. The other $\sim$95\% of the gamma-rays escape the ejecta.
The dominance of positrons, combined with the assumption that these are locally absorbed, means that most of the energy ($\sim$80\%) is deposited in the Fe/He clumps. Most of the gamma-rays are absorbed in the H zones, giving them $\sim$8\% of the total energy deposition. The remaining $\sim$12\% are split between the silicon, oxygen, and helium zones. All energy depositions are given in Table \ref{table:Zoneresults}.
\subsection{Temperature, ionization, and emissivities}
\label{Res:Tandion}
The solutions for the temperature and ionization in the various zones are given in Table \ref{table:Zoneresults}. From these steady-state solutions, we need to check the assumption that the recombination time-scale is shorter than
the time-scale for density change $n/\dot{n}=t/3\sim 2.6$ years. The last column in Table \ref{table:zonemasses} lists the recombination times ($1/\left[n_{\rm e} \alpha\right])$ in years for the dominant element in the various zones. We see that a steady-state treatment of all core zones as well as of the helium envelope zone is justified, whereas the hydrogen envelope is experiencing freeze-out as expected.
The temperature is 70-170 K in all core zones (Table \ref{table:Zoneresults}). The temperatures could be even lower if molecular cooling or dust cooling are important. This would not have any strong influence on the UV/optical/NIR spectra, however, because non-thermal processes dominate the output at those energies. Table \ref{table:coolers} lists the dominant coolers in the various zones. For the envelope, we use the temperatures listed in Table \ref{table:freeze}.
\begin{table}[htb]
\caption{Dominant cooling transitions for each zone.}
\centering
\begin{tabular}{l l l l}
\hline\hline
Zone & Cooler & Fraction\\
\hline
Fe/He & [Fe~II] 25.99 $\mu$m & 94\% \\
& [Fe~II] 14.98 $\mu$m* & 5\%\\
& [Fe I] 24.04 $\mu$m & 1.4\%\\
\hline
Si/S & [Si I] 68.47 $\mu$m & 63\%\\
& [Si I] 44.81 $\mu$m & 22\%\\
& [Si I] 129.68 $\mu$m & 15\%\\
\hline
O/Si/S & [O~I] 63.19 $\mu$m & 35\%\\
& [Si~I] 68.47 $\mu$m & 32\%\\
& [O I] 44.06 $\mu$m* & 17\%\\
\hline
O/Ne/Mg & [O I] 44.06 $\mu$m & 33\%\\
& [O I] 63.19 $\mu$m & 29\%\\
& [Si I] 68.47 $\mu$m & 16\%\\
\hline
O/C & [O~I] 63.19 $\mu$m & 41\%\\
& [O~I] 44.06 $\mu$m* & 40\%\\
& [Si I] 129.68 $\mu$m & 7\%\\
\hline
He (core) & [Fe~II] 25.99 $\mu$m & 46\%\\
& [Si I] 68.47 $\mu$m & 23\%\\
& [Si II] 34.81 $\mu$m & 16\%\\
\hline
H (core) & [Si~II] 34.81 $\mu$m & 59\%\\
& [O~I] 63.19 $\mu$m & 19\%\\
& [Fe II] 25.99 $\mu$m & 15\%\\
\hline
\end{tabular}
\tablefoot{Transitions that are strictly forbidden (and therefore give no line emission) are marked with an asterisk.}
\label{table:coolers}
\end{table}
\begin{table*}[htb]
\caption{Energy deposition, temperature, and electron fraction in the various zones in the standard model with M(\iso{44}Ti) = $1.5\e{-4}$\Mo.}
\centering
\begin{tabular}{l l l l l l l l l l l l }
\hline\hline
Zone & Fe/He & Si/S & O/Si/S & O/Ne/Mg & O/C & He & H \\
\hline
Deposition [\hbox{erg\,s$^{-1}$}] & $1.6\e{36}$ & $9.1\e{34}$ & $3.6\e{34}$ & $5.4\e{34}$ & $1.7\e{34}$ & $2.6\e{34}$\tablefootmark{a} & $1.5\e{35}$\tablefootmark{a}\\
Fractional deposition & 80\% & 4.7\% & 1.9\% & 2.8\% & 0.88\% & 1.4\%\tablefootmark{a} & $8.0$\%\tablefootmark{a}\\
Deposition external [\hbox{erg\,s$^{-1}$}] & $1.7\e{34}$ & $7.4\e{33}$ & $1.1\e{33}$ & $1.1\e{34}$ & $2.4\e{33}$ & $7.7\e{33}$\tablefootmark{a} & $1.3\e{34}$\tablefootmark{a}\\
\hline
Heating & 0.59 & 0.31 & 0.23 & 0.19 & 0.22 & 0.20\tablefootmark{b} & 0.14\tablefootmark{b}\\
Excitation & 0.16 & 0.27 & 0.21 & 0.21 & 0.18 & 0.23\tablefootmark{b} & 0.41\tablefootmark{b}\\
Ionization & 0.25 & 0.42 & 0.56 & 0.60 & 0.60 & 0.57\tablefootmark{b} & 0.45\tablefootmark{b}\\
\hline
Temperature [K] & 170 & 140 & 140 & 70 & 70 & 140\tablefootmark{b} & 100\tablefootmark{b}\\
Electron fraction & 0.17 & 0.020 & $2.9\e{-3}$ & $9.0\e{-4}$ & $5.6\e{-4}$ & $2.1\e{-4}$\tablefootmark{b} & $1.5\e{-4}$\tablefootmark{b} \\
\hline
\end{tabular}
\tablefoot{
'Deposition external' is the amount of absorbed radiation originating in other zones.\\
\tablefoottext{a}{All zones.}
\tablefoottext{b}{Core zone.}
}
\label{table:Zoneresults}
\end{table*}
We now discuss the individual zones in more detail to understand the emission lines that are produced. In Sect. \ref{sec:pos} we discuss how the energy deposition in the various zones changes if there is a non-local positron deposition.
\subsubsection{Fe/He zone}
\label{sec:fehe}
The high electron fraction in the Fe/He zone ($\sim$0.17) implies that a large part ($\sim$59\%) of the deposited energy goes into heating, re-emerging mainly in the [Fe~II]~26 $\mu$m line (Table 5). The reason that this line dominates over [Fe~I]~24 $\mu$m is that it has a much larger collision strength, $\sim$6 versus $\sim$0.02 \citep{Pelan1997,Zhang1995}. Because the Fe/He zone absorbs $\sim$80\% of the deposited energy (assuming on-the-spot positron absorption), and $\sim$60\% of this goes into heating, $\sim$50\% of the total deposited energy or $\sim 9\e{35}$~\hbox{erg\,s$^{-1}$} (or a factor two or so less considering the dust) should emerge in the [Fe~II]~26 $\mu$m line at this epoch. We discuss observations of this line in Sect \ref{sec:discussion}.
About 25\% of the energy goes into ionizations, with $\sim$14\% going to He I, $\sim$7\% to Fe I and $\sim$4\% to Fe II. Iron reaches an ionization degree of $\sim$40\%, while helium reaches $\sim$6\%. Fe III is largely neutralized by charge transfer reactions, mainly with Fe I. Nickel, which makes up $\sim$4\% of the zone mass, is almost fully ionized owing to charge transfer with Fe II. He I, Fe I, and Ni I therefore emit strong recombination lines. The recombination line spectrum of Fe~I is rather smooth with significant emissivity in the 2000-6000 \AA~range; this constitutes a major part of the emitted flux in this range (Sect. \ref{sec:sprop}). The He I recombination emission is mainly in the form of the two-photon continuum plus He~I \wl626. This flux is mostly continuum-absorbed by other elements and serves to additionally boost the ionization of iron and other metals (see also KF98 b). In the optical, the strongest recombination line is the 5876 \AA~line. Helium also emits IR lines at 1.083 $\mu$m and 2.058 $\mu$m. The 1.083 $\mu$m line is mainly excited by recombinations, while the 2.058 $\mu$m line arises mainly as a result of non-thermal excitations.
The remaining $\sim $16\% of the deposited energy excites resonance lines from the ground states of Fe~I ($\sim$9\%), Fe II ($\sim$3\%) and He I ($\sim$3\%). For the Fe I transitions, we use the approximate Bethe rates, whereas for Fe II and He I we have more accurate cross sections.
Based on modeling of the Fe II lines during the first 800 days \citep{Li1993iron}, the density of the Fe/He zone may be lower than the one we use here, but not higher. Lowering the density of this zone by a factor of two increases the heating fraction by about five percentage units, at the expense of excitations and ionizations. Assuming that the UV/optical/NIR flux scales in proportion to the excitation plus ionization energy, the UV/optical/NIR emission from the Fe/He zone then falls by $\sim$12\% by such a maneuver, and the total UV/optical/NIR emission by $\sim$5-10\%. Using $V_{\rm core}=2500$ \hbox{km\,s$^{-1}$}~instead of 2000 \hbox{km\,s$^{-1}$} (which lowers all densities by a factor of $\sim$2) lowers the total ejecta emission in the 2000-8000 \AA~range by $\sim$10\%.
\subsubsection{Si/S zone}
\label{sec:sis}
The higher density and the smaller amount of $^{44}$Ti ($\sim$5\% of the total) present in this zone compared to the Fe/He zone implies a lower electron fraction ($\sim$0.02). A smaller part of the deposited energy therefore goes to heating (31\%). Of the rest, 42\% goes into ionizations, mainly of the abundant Si~I and S~I. Still, these elements stay almost fully neutral because of rapid charge transfer with elements such as Fe~I, Mg~I and Ti~I. These recombine by further charge transfer reactions, and it is eventually calcium that is the dominantly ionized species. This means that quite a strong Ca~I recombination spectrum is emitted from this zone. In Sect. \ref{sec:eoct}, we show that this holds true even without charge transfer, because calcium is also strongly photoionized
About 27\% of the energy goes to non-thermal excitations, mainly for transitions in Si I and S I
\subsubsection{O/Si/S zone}
This is the smallest oxygen zone (0.16 M$_\odot$) but contains some $^{44}$Ti ($\sim$2\% of the total). The largest part of the energy (56\%) goes into ionizations, which through charge transfer reactions eventually leave iron and aluminium as the dominantly ionized elements.
The non-thermal excitations occur in similar amounts for O~I, Si~I and Mg~I.
\subsubsection{O/Ne/Mg zone}
\label{sec:onemg}
The O/Ne/Mg-zone is by far the most massive oxygen zone (1.89~M$_\odot$), but contain no $^{44}$Ti. The total energy deposition here is 2.8\% of the total. About half of the energy goes into ionizations of oxygen. A series of charge transfer reactions neutralize the oxygen ions and transfer the ionization to sodium, which becomes the most abundant ion. This zone therefore emits strong Na I recombination lines, including the doublet at 5890, 5896 \AA~as well as at 5683, 5688 \AA, 6154 \AA~and 8183, 8194 \AA.
In our model, the dominant charge transfer reactant with O II is C I, where a reaction to the excited 2p($^1$D) state in O I is close to resonance with an energy defect of only 0.4 eV. With the recipe for estimating rates given in \citet{Pequignot1986}, we estimate the reaction rate to be $1\e{-9}$ cm$^3$s$^{-1}$, which together with the relatively high carbon abundance in this zone (4.3\% by number) makes it the dominant reaction for neutralizing the oxygen ions. Because 2p($^1$D) is the upper state for the [O I] \wll6300, 6364 lines, this has the interesting effect of maximizing the doublet emission efficiency by producing a photon for each O I ionization, which is equivalent to $\sim$7\% of the deposited energy. Only $\sim$ 10\% of radiative recombinations pass through the 2p($^1$D) state at these densities (giving $\sim$0.7\% of the deposited energy), and since other competing charge transfer reactions likely do not occur to 2p($^1$D), the conversion efficiency is even lower if they would dominate.
The collisional cross sections for the [O I] doublet lines are moderate and only $\sim$1.5\% of the deposited energy goes to direct non-thermal excitation of 2p($^1$D). This state also receives some indirect contribution by non-thermal excitations to higher levels, the most important channel being via the 3d($^3$S) state. Collisions to this state contribute $\sim$0.4\% to 2p($^1$D) by cascading. The total energy emitted in [O~I] $\lambda \lambda$ 6300, 6364 is then $\sim$8.9\% (7\%+1.5\%+0.4\%) of the deposited energy if the charge transfer reaction with C I dominates, and only $\sim$2.6\% (0.7\%+1.5\%+0.4\%) otherwise.
Magnesium has an abundance (by number) that is $\sim$25 times lower than that of oxygen. But because the Mg I] \wl4571 transition has a large collisional cross-section, it receives about the same number of non-thermal excitations as the [O I] doublet.
About 3.7\% of the deposited energy goes to ionizing Mg I, but Mg II is neutralized by charge transfer with Ni I, which in turn ionizes Na I.
Sodium is five times less abundant still, but Na I has an even larger cross-section to its first excited state, which also receives $\sim$1.5\% of the energy through non-thermal excitations. This contributes further to the Na I \wll5890, 5896 emission, which is, however, dominated by recombinations.
\subsubsection{O/C zone}
This is the intermediate-mass oxygen zone (0.58 M$_\odot$), which receives 0.9\% of the total energy deposition. The low abundance of sodium in this zone, compared to the O/Ne/Mg zone, makes magnesium the dominantly ionized element. A strong Mg~I] \wl4571 recombination line is therefore emitted from this zone, although more magnesium resides in the O/Ne/Mg zone.
\subsubsection{He zone}
Because of mixing, the helium zone has both a core component and an envelope component (Sect. \ref{sec:modeling}). These components behave similarly.
He II is neutralized by charge transfer, mainly with Si I, and ionizations eventually end up in Na II, Mg II, Si II, and Fe II.
Non-thermal excitations go mainly to the He~I 2p($^1$P) state ($\sim$15\%), resulting in the 2.058 $\mu$m line, with only small contributions from recombinations for this line. As long as the positrons are locally absorbed in the Fe/He zone, all helium emission lines have stronger contributions from the Fe/He zone than from the He zone. As an example, the fraction of the total energy deposited in the ejecta going into non-thermal excitations of 2p($^1$P) is 0.16\% from the helium zones, and 2.3\% from the Fe/He zone, which means a 93\% contribution to the 2.058 $\mu$m lines from the Fe/He zone. Similar relations hold for all other helium lines as well. In Sect. \ref{sec:discussion} we explore this point in more detail. Because of the strong helium emission from the Fe/He zone, the He zone is not expected to leave any particular imprint on the spectrum at this epoch.
\subsubsection{H zone}
As with helium, hydrogen is present both in the envelope and as a mixed-in core component (Sect. \ref{sec:modeling}). The envelope zones are powered mainly by freeze-out recombinations. Here, hydrogen emits a low-temperature case B recombination spectrum. The freeze-out emission from the envelope is generally stronger than the emission from the steady-state core, and it provides the bulk of the hydrogen lines and continua in the model. The core component provides the peaks of the hydrogen lines, but also serves to emit ionizing radiation into the other core zones.
In the core H zone, steady-state still prevails and instantaneous gamma-ray deposition determines the output. Ionization of H I accounts for $\sim$36\% of the deposited energy. Because we are close to case B, most recombinations pass through $n=2$. About 28\% of the deposited energy goes into non-thermal excitations of the H I 2p state. Both in the core and in the (inner) envelope, 2s and 2p are still populated according to their statistical weights, i.e. as 1:3. In the core component, the Ly$\alpha$ optical depth is $\sim 4\e{8}$, which gives a ratio of two-photon to Ly$\alpha$ emissivity of $\sim$5. The two-photon continuum therefore dominates the decay of the $n=2$ state. The two-photon continuum can affect the spectrum in many ways. It can boost the ionization of metals with low ionization treshholds, both in the H zones themselves and in the other zones. It can also provide an important input into optically thick lines and affect the UV/optical/NIR spectrum by fluorescence. In wavelength windows where no or few lines exist, the emission may escape and contribute to the UV continuum. In the envelope, the Ly$\alpha$ optical depth is lower and Ly$\alpha$ dominates over the two-photon continuum.
H$\alpha$ from the core is powered in similar amounts by non-thermal excitations and recombinations ($\sim$4\% of the deposited energy each). About 70\% of recombinations lead to H$\alpha$ emission in a Case B scenario at 100 K \citep{Martin1988}.
\subsection{The spectrum}
\label{sec:sprop}
\begin{figure*}[t]
\centering
\includegraphics[width=1\linewidth]{15937_f3.eps}
\caption{Model spectrum at 2875 days for $\mbox{M(\iso{44}Ti)}=\mbox{$1.5\e{-4}$\Mo}$~and $\tau_{\rm d}=1$, together with the observed spectrum. The 1000-5,000 \AA~range. (a) The observed spectrum (red) versus the model spectrum (black). The observed spectrum was dereddened, corrected for redshift, and smoothed below 2200 \AA~and in the 7500-8500 \AA~range. (b) The contributions by various selected elements to the flux. (c) The part of the emerging flux that comes from scattering/fluorescence (red) and the part that is direct emission (blue). (d) The zones in which the photons had their last interaction; this is the zone of emission if they did not scatter and the zone where they last experienced scattering or fluorescence if they did. (e) The zones from where the photons were emitted before any scattering or fluorescence occurred. The color codings for panels 3-5 are on the right-hand side. Ly$\alpha$ extends to to a peak flux of $1.4\e{-14}$ erg s$^{-1}$ \AA$^{-1}$ cm$^{-2}$. See also Sect. \ref{sec:sprop} in the text.}
\label{fig:s1}
\end{figure*}
\begin{figure*}[htb]
\centering
\includegraphics[width=1\linewidth]{15937_f4.eps}
\caption{Same as Fig. \ref{fig:s1} but for the 5000-9,000 \AA~range.}
\label{fig:s2}
\end{figure*}
\begin{figure*}[htb]
\centering
\includegraphics[width=1\linewidth]{15937_f5.eps}
\caption{Same as Fig. \ref{fig:s1} but for the 0.9-2.1 $\mu$m range.}
\label{fig:s4}
\end{figure*}
The model spectrum, for $\mbox{M(\iso{44}Ti)}=\mbox{$1.5\e{-4}$\Mo}$~and $\tau_{\rm d} = 1$, is shown together with the observed spectrum in Figs. \ref{fig:s1}-\ref{fig:s4}. These figures contain five panels. Panel (a) shows the model spectrum (black) overlaid on the observed spectrum (red). The observed spectrum was dereddened and corrected for redshift (Sect. \ref{sec:TOS}). It was also smoothed below 2200 \AA~and in the 7500-8500 \AA~range. Panel (b) identifies the contribution by individual elements to the spectrum. These are the elements that emitted the photon if it escaped directly, and the element that produced the last scattering/fluorescence if it did not. Panel (c) shows the part of the flux that was processed by scattering and/or fluorescence in red, and the part that comes from direct emission in blue. Panel (d) shows in which zones the photons were last processed. This is the zone of emission for photons that escaped directly, and the zone of the last scattering/fluorescence for those which did not. Finally, panel (e) shows from which zones the photons were originally emitted (which is the same as in panel (d) for the photons that escaped directly).
We see that the model reproduces most of the features in the observed spectrum, which bolsters our confidence in the fundamental assumptions in the model and to the code. The main exceptions occur in the 3000-3500 \AA~range, which we discuss below.
Apart from this range, all major observed lines are present in the model spectrum, and, equally important, the model does not produce any strong lines that are not observed.
\subsubsection{Scattering and fluorescence}
\label{sec:saf}
Although the line blocking effect in the photospheric phase has been extensively modeled before \citep[e.g.][]{Karp1977,Lucy1987,Wagoner1991,Mazzali1993}, an analysis of the actual fluorescence in the nebular phase is rarer.
\citet{Li1996} study the process at 300 and 800 days, using a He I two-photon continuum that scatters in a single-density hydrogen envelope. They conclude that the process transfers a significant fraction of the UV emission into a quasi-continuum
in the optical and NIR up to at least 800 days. The importance of the effect was found to be sensitive to the adopted temperature, which determines the population of the lower levels, and thereby the number of optically thick lines.
That especially resonance lines can be optically thick even at eight years
can be seen with a simple estimate. The Sobolev optical depth
for a line from the ground state can be written as
\begin{equation}
\tau \sim 3\e{3} f_{\rm abs} \left(\frac{\lambda}{5000~\AA}\right)\left(\frac{X}{10^{-5}}\right)\left(\frac{n}{10^6~\mbox{cm}^{-3}}\right)~,
\end{equation}
where $X$ is the number fraction of the element and $n$ is the total number density, which is in the range of $10^4-10^7~\mbox{cm}^{-3}$ in the core at this epoch. Obviously, even elements with low
abundances can still be optically thick in their ground-state resonance lines.
By looking at the panels labeled (b) in Figs. \ref{fig:s1}-\ref{fig:s4}, we see that scattering and fluorescence are indeed important for the spectral formation at eight years. The fraction of the emerging flux that has been processed by line transfer (red) is $\sim$60\% in the UV ($<$4000 \AA), $\sim$30\% in the optical (4000-7500 \AA) and $\sim$30\% in the NIR (7500-21,000 \AA). Running the model with the line transfer switched off resulted in a $\sim$30\% increase in the UV flux, a $\sim$10\% decrease in the optical flux, and a $\sim$30\% decrease in the NIR flux.
Few lines are unaffected by the line transfer. Those with the highest fraction of direct emission in the flux are Mg I] \wl4571, H$\beta$, Fe~I] \wl5012, [Fe~I] \wl5697, the [O I]/Fe I complex at $\sim$6300-6400 \AA, H$\alpha$+Ca I \wl6573, He I 1.083 $\mu$m, [Si I] 1.60, 1.64 $\mu$m, He I 2.058 $\mu$m, and a few others in the NIR part of the spectrum.
By inspecting the bottom-most panels in Figs. \ref{fig:s1}-\ref{fig:s4}, we see that a substantial fraction of the emerging flux, especially in the UV, has its origin in the hydrogen envelope. This is mainly freeze-out energy that is emitted as Ly$\alpha$ and two-photon continuum, and is then processed by scattering and fluorescence. Below $\sim$2500 \AA~essentially the whole spectrum originates in the envelope.
An alternative way of illustrating the fluorescence process is to plot the escape wavelengths versus the emitted wavelengths, which we did in Fig. \ref{fig:lambdalambda} for all photons that have been scattered. The transformation of high-energy photons to low-energy ones can clearly be seen here, and it is clear that most of the fluorescent output originates as UV emission ($\lambda_{\rm in} < 4000$ \AA). The low occupation numbers of excited states makes any inverse fluorescence, i.e. conversion to shorter wavelengths, unimportant at this epoch, and the line transfer produces a systematic reddening of the spectrum. The contribution by many input wavelengths to each output wavelength shows that the fluorescence process is insensitive to the input spectrum, as also found by \citet{Li1996}. Some of the more distinct input lines are Ly$\alpha$ and Fe I/Fe II emission lines at $\sim$2500 \AA, $\sim$2900 \AA~and $\sim$3900 \AA. Ly$\alpha$ powers the emission in Mg II 2795, 2802 (Sect. \ref{sec:lineids})
\subsubsection{Line identifications}
\label{sec:lineids}
We now go through the spectrum, identifying and commenting on the most prominent lines. We use the same line peak wavelengths as in C97.
\begin{figure}[t]
\centering
\includegraphics[width=1\linewidth]{15937_f6.eps}
\caption{The fluorescence process illustrated by plotting emerging versus emitted wavelengths ($\lambda_{\rm out}$ and $\lambda_{\rm in}$, respectively) for all escaping photons that were absorbed by a line at least once. The intensities are normalized in each $\lambda_{\rm out}$-bin ($\Delta \lambda/\lambda = 0.03$), so that the dominant $\lambda_{\rm in}$ for each $\lambda_{\rm out}$ can be seen as the yellow/red regions. Contours are at 0.1, 1, 10, 50 and 90\%. Very little fluorescence occurs for $\lambda_{\rm in}> 7000~\AA$, which is therefore excluded from the plot.}
\label{fig:lambdalambda}
\end{figure}
\paragraph{UV} As panel (e) in Fig. \ref{fig:s1} illustrates, most of the emerging far-UV flux (below $\sim$ 2400 \AA) has its origin as emission from the envelope. Panel (b) shows that most of this flux has been reprocessed by scattering and fluorescence.
Longward of $\sim$2400 \AA~the core begins to make significant contributions. The feature at 2400 \AA~is mainly scattering in the Fe II UV 2 multiplet, with similar contributions from the envelope and the iron zone. The feature is blended with an Fe I emission line on the red side. The feature at 2600 \AA~is also a blend of Fe II scattering (UV 1) and Fe I emission. The strong line at 2513 \AA~appears to be scattering in Si I UV 1, complemented by Fe I.
The feature at $\sim$2790 \AA~is scattering by Mg II \wll2795, 2802 in the envelope. We investigated the formation of this line and found it to be dominantly powered by Ly$\alpha$, which is absorbed in the Mg II 1240 \AA~line, which then produces fluorescence at 2795, 2802 \AA, 2798 \AA, 2928, 2936 \AA, 9218 \AA, and 10,914 \AA.
There is some flux missing on the blue side of the Mg II feature, but we find no Fe II contributions that could improve this, as suggested in C97.
Next to the Mg II feature follows emission/scattering in the Mg I \wl2852 line. This feature has a broad component that comes from scattering in the envelope (the line is optically thick out to $\sim$6000 \hbox{km\,s$^{-1}$}~even though magnesium is $\sim$99\% ionized in the envelope), and a narrow component that is dominated by emission from the O/Ne/Mg zone. On the red side there is some contribution from iron and other elements.
The 3000-3500 \AA~range is not well reproduced by the model, and no definitive identifications can be made. Panels (d) and (e) in Fig. \ref{fig:s1} show that most of the flux here originates in the Fe/He zone. The model produces strong lines at 2986 \AA~(Fe I), 3090 \AA~(Fe I and Al I), 3160 \AA~(Ti II scattering), 3270 \AA~(Fe II), 3310 \AA~(Ni I scattering), 3400 \AA, 3440 \AA~and 3490 \AA~(emission and scattering mainly from iron, cobalt, and nickel). The overproduction of several lines in this region could conceivably be due to the omission of trace elements that could provide additional line absorptions. We surveyed the line lists in the 3000-3500 \AA~range of all omitted elements with $Z<30$, however, and there are no lines from these elements that could be important scatterers at this epoch.
The discrepancy could also be caused by the lack of accurate collision cross-sections for the non-thermal excitations in Fe I (we use the Bethe approximation). Although several of the lines are produced by scattering/fluorescence, the origin is often an Fe I emission line. As we see in Sect. \ref{sec:eoct}, charge transfer can also be responsible by introducing uncertainties in the ionization balance. Finally, it is possible that the dust extinction is wavelength-dependent and should be higher in the UV. This would affect the UV lines coming from the core more than the ones coming from the envelope
We identify the feature at 3650 \AA~as $\sim$50\% Balmer continuum and $\sim$50\% scattering by mainly Fe I and Cr I in the iron core. The 'contamination' of the Balmer continuum makes the attempts to determine of the hydrogen zone temperature in W96 and C97 unsuitable, since they are based on an assumption of pure Balmer emission. The temperatures derived there are also significantly higher than in the ones calculated in our model ($\sim$100 K). Taking the contaminations of both the Balmer continuum and H$\alpha$ into account (see below), the observed ratio between the Balmer continuum and H$\alpha$ is 0.25, which is close to the theoretical ratio of 0.32 obtained by using the Case B recombination rates at 80 K in \citet{Martin1988}.
The strong observed line at 3730 \AA~is reproduced by the model and appears to be scattering and emission from Fe~I. We find the [O II] $\lambda$ 3727 contribution to be negligible, agreeing with C97 that there is no need for this line, a possibility raised in W96.
\paragraph{Optical} The 4000-4500 \AA~range is a blend of Fe~I, Ca~I, H~I, and others. The strongest line in this range is at 4223 \AA, which the model identifies as Ca I \wl4226 emission from the Si/S zone. As we discussed in Sect. \ref{sec:sis}, Ca I is likely to emit most of the ionization energy in this zone, a conclusion which is supported by the presence of the 4226 \AA~line in the observed spectrum. This line is sensitive to density; $A\beta$ has a value of $\sim$600 s$^{-1}$ for the density we use here (a number density ten times lower than in the O/Ne/Mg zone), which dominates the rate of $A_{\rm tot}=140$ s$^{-1}$ for other radiative branching paths. But if the Si/S zone would have the same density as the oxygen zones, $A\beta$ would be only $\sim$10 s$^{-1}$ (calcium is also less ionized at higher density and so the scaling is not proportional), and the other deexcitation paths would then dominate. The presence of this line in the spectrum, at this epoch, therefore requires a number density in the Si/S zone at least $\sim$10 times lower than the one of the oxygen zones.
Because the Si/S zone contains some ($\sim$5\%) of the \iso{56}Ni, according to the explosion model we use, this is consistent with the physical picture that radioactive decays cause expansions in the early phase of the supernova, as was found to have happened for the Fe/He clumps by \citet{Li1993iron}.
The rest of the 4200-4500 \AA~ 'plateau' is mainly emission and scattering from the Fe/He zone, complemented with H$\gamma$ at 4343 \AA. The Fe II contribution, which is suggested to dominate the features at 4223 \AA, 4339 \AA, and 4453 \AA~in C97, is only at the $\sim$10\% level in our model.
Mg~I] \wl4571 is one of the lines with the smallest contribution from scattering/fluorescence. The model underproduces the flux by a factor of $\sim$2-3, possibly as a result of the uncertain ionization balance in the O/Ne/Mg zone (Sect. \ref{sec:onemg}). Although most of the synthesized magnesium resides in this zone, its contribution to the Mg I] \wl4571 line is moderate because Mg II is largely neutralized by charge transfer with nickel and sodium, leaving only non-thermal excitations to power the line. It is possible that we use rates that are too high for some of these charge transfer reactions, and that Mg II instead recombines radiatively, in which case the Mg~I] \wl4571 line would become stronger. Our model gives emission of similar magnitudes from the O/Ne/Mg zone, the O/C zone, and the He and H zones.
The feature at 4850 \AA~is $\sim$2/3 H$\beta$ and $\sim$1/3 Fe~I in the model. Removing the Fe I contamination gives a Balmer decrement of $\sim$5, which agrees well with a Case B recombination scenario at $\sim$100 K.
The 5000-5700 \AA~range is a quasi-continuum dominated by Fe I lines from the iron core, with about 3/4 of the flux being unscattered. The range is well reproduced by the model, apart from an overproduction of the Fe I line at 5007 \AA~(the narrow observed line does not come from the ejecta, but from the northern loop).
The line at 5900 \AA~is emission and scattering in Na~I 5896 \AA. The scattered part is predominantly from emission in Na I 5890 \AA, but also from He I 5876 \AA~emitted in the Fe/He zone. Panel (e) in Fig. \ref{fig:s2} shows that most of the energy originates in the O/Ne/Mg zone, which means that the helium contribution is minor. Na I emits strongly from the O/Ne/Mg zone because of a series of charge transfer reactions and a high photoionization rate that make it the dominant ion in this zone (Sect. \ref{sec:onemg}). Apart from the 5890, 5896 \AA~doublet, the Na I recombination spectrum also includes lines at 5690 \AA, 6154~\AA, 8190 \AA~and 1.14 $\mu$m. There are indeed observed lines at $\sim$5690 and $\sim$6154 \AA, but blending with Fe~I and Ca~I makes identifications uncertain. At 8190 \AA~the spectrum is too noisy to clearly detect a line, and in addition there is also contamination from other lines in the model. There is a clear detection at 1.14 $\mu$m in the day 2112 spectrum in \citet{Fassia2002} (not shown here), but all in all it is difficult to tell whether a significant Na I recombination spectrum is emitted by the ejecta or not. The model overproduces the Na I \wll5890, 5896 doublet, and it is possible that the ionizations instead go to magnesium in the O/Ne/Mg zone, especially because the Mg I] \wl4571 line is underproduced. However, in Sect. \ref{sec:eoct} we show that the Na I doublet is overproduced also if the charge transfer reactions are switched off.
[O I] \wll6300, 6364 is contaminated, which the deviation of the observed ratio from the expected 3:1 value verifies. In the model, most of the contaminating flux is from Fe I. The complex is well reproduced by the model, although the good fit relies on the assumed high rate for the O II + C I $\rightarrow$ O I(2p$^1$D) + C II charge transfer reaction, as discussed in Sect. \ref{sec:onemg}. If this reaction is not the dominant channel for O II recombinations, the [O I] $\lambda \lambda$ 6300, 6364 emission would decrease by a factor of $\sim$4. The uncertainties regarding this reaction combined with the Fe I contamination make the oxygen doublet unsuitable for drawing any strong conclusions at this epoch, including attempts to determine the oxygen mass.
Interestingly, H$\alpha$ is significantly contaminated by Ca~I 6572 \AA~in the model. This line originates, like the Ca I 4226 \AA~line, mainly in the Si/S zone, but also in the O/Si/S zone. Being the first transition from the ground state in Ca I, the line has a high effective recombination rate and is about twice as strong as the H$\alpha$ emission from the core H zone. Note that this component is needed to satisfactorily reproduce the 'H$\alpha$' profile, which would otherwise be too flat-topped. For the high-velocity part of H$\alpha$, the good fit of the model is an important verification that our treatment of the freeze-out in the envelope is accurate.
The [Ca~II] $\lambda \lambda$ 7291,7323 doublet is well reproduced and results from pumping in the H and K lines. Fig. \ref{fig:s2}d-e show that the fluorescence is produced by both synthesized and primordial calcium. The H, K lines are optically thick in all core zones, and in the envelope out to $\sim$7000 \hbox{km\,s$^{-1}$}. Fig. \ref{fig:lambdalambda} shows that most of the photons start as emission close to the H and K lines. The flux in the Ca II $\lambda \lambda$ 8600 IR triplet is similar to the flux in the doublet, as it must be for a H, K pumping scheme.
\paragraph{Infrared} In the infrared, some of the observed lines are dominated by narrow circumstellar components seen at low resolution. We identify the 1.083 $\mu$m line, observed in the day 2112 spectrum \citep{Fassia2002}, with He I emission from the Fe/He zone.
The broad component has an observed flux (Table 2 in \citet{Fassia2002}) similar to the model flux. Pa$\gamma$ $\lambda$ 1.0938 $\mu$m also has a strong narrow component. The model shows the ejecta emission here to be dominated by [Si I] 1.099 $\mu$m. The model flux is 1.5 times the the observed broad component at 2112 days.
Further, we identify 1.14 $\mu$m with Na I, 1.20 $\mu$m with Si I, 1.25 $\mu$m with [Fe II], and 1.28 $\mu$m with Pa$\beta$.
The strong 1.44 $\mu$m line is [Fe I] emission from the core, with a model flux 1.8 times the observed value. Also 1.50 $\mu$m and 1.53 $\mu$m are Fe I emission. 1.60 and 1.64 $\mu$m are dominantly [Si~I] emission from the Si/S zone, and 2.058 $\mu$m is He I from the iron core. The 2.058 $\mu$m line is overproduced in the model. We also refer to \citet{Kjaer2010}, where we discuss the IR spectrum at an age of 19 years, with most line identification being the same as at this epoch.
\paragraph{Summary of the spectrum}
In general, we find good agreement with the line identifications in W96 and C97. The most important difference is that we find many lines to be from Fe~I rather than Fe~II. This is what to expect since Fe~I emits not only a recombination line spectrum, but also receives more non-thermal excitations because Fe I atoms outnumber Fe~II ions by a factor of $\sim$1.5 (see Sect. \ref{sec:fehe}). In our model, $\sim$30\% of the flux between 2000 and 8000 \AA~is from Fe~I, and only $\sim$9\% is from Fe~II (including both emission and scattering). We do not, however, include ionizations to excited states in Fe II, which causes some underestimate of the Fe II emission.
\subsection{The \iso{44}Ti mass}
\label{sec:ti-mass}
The character of the spectrum, i.e. the line ratios and the ratio of lines
to continuum, is not very sensitive to the absolute level of the radioactive powering. The character is
partly determined by the relative amounts of energy that go into
excitations and ionizations, which is insensitive to the total
deposition. It is also influenced by the ionization balance of
the dominant elements in each zone, which is likewise
insensitive to the total deposition; the ionization fraction of
the dominant element approximately scales as the square root of the
deposition. Furthermore, there is a negative feedback present because a higher electron fraction reduces the fraction that goes
into ionizations (and excitations) at the expense of the heating
fraction.
The $^{44}$Ti mass therefore mainly determines the absolute
flux level of the spectrum, rather than its form.
Unfortunately, it is difficult to distinguish the $^{44}$Ti mass from
the uncertain dust properties. We
therefore calculated the integrated flux in the 2000-8000 \AA~range for
different combinations of M($^{44}$Ti) and the dust optical depth $\tau_{\rm d}$ (Sect. \ref{sec:dust1}). We did not
include the NIR because of the uncertainties in the calibration of
these observations (Sect. \ref{sec:TOS}). Fig. \ref{fig:ti44contours} shows the $^{44}$Ti mass that reproduces the observed flux in the 2000-8000 \AA~range as function of $\tau_{\rm d}$, together with the masses that produce the correct flux plus/minus 10\% and 20\%.
Note that the emerging flux is less than proportional to the \iso{44}Ti mass.
The main reason is that a
significant fraction of the flux comes from freeze-out emission rather than from instantaneous
\iso{44}Ti deposition. Another reason is that a higher deposition increases
the electron fraction, which increases the fraction of the flux emerging in the FIR.
A $^{44}$Ti mass of $1.5\e{-4}$\Mo~is our best match for $\tau_{\rm d}=1$, which is
our best estimate for the dust extinction from the line profiles. This result is only valid
for the assumption of local positron deposition. The dust
optical depth has a $\pm$30\% uncertainty (Sect. \ref{sec:dust1}), which according to Fig. \ref{fig:ti44contours} gives a $\pm$20\% uncertainty for the best fitting \iso{44}Ti mass.
In Sect. \ref{sec:fehe} we noted
that the uncertainty in the density of the Fe/He zone adds an uncertainty of $\pm$10\% for the model flux, and in Sect. \ref{sec:TOS} we saw that the observed flux level has a $\pm$15\% uncertainty from errors in distance, reddening, and flux calibration. Together, these give a $(15\%^2+10\%^2)^{1/2}=18\%$ uncertainty for the ratio between modeled and observed flux. According to the contours in Fig. \ref{fig:ti44contours}, this corresponds to a $\pm$30\% uncertainty in the \iso{44}Ti mass. Together with the dust error, the total uncertainty for the \iso{44}Ti-mass
becomes $\sim (20\%^2+30\%^2)^{1/2}=36\%$.
The $^{44}$Ti mass is then constrained to $(1.0-2.0)\cdot 10^{-4}$~M$_\odot$.
\begin{figure}[!h]
\centering
\includegraphics[width=1\linewidth]{./15937_f7_new2.eps}
\caption{Relationship between the $^{44}$Ti-mass and the dust optical depth $\tau_{\rm d}$ required to match the total luminosity in the 2000-8000 \AA~range (solid black line). Also plotted are the 10\% and 20\% contours of $|\left(L_{\rm model}-L_{\rm obs}\right)|/L_{\rm obs}$ (dashed lines).}
\label{fig:ti44contours}
\end{figure}
\subsection{Positron trapping}
\label{sec:pos}
An described in Sect. \ref{sec:edep}, an interesting and important question is whether the positrons are indeed locally absorbed, as assumed so far, or if they leak into the other zones in the core before they are absorbed. We recomputed the spectrum for a model where we assume that the positrons are not absorbed on the spot, but in proportion to the electron content (free and bound) of each zone. This treatment is based on the Bethe stopping formula (Eq. (\ref{eq:bethe})), where we neglect the small differences due to different effective ionization potentials. The number of electrons per zone is proportional to $M\bar{Z}/\bar{A}$, where $M$ is the zone mass and $\bar{Z}$ and $\bar{A}$ are the average nuclear charge and atomic weight, respectively.
For this alternative model, the \iso{44}Ti mass needed to give the correct flux in the 2000-8000 \AA~range for $\tau_{\rm d}=1$ is only $8\e{-5}~$M$_\odot$, which we therefore used. The reason for the lower \iso{44}Ti mass is that a smaller fraction of the positron energy is now emitted by far-IR cooling lines (mainly [Fe II] 26 $\mu$m) by the low-density Fe/He zone, and that the oxygen and hydrogen zones have some strong lines in the 2000-8000 \AA~range. Consequently, less \iso{44}Ti is needed to get the same UV/optical/NIR luminosity.
In this model, the zones obtain total energy depositions (positrons + gamma-rays) given in Table \ref{table:leakage}, where we also give the relative change in the deposition fraction with respect to the on-the-spot model. From this, we expect positron leakage to lead to a severe weakening of emission lines from the Fe/He zone, and a boost by up to a factor of $\sim$10 for emission lines from the O/Ne/Mg, O/C, core He, and core H zones. The $^{44}$Ti content of the silicon zones (Si/S and O/Si/S) however, is such that the deposition here does not change much between the two scenarios.
\begin{table}[h!]
\caption{Fractional energy deposition per core zone in the positron-leakage model (which roughly corresponds to the fraction of the electrons that reside in each zone), and the change in fractional deposition relative to the on-the-spot model. About 5\% of the total deposition occurs outside the core.}
\centering
\begin{tabular}{l | l l}
\hline\hline
Zone & Fractional deposition & Relative change \\
\hline
Fe/He & 1.3\% & 0.017 \\
Si/S & 2.1\% & 0.43 \\
O/Si/S & 2.4\% & 1.2\\
O/Ne/Mg & 28\% & 10\\
O/C & 8.7\% & 10\\
He(core)& 4.9\% & 10\\
H(core) & 48\% & 11\\
\hline
\end{tabular}
\label{table:leakage}
\end{table}
Fig. \ref{fig:positrons} shows the resulting model spectrum compared to the standard on-the-spot model (which is the same as in Figs. \ref{fig:s1}-\ref{fig:s4}). On the whole, the spectra are quite similar, but contain some important differences as well.
The hydrogen emission from the core is boosted by a factor of several, as expected. H$\alpha$ and the Balmer continuum are now overproduced. However, because the fraction of the hydrogen that we mix into the core (25\%) is not tightly constrained, there is some room to improve the H lines by reducing this in-mixing. That would, however, increase the deposition in the other core zones and strengthen the lines from them.
The [O I] $\lambda \lambda$ 6300, 6364 doublet becomes much too strong, but this result is sensitive to the uncertain O II + C I $\rightarrow$ O I (2p$^1$D) + C II charge transfer rate used, as discussed in Sect. \ref{sec:onemg}. If this reaction is switched off, the agreement is actually better in this model than in the on-the-spot model. The Mg~I] $\lambda$ 4571 line is mainly emitted from the oxygen zones, and becomes about as much overproduced as it was underproduced before. The Mg I $\lambda$ 2852 line, however, now becomes much too strong (it is truncated in the figure). This line is powered by non-thermal excitations, and is therefore independent of the uncertain ionization balance in the O/Ne/Mg zone (magnesium is dominantly neutral in all scenarios). This line should therefore be a more reliable indicator than the recombination-driven 4571 \AA~line, and all this suggests that too much energy is now being deposited in the oxygen zones. The Na~I recombination lines at
5890, 5896 \AA, and 6160 \AA~from the O/Ne/Mg zone become overproduced as well. The 5890, 5896 \AA~doublet, which was already too strong without leakage, is now a factor $\sim$4 too strong.
For the rest of the spectrum, the two models produce surprisingly similar spectra. Because fluorescence makes important contributions at almost all wavelengths, turning off the iron emission lines from the Fe/He zone typically lowers the line fluxes by no more than a factor of about two. An example of this situation is the 5000-5500 \AA~range. The iron lines with the smallest contamination by fluorescence in the UV/optical are the lines at 3270 \AA, 3490 \AA, 5012 \AA, and 5159 \AA, and 5697 \AA. All of these lines seem somewhat better reproduced in the leakage model
Because the flux calibration in the NIR is uncertain (Sect. \ref{sec:TOS}), we refrain from comparing the models in this range. All in all, we think the on-the-spot model produces a spectrum that better agrees with the observations, although the uncertainties mentioned above prohibit a definitive ruling out of the leakage scenario. The average $\chi^2$-value (per spectral bin) is about twice as high in the leakage model as in the on-the-spot model.
\begin{figure}[h!]
\centering
\includegraphics[width=1\linewidth]{15937_f8.eps}
\caption{Standard model (assuming on-the-spot positron absorption) (upper, black) compared to a model assuming positron deposition in proportion to the electron content of each core zone (lower, black). The latter uses a 40\% smaller $^{44}$Ti mass (in order to produce the correct total flux in the 2000-8000 \AA~range). The Mg I 2852 \AA~line continues up to $\sim 3.1\e{-15}$ in the bottom panel. The observed spectrum in red, same as in Fig. \ref{fig:s1}.}
\label{fig:positrons}
\end{figure}
\subsection{Fragmentation level}
\label{sec:fragmentation}
The fragmentation level, $N_{\rm cl}$, is a proxy for the hydrodynamical mixing. This
parameter determines the length of the paths that the photons travel in their parent zone before they have the chance to enter the other core zones. A high fragmentation level, implying quick exposure to other zones, will lead to a higher degree of radiative energy exchange between the zones. If one were able to constrain this parameter it would be very useful for evaluating multi-dimensional explosion models.
To check the sensitivity of the spectrum to this parameter, we compared spectra for models using $N_{\rm cl}=10$ and $N_{\rm cl}=10^5$. The results were very similar, and we therefore conclude that the fragmentation parameter has no significant influence on the spectrum at this epoch. One possible explanation for this is that few line optical depths can be expected to be close enough to unity that an abundance change (by switching zones) between a high value in the zones where the element is synthesized to the (low) primordial values where it is not, will make any difference. Another could be that few of the line absorptions occur before the photon has exited its parent zone, even in the low fragmentation limit. Finally, a significant part of the line transfer occurs in the envelope, where there is no mixing in the model.
\subsection{Effects of charge transfer}
\label{sec:eoct}
Finally, the uncertainty in many of the charge transfer rates prompted us to examine the influence of these reactions on the spectrum. It is also interesting to see how important these reactions are for supernova spectra in general. We recomputed our standard model ($\mbox{M(\iso{44}Ti)}=\mbox{$1.5\e{-4}$\Mo}$, $\tau_{\rm d}=1$, and on-the-spot positron deposition) with all charge transfer reactions switched off. The resulting spectrum is compared to the standard one in Fig. \ref{fig:wwoutct}.
The total flux in the 2000-8000 \AA~range changes by only a few percent, so charge transfer should not affect our conclusions regarding the \iso{44}Ti mass. Several lines differ significantly between the two models though. The [O I] $\lambda \lambda$ 6300, 6364 doublet is weakened for reasons discussed in Sect. \ref{sec:onemg}, and the Fe I lines now dominate the doublet. At the same time, the O I \wl7775~recombination line, which is not observed, emerges in the model. This suggests that the oxygen ions are indeed neutralized by charge transfer reactions, as in our standard model.
The important Ca I lines (4226 \AA~and 6572 \AA) do not decrease by much, which shows that charge transfer is not critical for them. Ca I is significantly ionized by the internal radiation field, and so emits strong recombination lines even without charge transfer.
\begin{figure}[h!]
\centering
\includegraphics[width=1\linewidth]{15937_f9.eps}
\caption{The standard model (upper, black) compared to a model with all charge transfer reactions switched off (lower, black). Observed spectrum in red.}
\label{fig:wwoutct}
\end{figure}
The Na I \wl5896 emission is reduced somewhat, but is still too strong. Just as with Ca I, Na I is significantly ionized by the internal radiation field, and consequently strong recombination emission still occurs. An important conclusion is therefore that the emission lines from elements with low ionization thresholds are is not very sensitive to the charge transfer, because low ionization thresholds also correlate with high radiative ionization rates.
For many lines, it is not possible to directly say why they change. The 'picket fence' of absorption lines changes as the ionization balance changes, which opens some wavelength windows and closes others. Consider, for example, the Fe I emission line at $\sim$2990 \AA. This emission line is still there in the model without charge transfer, but now a Ni I absorption line has become optically thick that absorbs it. The UV spectrum is actually somewhat better reproduced in the model without charge transfer, which may indicate that we overestimate some of our adopted rates. Unknown charge transfer rates clearly constitute one of the main obstacles to accurate supernova spectral modeling today, and more calculations and measurements of these would be highly desirable.
\section{Discussion}
\label{sec:discussion}
Although most of the optical and infrared emission escapes freely in the nebular phase of supernovae, the UV emission does not. The freeze-out energy in the envelope, as well as non-thermal excitations and ionizations in the core, produce strong UV emission that is significantly absorbed by atomic lines and emerges by fluorescence at longer wavelengths. By this process, the emerging flux in the optical and NIR is enhanced by $\sim$10\% and $\sim$30\% compared to a purely nebular spectrum at the epoch we have looked at here. While the non-local line transfer can have a large impact on individual lines at late times, the fact that it has a moderate impact on the supernova photometry is important for judging the accuracy of models without this effect, which has previously been difficult.
While the fraction of UV photons that are absorbed decreases with time, the fraction of the optical/NIR spectrum that comes from fluorescence may actually increase \citep{Li1996}. The reason for this is that with time the cooling lines move from the optical/NIR into the FIR, and non-thermal processes do not produce very strong optical/NIR emission. Our finding that fluorescence is important also at eight years illustrates this point well.
We showed that the only FUV flux that escapes comes from far out in the hydrogen envelope. A consequence of this is that the ejecta appear larger in the UV than in the optical \citep{Jakobsen1994, Wang1996}. At longer UV wavelengths both the envelope and the core contribute to the spectrum.
From a mathematical perspective, it is not a very well-conditioned problem to attempt exact modeling of the UV lines, because the output depends sensitively on the exact shape of the 'picket fence' of absorption lines, which in turn depends on many uncertain parameters such as abundances, densities, and charge transfer rates. Considering this complexity, it is probably not realistic or meaningful to attempt any modeling of the UV lines in much more detail than we have done here.
As discussed in Sect. \ref{sec:edep}, the positrons emitted in the $^{44}$Ti decay will enter the neighboring oxygen, helium, and hydrogen zones at late times unless a magnetic field exists to trap them. This would have significant consequences for the FIR cooling lines. The model flux in the [Fe II] 26 $\mu$m line, discussed in Sect. \ref{sec:fehe}, corresponds to a peak flux of $\sim$1.8 Jy at day 2875. ISO observations at day 3425 failed to detect this line, giving a upper limit for the peak flux of 0.64 Jy \citep{Lundqvist2001}. Spitzer observations at day 6190 \citep{Bouchet2006} show a distinct broad feature (FWHM $\sim$ 3800 \hbox{km\,s$^{-1}$}) at 26 $\mu$m, but with a peak flux of only $\sim$0.02 Jy. The slow decay of $^{44}$Ti should make the situation at 6190 days similar to the one at 2875 days, and it is therefore surprising that the observed flux in this line is about two orders of magnitude weaker than expected. Possible explanations are that the positrons are indeed leaking out into the other core zones, or that dust is cooling the Fe/He zone. In the positron leakage model, the flux in the 26 $\mu$m line is $\sim$0.16 Jy at 2875 days (for a \iso{44}Ti mass of $8\e{-5}$ M$_\odot$~and no dust correction). As we found in Sect. \ref{sec:pos}, a positron leakage scenario produces a UV/optical spectrum with several lines that are clearly too strong. Our favored explanation for the weak [Fe II] 26 $\mu$m line is therefore that dust cooling occurs in the Fe/He zone. The possibilities of having dust cooling in the ejecta was discussed in KF98 a and in \citet{Lundqvist2001}.
In the case of magnetic confinement of the positrons, all helium lines are dominated by helium in the Fe/He zone, which is produced by freeze-out in the explosion. Although we do not identify any distinct observed lines with helium in the eight-year spectrum, the 19-year spectrum in \citep{Kjaer2010} contains a strong and distinct 2.058 $\mu$m line, which, as we speculated there, probably shows that $\alpha$-rich freeze-out has indeed happened in the Fe/He zone; without it the local positron deposition could not produce such a strong line.
Together with the \iso{56}Ni mass of $0.069\ \hbox{$~\rm M_{\odot}$}$ \citep{Bouchet1991a} and the \iso{57}Ni mass of $\sim 3.3\e{-3} \hbox{$~\rm M_{\odot}$}$
\citep{Fransson1993,Varani1990,Kurfess1992}, the determination of the \iso{44}Ti mass should put considerable constraints on
supernova explosion models. Our result of $1.5_{-0.5}^{+0.5}\e{-4}\Mo$~ agrees with the result of $(1-2)\e{-4}$~M$_\odot$~found in C97,
although that result was based on an analysis of [Fe II] lines, whose identifications are different in our model on
several instances, and on the [O I] $\lambda \lambda$ 6300, 6364 doublet, whose emission is highly sensitive to uncertain charge transfer reactions, and which is also contaminated by Fe I lines. It is therefore a bit of a coincidence that we get similar results for the \iso{44}Ti mass. \citet{Fransson2002} determine the \iso{44}Ti mass to $(0.5-2)\e{-4}$~M$_\odot$~by comparing the observed B and V band HST photometry between 1800 and 3500 days to the output from a time-dependent model without non-local radiative transfer (as in KF98 a, b). We showed here that this transfer, although it is important for individual lines, only boosts the output in the optical/NIR range by 10-30\% at 2875 days (Sect. \ref{sec:saf}), which implies that the results in FK02 should have a corresponding accuracy. An inspection of Fig. 4 in the FK02 paper shows that the best-fitting \iso{44}Ti mass is in the $(1.5-2)\e{-4}$~M$_\odot$~range, which agrees well with our results here; ignoring non-local radiative transfer should give a somewhat over-estimated \iso{44}Ti mass, but not by much. Our result also agrees with the range of $(0.8-2.3)\e{-4}$~M$_\odot$~found by \citet{Motizuki2004}. Their result is, however, is based on a comparison betwen the deposited energy and the bolometric luminosity at 3600 days, the latter being highly uncertain since most of the flux comes out in the unobservable FIR.
The \iso{44}Ti isotope is a result of the
$\alpha$-rich freeze-out \citep[e.g.][]{Woosley1973}, which is a consequence of
the slow rate for the triple-alpha reaction, which cannot keep the
\iso{4}He abundance at the NSE value as the temperature and density
decreases. The high \iso{4}He abundance in turn results in a higher
abundance of \iso{44}Ti than is the case in NSE. The
\iso{44}Ti/\iso{56}Ni ratio therefore sensitively depends on the
entropy \citep[e.g.][]{Woosley1991}. High temperature and/or low
density increases this ratio.
The explosion model we use (from WH07) has a mass ratio
\iso{44}Ti/\iso{56}Ni = $3.1\e{-4}$, which gives a \iso{44}Ti
mass of only $2.1\e{-5}$ M$_\odot$~(for M(\iso{56}Ni) = 0.069 M$_\odot$), which is clearly ruled out according to our modeling. For a spherically symmetric explosion of a 20 M$_\odot$~star, WW95 find an ejected mass of $(1-2)\e{-5}$~M$_\odot$. On the other hand, \citet{Thielemann1996} find a much larger mass of $1.7\e{-4}$~M$_\odot$, closer to the value we derive here. The differences are partly caused by the different nuclear reaction rates used, but mainly because WW95 initiate the explosion by injecting momentum (piston-driven), while \citet{Thielemann1996} deposit thermal energy \citep{Timmes1996}. The solar abundance of the decay product \iso{44}Ca requires a \iso{44}Ti-production higher than in WW95 by a factor $\sim$ 3, averaged over all progenitor masses, suggesting that the yields in those models are indeed low \citep{Timmes1996}.
\citet{Nagataki1998}, \citet{Nagataki2000} and \citet{Maeda2003} have investigated
the effects of asymmetries in the explosion. These (simplified) models have
considerably higher temperatures and entropies in the polar
direction. Consequently, the \iso{44}Ti production occurs mainly in
this direction, and the total \iso{44}Ti mass can be considerably
higher than in a spherically symmetric model.
Further, the rates of several of the nuclear reactions involved in the
\iso{44}Ti production are uncertain. \cite{Nassar2006} find an
increased rate for the important
\iso{40}Ca$(\alpha,\gamma)$\iso{44}Ti reaction, resulting in a factor of two
higher \iso{44}Ti abundance. The theoretical range of \iso{44}Ti mass
is therefore highly uncertain, as discussed by \cite{The2006} for Cas
A.
In addition to \iso{56}Ni, \iso{57}Ni and \iso{44}Ti, other radioactive
isotopes may be present in the ejecta. In particular, \iso{60}Co and \iso{22}Na may have sizeable
abundances. These isotopes are produced by neutron and proton capture, respectively, during carbon burning, and have uncertain yields that are sensitive to the convective prescription \citep{Timmes1996}.
The \iso{60}Co-isotope has a decay time of 7.605 years, and may have a similar mass to \iso{44}Ti (computed as $2.5\e{-5}$~M$_\odot$~in WW95 and $6.5\e{-5}$~M$_\odot$~ in WH07). The gamma-ray luminosity for these two masses are 56\% and 150\% of the \iso{44}Ti gamma-ray luminosity, respectively, and the electron deposition is 12\% and 30\% of the \iso{44}Ti positron deposition, respectively
~\iso{60}Co may therefore make a non-neglegible contribution to the powering at eight years.
The $\beta$-decays occur as electrons with energies
up to 317 keV, with an average energy of 96.5 keV. This is significantly
lower than the energy of \iso{44}Ti-positrons, and the stopping range is therefore
much shorter. A major difference is that most of the \iso{60}Co electron
energy should be deposited in the oxygen-rich zones. These account for
approximately half of the total column density of the core, and from
Fig. \ref{fig:stopping_power} it can be seen that they can easily
trap most of these electrons. The increased deposition in the oxygen-rich regions means that an appreciable \iso{60}Co mass may mimic the
effects of the positron leakage in the \iso{44}Ti dominated models
(Sect. \ref{sec:pos}). It will in particular increase the O I, Mg I, and Na I line fluxes noticeably, further complicating the question of where the positrons are deposited.
\iso{22}Na has a decay time of 3.75 years, emitting a gamma-ray
at 1.274 MeV and in 90\% of the cases a positron, while 10 \% of the
cases occur by electron capture. The positrons have an average energy
of 215.5 keV and a maximum energy of 545.4 keV.
The \iso{22}Na mass in the explosion model is $4.5\e{-6}$ M$_\odot$, in \citet{Woosley1989} it is $2.0\e{-6}$, in WW95 it is $3.0\e{-7}$~M$_\odot$~and in \citet{Thielemann1996} it is $1.3\e{-7}$~M$_\odot$. This gives gamma-ray luminosities of 0.5\%-17\% of the \iso{44}Ti gamma-ray luminosity, and positron depositions of 0.2\%-5.8\% of the \iso{44}Ti positron deposition. Based on these models it is therefore unlikely that \iso{22}Na is important for the energy budget at the epoch studied here.
It is, unfortunately, difficult to constrain the contribution from \iso{60}Co and \iso{22}Na to the eight-year-spectrum because of the uncertainties in the modeling of the oxygen zones, and the possibility of non-local positron/electron absorption occurring. We assume both of these isotopes to make neglegible contributions to the powering at eight years.
The shorter decay time scales of \iso{60}Co and \iso{22}Na, compared to \iso{44}Ti, means that their influence varies with time. The modeling of the spectrum and light
curve at other epochs should therefore have the potential to distinguish between
the \iso{44}Ti and the \iso{60}Co/\iso{22}Na input. Unfortunately, only a few
spectra exist at epochs later than the one studied here, and they are
increasingly contaminated by the flux from the ring collisions. Modeling of the NIR
spectrum at 19 years has shown that the combination of $\mbox{M(\iso{44}Ti)}=1\e{-4}$~M$_\odot$~
and $\tau_{\rm d}=1$ matches the observed NIR spectrum at this epoch reasonably well \citep{Kjaer2010}.
However, this spectrum may be affected both by scattered light and X-ray input from the circumstellar ring \citep{Larsson2011}.
Small or moderate masses of \iso{60}Co and \iso{22}Na are supported by the modeling of the SN 1987A photometry (FK02), which shows
that the light curves do not deviate significantly from a pure \iso{44}Ti input between 1800 and 3500 days.
\section{Conclusions}
\label{sec:conclusions}
Our main conclusions are:
\begin{itemize}
\item A spectral model based on the determination of the positron and gamma-ray degradation, the ionization, excitation and temperature structure of the ejecta, and the multi-line radiative transfer, reproduces most of the observed lines in the eight-year UV/optical/NIR spectrum of SN 1987A to good accuracy. We showed that many regions in the spectrum are produced by Fe I lines. The contribution by Fe II lines is small.
\item Even many years after explosion, scattering and fluorescence are important processes for the spectral formation in Type-II supernovae. At an age of eight years, $\sim$30\% of the emerging optical/NIR flux and $\sim$60\% of the UV flux in SN 1987A is produced by fluorescence. Non-local line transfer decreases the UV flux by $\sim$30\%, increases the optical flux by $\sim$10\%, and increases the NIR flux by $\sim$30\%, compared to a nebular treatment.
\item Most of the flux in the hydrogen lines and continua, as well as significant parts of the UV spectrum, are at eight years produced by freeze-out emission from the envelope rather than by instantaneous \iso{44}Ti deposition. In the far-UV ($\lesssim 2500~\AA$), essentially the whole spectrum is produced in the envelope.
\item {A scenario of positron leakage from the Fe/He clumps produces too strong emission in several lines from the hydrogen and oxygen zones. We therefore support the conclusion in \citet{Chugai1997} that the positrons remain trapped by a magnetic field.}
\item{ The [Fe II] 26 $\mu$m line in Spitzer observations at day 6190 is much weaker than expected. Possible explanations are dust cooling or that positron leakage is occuring at that epoch.}
\item {Modeling the dust as a gray absorber with radial optical depth $\tau_{\rm d}=1$, which is our best estimate from the line peak blue-shifts, the amount of $^{44}$Ti synthesized in the explosion that best reproduces the total flux between 2000 and 8000 \AA~is $1.5\e{-4}$\Mo. We estimate the uncertainty in this result to $\pm 0.5\e{-4}$~M$_\odot$.}
\item One of our line identifications is the Ca I \wl4226 line. If this identification is correct, the density in the Si/S zone must be at least $\sim$10 times lower than in the oxygen zones for the line to be as strong as observed at this epoch. Because the Si/S zone contains some of the $^{56}$Ni, this adds further evidence to the picture that the early radioactive decays cause expansion in the ejecta.
\end{itemize}
\begin{acknowledgements}
We thank Alexander Heger for providing us with the explosion model, Peter Meikle for providing the NIR data, Cathy Ramsbottom for providing Fe II collisional cross sections, Sultana Nahar for consultations on radiative oxygen recombination, and Claes-Ingvar Bjornsson for useful discussions. We especially thank Mattias Ergon for comments on the manuscript and many hours of useful discussions. This work is supported by the Swedish Research Council and the Swedish National Space Board.
\end{acknowledgements}
|
\section{Introduction}
\label{Sec1}
Recent progress in the fabrication of a near-field scanning thermal microscope (NSThM) enables one to
measure the radiative heat transfer between a cooled sample and a hot probe directly in the near-field regime, i.e.\
for distances in the nanometer range~\cite{KittelEtAl05,WishnathEtAl08}.
Due to thermally excited evanescent waves, in this regime one expects an energy transfer several orders of magnitude greater
than the black body value~\cite{PolderVanHove71}. For estimating the heat current in such a device several theoretical models
are available, which describe the probe as a dielectric sphere, and the sample as a semi-infinite dielectric body with a flat
surface~\cite{Dorofeyev98,JPendry99,MuletEtAl01,A.I.VolokitinB.N.J.Persson01,ChapuisEtAl08,Dorofeyev08,DedkovKyasov07}. These
models can now be tested against the data provided by the NSThM.
In the literature, the near-field radiative heat transfer between a sphere and a
structured surface has not been studied so far, although such a geometry is of considerable practical relevance and theoretical
interest. Therefore, in the present paper we analyze the near-field radiative heat transfer
between a spherical probe and planar samples with surface structures such as depicted in
Fig.~\ref{Fig:Configuration}. In particular, we discuss numerical results obtained
for a planar surface structured by an infinite bar and a square pad, respectively.
In this work, we use the general formulation of the near-field radiative heat transfer between a probe
described as a spherical metallic nanoparticle within the dipole approximation and a second material
as developed in Refs.~\cite{DedkovKyasov07,Dorofeyev08,ChapuisEtAl08}, which is based on
Rytov's fluctuational electrodynamics~\cite{RytovEtAl89}. This formulation allows one to take the material's properties of the
probe into account in terms of its electric and magnetic dipole moments, and the properties of the sample material, which is
assumed to be a semi-infinite body with a given surface structure, through the local density of states~\cite{JoulainEtAl03} above
that medium. Since the dipole moments of the probe or nanoparticle are known it remains to calculate the local
density of states (LDOS) above the sample material. This is done within a perturbative
approximation employing the Ewald-Oseen extinction theorem, as described in detail in Ref.~\cite{G.S.Agarwal1977}.
It should be mentioned that our work is closely related to Ref.~\cite{C.HenkelV.Sandoghdar98}, where the
changes of linewidth and the lineshift for a molecule near a structured surface were calculated, since the electric
LDOS can in principle be read off from the Green's function calculated therein. However, in our work not only
the electric but also the magnetic fields are determined, since it is known~\cite{YuMartynenko05,ChapuisEtAl08} that for
metallic nanoparticles these magnetic fields can cause a heat transfer much greater than that due to the electric fields,
as a result of the induction of Foucault's currents.
\begin{figure}[Hhbt]
\centering
\begin{minipage}[t]{0.9\textwidth}
\epsfig{file=Fig1.eps, width = 0.6\textwidth}
\caption{Sketch of a probe-sample configuration with a square pad on a flat surface.}
\label{Fig:Configuration}
\end{minipage}
\end{figure}
This paper is organized as follows: In Sec.~II we introduce the dipole
model for the radiative heat transfer between a metallic nanoparticle
and an arbitrary second material, and outline the strategy of the
following calculations. In order to determine the electromagnetic LDOS,
we deduce the appropriate dyadic Green's functions in Sec.~III by
means of a perturbative expansion due to Agarwal~\cite{G.S.Agarwal1977}, which we
terminate after the first order. Within this approach, the LDOS itself
is calculated and numerically evaluated for different surface profiles
in Sec.~IV, where we also deduce tentative criteria justifying the
restriction to the lowest-order contributions. In Sec.~V we then
present some numerical results for the near-field heat transfer
between a metallic nanosphere and a structured surface, and predict
signatures that should be observable with a NSThM operated in either
constant-height or constant-distance mode.
\section{Heat transfer between a Nanoparticle and a second medium}
\label{Sec2}
It has recently been shown~\cite{DedkovKyasov07,ChapuisEtAl08,Dorofeyev08} that the near-field
radiative heat transfer between a metallic nanoparticle
with radius $R$ and temperature $T_{\rm P}$ and a dielectric material with
temperature $T_{\rm S}$ can be described within the dipole model for particle-sample distances $a \gg R$. In this model
the electric polarisability $\alpha_{\rm P}$ and the magnetic polarisability $\mu_{\rm P}$ of the particle are given by~\cite{Landau60}
\begin{equation}
\alpha_{\rm P}(\omega,R) = 4 \pi R^3 \frac{\epsilon_{\rm P}(\omega)-1}{\epsilon_{\rm P}(\omega)+2}
\label{Eq:alpha_E}
\end{equation}
and
\begin{equation}
\mu_{\rm P}(\omega,R) = -\frac{R^3}{2}\biggl[1 - 3\frac{d_{\rm s}'^2(\omega)}{R^2} +
3 \frac{d_{\rm s}'(\omega)}{R}
\cot\biggl( \frac{R}{d_{\rm s}'(\omega)} \biggr) \biggr]
\label{Eq:alpha_H}
\end{equation}
with the permittivity of the probe $\epsilon_{\rm P}(\omega)$ and $d_{\rm s}'(\omega) := c (\omega \sqrt{\epsilon_{\rm P}(\omega) - 1})^{-1}$; as usual, $c$ denotes the velocity of light in vacuum. Assuming that the two bodies are in local thermal
equilibrium, the mean energy rate $P$ flowing from the hot to the cold body is given by the relation
\begin{equation}
P = \int_0^\infty\!\!{\rm d} \omega\, \bigl[ \Theta(\omega,T_{\rm P}) - \Theta(\omega,T_{\rm S}) \bigr] 2 \omega \bigl(
\alpha_{\rm P}''(\omega,R) D^{\rm E}(\omega,\mathbf{r}) + \mu_{\rm P}''(\omega,R) D^{\rm H}(\omega,\mathbf{r})\bigr).
\label{Eq:DipolModel}
\end{equation}
Here, the sign of the mean energy rate $P$ is determined by the difference of the Bose-Einstein functions
$\Theta(\omega,T):= \hbar \omega (\exp(\hbar \omega \beta) - 1)^{-1}$ with the
temperature $\beta^{-1} := k_{\rm B} T$, so that the sign is positive for $T_{\rm P} > T_{\rm S}$ and negative otherwise.
The material's properties of the nanoparticle are taken into account in Eq.~(\ref{Eq:DipolModel})
by means of the imaginary part (as indicated by the double prime) of the electric and magnetic
polarisabilities, $\alpha_{\rm P}''(\omega,R)$ and $\mu_{\rm P}''(\omega,R)$, whereas the material's and geometrical
properties of the second medium enter into this expression via the electric and magnetic local density of states
$D^{\rm E}(\omega,\mathbf{r})$ and $D^{\rm H}(\omega,\mathbf{r})$ at the point $\mathbf{r}$ above this medium, where the probe
is located. Therefore, the thermal radiative heat transfer between a nanoparticle and an arbitrary second medium can be calculated if the
electromagnetic local density of states above that medium is known. We point out that in the opposite limit
$a \ll R$ the dipole model discussed here is not valid, whereas the so called ``proximity approximation'' should
prove to be useful then~\cite{VolokitinPersson2007}.
Now the local density of states can be calculated with the help of the relations~\cite{JoulainEtAl05}
\begin{equation}
D^{\rm E}(\omega,\mathbf{r}) = \frac{\omega}{\pi c^2} \epsilon_{\rm S}'' (\omega) \int\!\! {\rm d}^3 r'\, {\rm Tr}\biggl[ \GE{r}{r'} \GE{r}{r'}^\dagger \biggr]
\label{Eq:non_equi_electric_LDOS}
\end{equation}
and
\begin{equation}
D^{\rm H}(\omega,\mathbf{r}) = \frac{\omega}{\pi c^2} \frac{\epsilon_{\rm S}'' (\omega)}{\omega^2 \mu_0^2} \int\!\! {\rm d}^3 r'\, {\rm Tr}\biggl[ \nabla\times\GE{r}{r'}(\nabla\times\GE{r}{r'})^\dagger \biggr],
\label{Eq:non_equi_magnetic_LDOS}
\end{equation}
in the case of local equilibrium inside a heat-radiating local medium surrounded by vacuum. Here
$\mu_0$ is the magnetic permeability of the vacuum, and $\epsilon_{\rm S}(\omega)$ is the relative permittivity of the
material considered. These relations can in principle be evaluated by determining the electric dyadic Green's function
$\GE{r}{r'}$ with the source points $\mathbf{r}'$ inside the medium and the observation
point $\mathbf{r}$ outside the medium, implementing the tensor product with its hermitian conjugate and integrating over
the volume of the medium. Since, here only a local equilibrium inside the medium is assumed, one can use these relations
to investigate for example the heat transfer between bodies kept at different temperatures.
Considering a medium surrounded by vacuum, the non-equilibrium expressions in Eqs.~(\ref{Eq:non_equi_electric_LDOS}) and
(\ref{Eq:non_equi_magnetic_LDOS}) can be decomposed into an evanescent and a propagating part by transforming the
volume integral into a surface integral~\cite{VolokitinPersson2007}.
This evanescent part of the local density of states coincides with the equilibrium expression, since the
evanescent modes are bound to the surface of the medium and are therefore not
relevant for preserving a global equilibrium situation. Thus, the evanescent part of the local density of states
above a material surrounded by vacuum can also be calculated by means of the equilibrium
expressions~\cite{JoulainEtAl05,VolokitinPersson2007}
\begin{equation}
D^{\rm E}(\omega,\mathbf{r}) = \frac{\omega}{\pi c^2} \Im \,{\rm Tr}\, \GER{r}{r}
\label{Eq:equi_electric_LDOS}
\end{equation}
and
\begin{equation}
D^{\rm H}(\omega,\mathbf{r}) = \frac{\omega}{\pi c^2} \Im \, {\rm Tr}\, \GHR{r}{r},
\label{Eq:equi_magnetic_LDOS}
\end{equation}
which state that the electric and magnetic local density of states at the point $\mathbf{r}$ is given by
the imaginary part of the trace of the
renormalised Green's functions $\GER{r}{r}$ and $\GHR{r}{r}$, where the renormalisation procedure is defined as~\cite{Landau92}
\begin{equation}
\mathds{G}^{{\rm E}/{\rm H}}_{\rm R} (\mathbf{r,r}) = \lim_{\mathbf{r} \rightarrow \mathbf{r}'} \biggl[\mathds{G}^{{\rm E}/{\rm H}} (\mathbf{r,r'}) - \mathds{G}^{{\rm E}/{\rm H}}_0 (\mathbf{r,r'}) \biggr]
\label{Eq:renormalisation}
\end{equation}
with $\mathds{G}^{{\rm E}/{\rm H}}_0 (\mathbf{r,r'})$ denoting the Green's function of the vacuum. Since the Green's function with
the observation and source point located above the medium consists
of an incident and a reflected part, with the incident part coinciding with $\mathds{G}^{{\rm E}/{\rm H}}_0 (\mathbf{r,r'})$, the renormalised
dyadic Green's function coincides with the reflected Green's function, so that the index ``R'' can be understood as both ``renormalised'' and ``reflected''.
Obviously, the relations in Eqs.\ (\ref{Eq:equi_electric_LDOS}) and (\ref{Eq:equi_magnetic_LDOS}) are much easier to evaluate than the
expressions in Eqs.\ (\ref{Eq:non_equi_electric_LDOS}) and (\ref{Eq:non_equi_magnetic_LDOS}),
giving reliable results for such distances above the material, at which the evanescent modes dominate the local density of states.
Since we are especially interested in the evanescent regime, we focus on the equilibrium
relations in order to determine the radiative heat transfer between a nanoparticle and a structured surface, keeping in mind that
the results hold in the evanescent regime only. Thus, it is necessary
to calculate the reflected electric and magnetic Green's function with observation points $\mathbf{r}$ and source points
$\mathbf{r}'$ located outside the material of interest. For an
electric source current $\mathbf{j}_{e0}$ and a magnetic source current $\mathbf{j}_{m0}$ located at
the point $\mathbf{r}'$, the reflected fields and reflected dyadic Green's functions are connected by the relations
\begin{equation}
\mathbf{E}_{\rm R}(\omega,\mathbf{r}) = {\rm i} \omega \mu_0 \GER{r}{r'} \mathbf{j}_{e0}
\label{Eq:E_field_greens_function}
\end{equation}
and
\begin{equation}
\mathbf{H}_{\rm R}(\omega,\mathbf{r}) = {\rm i} \omega \mu_0 \GHR{r}{r'} \mathbf{j}_{m0}.
\label{Eq:H_field_greens_function}
\end{equation}
Furthermore, in vacuum the electric and magnetic dyadic Green's functions are related by~\cite{Felsen94}
\begin{equation}
\GHR{r}{r'} = - \frac{1}{k_0^2} \nabla\times\GER{r}{r'}\times\nabla'
\label{Eq:Green_electric_magnetic}
\end{equation}
with $k_0 := \omega/c$.
In the following, we will therefore perturbatively evaluate the reflected electric
field generated by an electric current $\mathbf{j}_{e0}$ located at the source point $\mathbf{r}'$
above a semi-infinite medium with a structured surface,
determine the electric dyadic Green's function by means of Eq.\ (\ref{Eq:E_field_greens_function}), and from this result
deduce the magnetic dyadic Green's function from Eq.\ (\ref{Eq:Green_electric_magnetic}). Finally, we calculate
the local densities of states above the structured surface with Eqs.\ (\ref{Eq:equi_electric_LDOS}) and (\ref{Eq:equi_magnetic_LDOS}).
These densities, in their turn, then allow us to determine the near-field radiative heat transfer between
a metallic nanoparticle and a semi-ininite medium with a structured surface.
\section{Green's function above a structured surface}
\label{Sec3}
In order to calculate the electromagnetic fields above a structured surface, we assume that the
surface profile is given by an expression $h f(x,y)$, where the dimensionless function $f(x,y)$
varies between zero and unity, and $h$ is the characteristic scale of the profile variation, as sketched
in Fig.~\ref{Fig:profil_geometry}. Moreover,
we assume that $h$ be small compared to all other relevant length scales of the problem, so that
we can apply a perturbation expansion. This approach has been worked out in great detail by
Agarwal~\cite{G.S.Agarwal1977}, so that it suffices here to mention only those elements that are indispensable to follow
our line of reasoning.
Within this approach the so-called Ewald-Oseen extinction theorem \cite{PattanayakWolf72,M.Vesperinas06} is employed,
allowing us to restate the boundary conditions of the electromagnetic fields in the given geometry as integral equations.
For the case of a non-magnetic, isotropic, local, and linear material this theorem states that for
observation points $\mathbf{r}$ outside that material, i.e. $\mathbf{r} \notin V$
(see Fig.\ \ref{Fig:profil_geometry}), one has
\begin{equation}
\begin{split}
\mathbf{E}(\mathbf{r}) &= \mathbf{E}_{\rm I}(\mathbf{r}) + \mathbf{E}_{\rm R}(\mathbf{r})\\
&=\mathbf{E}_{\rm I} + \frac{1}{k_0^2} \boldsymbol{\nabla}\times\boldsymbol{\nabla}
\times \int_{\partial V} {\rm d} S' \left[\mathbf{E}_{\rm T}(\mathbf{r}')
\frac{\partial g(\mathbf{r}-\mathbf{r}')}{\partial \mathbf{n}'} -
g(\mathbf{r}-\mathbf{r}') \frac{\partial \mathbf{E}_{\rm T}(\mathbf{r}')}
{\partial \mathbf{n}'}\right].
\end{split}\label{Eq:Ewald1}
\end{equation}
Here, the field $\mathbf{E}(\mathbf{r})$ outside the medium is simply the sum of the incident field
$\mathbf{E}_{\rm I}(\mathbf{r})$ and the reflected field $\mathbf{E}_{\rm R}(\mathbf{r})$. The latter
is described by the surface integral in Eq.~(\ref{Eq:Ewald1}), so that the reflected field
$\mathbf{E}_{\rm R}(\mathbf{r})$ can be calculated by means of the free Green's function
\begin{equation}
g(\mathbf{r}-\mathbf{r}') = \frac{{\rm e}^{{\rm i} k_0 |\mathbf{r}-\mathbf{r}'|}}{4 \pi |\mathbf{r}-\mathbf{r}'|}
\end{equation}
and the transmitted field $\mathbf{E}_{\rm T}(\mathbf{r})$. In addition, $\frac{\partial}{\partial \mathbf{n}'}$
symbolizes the normal derivative, taken in the direction of the unit normal of the surface,
\begin{equation}
\mathbf{n}' = - \frac{\mathbf{e}_{\rm z} + h \boldsymbol{\nabla}'_\parallel f(x',y')}{\sqrt{1 + h^2
|\boldsymbol{\nabla}'_\parallel f(x',y')|^2}},
\end{equation}
with $\boldsymbol{\nabla}'_\parallel := (\partial_{x'},\partial_{y'},0)^t$ and $\mathbf{e}_z$ the unit vector in $z$-direction.
\begin{figure}[Hhbt]
\centering
\begin{minipage}[t]{0.75\textwidth}
\centering
\epsfig{file=Fig2.eps, width = 0.6\textwidth}
\caption{Sketch of an example structure, showing a particular value of the surface profile $h f(x,y)$
at $x = x_0$ and $y = 0$.}
\label{Fig:profil_geometry}
\end{minipage}
\end{figure}
On the other hand, for observation points $\mathbf{r}$ within the medium, i.e.\ $\mathbf{r} \in V$, the Ewald-Oseen
theorem gives
\begin{equation}
0 = \mathbf{E}_{\rm I} + \frac{1}{k_0^2} \boldsymbol{\nabla}\times\boldsymbol{\nabla}
\times \int_{\partial V} {\rm d} S' \left[\mathbf{E}_{\rm T}(\mathbf{r}')
\frac{\partial g(\mathbf{r}-\mathbf{r}')}{\partial \mathbf{n}'} -
g(\mathbf{r}-\mathbf{r}') \frac{\partial \mathbf{E}_{\rm T}(\mathbf{r}')}
{\partial \mathbf{n}'}\right].
\label{Eq:Ewald2}
\end{equation}
In fact, the surface integrals in Eqs.~(\ref{Eq:Ewald1}) and (\ref{Eq:Ewald2}) have the same structure, but
one has to keep in mind that $\mathbf{r}\notin V$ in Eq.~(\ref{Eq:Ewald1}), whereas $\mathbf{r} \in V$
in Eq.~(\ref{Eq:Ewald2}), so that both integrals give different results.
By means of Eq.~(\ref{Eq:Ewald2}) the transmitted field $\mathbf{E}_{\rm T}(\mathbf{r})$ can be computed if
the incident field $\mathbf{E}_{\rm I}(\mathbf{r})$ is given, so that the
sought-after reflected field $\mathbf{E}_{\rm R}(\mathbf{r})$ can then be determined with the
surface integral in Eq.~(\ref{Eq:Ewald1}).
Now the transmitted and reflected fields in Eqs.~(\ref{Eq:Ewald1}) and (\ref{Eq:Ewald2})
are expanded in a power series
\begin{equation}
\mathbf{E}_{\rm T/R}(\mathbf{r}) = \sum_{n=0}^{\infty} \mathbf{E}^{(n)}_{\rm T/R} h^n.
\end{equation}
Furthermore, all quantities in the Ewald-Oseen extinction theorem are expanded in a Taylor
series with respect to the small quantity $h f(x,y)$, giving
\begin{equation}
\mathbf{E}_{\rm T}(\mathbf{r}') = \mathbf{E}_{\rm T}^{(0)}(x',y',z'=0) - h f(x',y')
\frac{\partial\mathbf{E}_{\rm T}^{(0)}}{\partial z'}(x',y',z'=0) + \mathcal{O}(h^2)
\end{equation}
for the transmitted field and
\begin{equation}
g(\mathbf{r}-\mathbf{r}') = g(\mathbf{r}-\mathbf{r}')\Big|_{z'=0} -h f(x',y') \frac{\partial}
{\partial z'} g(\mathbf{r}-\mathbf{r}')\Big|_{z'=0} + \mathcal{O}(h^2)
\end{equation}
for the Green's function. The normal derivative times the surface element ${\rm d} S'$ yields
\begin{equation}
{\rm d} S' \frac{\partial}{\partial \mathbf{n}'} = - dx' dy'
\left[\mathbf{e}_{\rm z} + \boldsymbol{\nabla}'_\parallel h f(x',y') \right]\cdot
\boldsymbol{\nabla}'\;.
\end{equation}
Substituting these expressions into Eqs.~(\ref{Eq:Ewald1}) and (\ref{Eq:Ewald2}), and comparing
coefficients yields constitutive equations for the fields in different orders of $h$.
Next, it is useful to expand the fields in plane waves according to
\begin{equation}
\begin{split}
\mathbf{E}_{\rm T}(\mathbf{r}) &= \int \!\!\! \frac{{\rm d}^2 \kappa}{(2\pi)^2}\, {\rm e}^{{\rm i} (\boldsymbol{\kappa}\cdot\mathbf{x}
+k_z z)} \boldsymbol{\mathcal{E}}_{\rm T}(\boldsymbol{\kappa})\\
\mathbf{E}_{\rm R}(\mathbf{r}) &= \int \!\!\! \frac{{\rm d}^2 \kappa}{(2\pi)^2}\, {\rm e}^{{\rm i} (\boldsymbol{\kappa}\cdot\mathbf{x}
-k_{z0} z)} \boldsymbol{\mathcal{E}}_{\rm R}(\boldsymbol{\kappa})\\
\mathbf{E}_{\rm I}(\mathbf{r}) &= \int \!\!\! \frac{{\rm d}^2 \kappa}{(2\pi)^2}\, {\rm e}^{{\rm i} (\boldsymbol{\kappa}\cdot\mathbf{x}
+k_{z0} z)} \boldsymbol{\mathcal{E}}_{\rm I}(\boldsymbol{\kappa})
\end{split},
\end{equation}
and to utilize the Weyl expansion
\begin{equation}
g(\mathbf{r}-\mathbf{r}') = \int \!\!\! \frac{{\rm d}^2 \kappa}{(2 \pi)^2} \, \frac{{\rm i}}{2 k_{z0}}
{\rm e}^{{\rm i}(\boldsymbol{\kappa}\cdot(\mathbf{x}-\mathbf{x}') + k_{z0} |z-z'|)}
\label{Eq:Weyl}
\end{equation}
for the Green's function, having introduced the notation $k_{z0}^2 = k_0^2 - \boldsymbol{\kappa}^2$,
$k_z^2 = k_0^2 \epsilon_r - \boldsymbol{\kappa}^2$, $\mathbf{x}:= (x,y)^t$,
and $\boldsymbol{\kappa} = (k_x,k_y)^t$. It has to be
emphasized that the use of these expansions down to the surface relies on the Rayleigh
hypothesis, as discussed in Ref.~\cite{M.Vesperinas06}.
With the help of the plane-wave expansions the relations for the Fourier components of the
reflected fields in terms of the incident field are easily calculated; for details we refer again
to the work by Agarwal~\cite{G.S.Agarwal1977}. For the zeroth-order field one obtains the expression
\begin{equation}
\boldsymbol{\mathcal{E}}_{\rm R}^{(0)}(\boldsymbol{\kappa}) = -\left[\frac{k_z-k_{z0}}{k_z+k_{z0}} \mathds{1}
+ \frac{2 k_{z0} (k_{z} - k_{z0})}
{k^2_0(k_{z0}\varepsilon_{r} + k_z)} \left(\boldsymbol{\kappa} - k_{z}k_{z0}\mathbf{e}_z\right)
\right]\boldsymbol{\mathcal{E}}_{\rm I}(\boldsymbol{\kappa}),
\label{Eq:Ref0}
\end{equation}
and for the first-order field
\begin{equation}
\boldsymbol{\mathcal{E}}_{\rm R}^{(1)}(\boldsymbol{\kappa}) = {\rm i} (\varepsilon_{{\rm S}} -1 ) \int\!\frac{{\rm d}^2 \kappa'}
{(2\pi)^2}\, F(\boldsymbol{\kappa} -\boldsymbol{\kappa}')\, \mathds{L}(\boldsymbol{\kappa})\,
\biggl[\frac{2 k_{z0}'}{k_z' + k_{z0}'} \mathds{1}
+ \frac{2 k_{z0}' }{k_0^2 (k_{z0}' \epsilon_{\rm S} + k_z')}\mathbf{K}_0'\otimes\mathbf{K}' \biggr]\boldsymbol{\mathcal{E}}_{I}(\boldsymbol{\kappa}'),
\label{Eq:Ref1}
\end{equation}
with $\mathbf{K}' := (\boldsymbol{\kappa}',k_z')^t$, $\mathbf{K}_0' := (\boldsymbol{\kappa}',k_{z0}')^t$, and the dyadic operator
\begin{equation}
\begin{split}
\mathds{L}(\boldsymbol{\kappa}) &:= \frac{1}{k_{z0}\varepsilon_{\rm S} - k_z}\Big[(\kappa^2 +k_zk_{z0})
(\mathds{1} - \mathbf{e}_z\otimes \mathbf{e}_z) - \boldsymbol{\kappa}\otimes\boldsymbol{\kappa}
+ k_z \mathbf{e}_z \otimes \boldsymbol{\kappa} + \varepsilon_r \kappa^2 \mathbf{e}_z \otimes
\mathbf{e}_z\\
&\qquad\qquad\qquad +\varepsilon_r k_{z0} \boldsymbol{\kappa}\otimes\mathbf{e}_z\Big].
\end{split}
\end{equation}
Here $\otimes$ symbolizes the dyadic product of two vectors, and $\mathds{1}$ the unit dyad. Of course, also higher orders
can be calculated within this approach~\cite{G.S.Agarwal1977}, but the higher-order contributions become increasingly
cumbersome. Thus, this approach is particularly useful if meaningful results can already be obtained to first order.
Hence, we restrict ourselves here to these first-order fields, and try to give approximate criteria justifying this
termination of the series later on.
From Eq.~(\ref{Eq:E_field_greens_function}) it is clear that the Fourier component of the electric dyadic Green's
function for the reflected fields given in Eqs.~(\ref{Eq:Ref0}) and (\ref{Eq:Ref1}) can be read off if
we consider the incident electric fields
$\boldsymbol{\mathcal{E}}_{\rm I}$ generated by a delta-like source current $\mathbf{j}_{e0}$
located at $\mathbf{r}''$ and put the Fourier component of the result into Eqs.~(\ref{Eq:Ref0}) and (\ref{Eq:Ref1}).
According to Eq.~(\ref{Eq:E_field_greens_function}), this field can be stated as
\begin{equation}
\mathbf{E}_{\rm I}(\omega,\mathbf{r}) = {\rm i} \omega \mu_0 \mathds{G}_0^{\rm E} (\mathbf{r},\mathbf{r}'')
\cdot \mathbf{j}_{e0},
\label{Eq:incident_field0}
\end{equation}
where here $\mathds{G}_0^{\rm E}$ is the free dyadic Green's function, which can directly
be obtained from the relation
\begin{equation}
\mathds{G}_0^{\rm E}(\mathbf{r},\mathbf{r}'') = \left(\mathds{1} + \frac{\boldsymbol{\nabla}\otimes\boldsymbol{\nabla}}
{k_0^2}\right) g(\mathbf{r}-\mathbf{r}'').
\end{equation}
Inserting the Weyl expansion from Eq.~(\ref{Eq:Weyl}) for the Green's function $g(\mathbf{r}-\mathbf{r}'')$,
such that $z''< z$, yields
\begin{equation}
\mathds{G}_0^{\rm E}(\mathbf{r},\mathbf{r}'') = \int \frac{{\rm d}^2 \kappa}{(2 \pi)^2}
\frac{{\rm i} {\rm e}^{{\rm i}(\boldsymbol{\kappa}\cdot(\mathbf{x}-\mathbf{x}'') + k_{z0}(z-z''))}}{2 k_{z0}}
\Big(\mathbf{e}_{\perp}\otimes \mathbf{e}_{\perp} + \mathbf{e}_{\parallel}(-k_{z0}) \otimes
\mathbf{e}_{\parallel}(-k_{z0})\Big).
\label{Eq:free_Green}
\end{equation}
Here we have defined the unit vectors in vacuum for the TE- and TM-modes as
\begin{equation}
\mathbf{e}_{\perp} := \frac{1}{\kappa}\left(\,k_y\, , \,-k_x\, , \,0\, \right)^t
\end{equation}
and
\begin{equation}
\mathbf{e}_{\parallel}(k) := \frac{1}{\kappa k_0}
\left(\,k_x k\, , \,k_y k\, , \,\kappa^2\,\right)^t.
\label{Eq:unit_tm}
\end{equation}
Using this expression for the free Green's function in Eq.~(\ref{Eq:incident_field0}) allows us to identify
the Fourier component of the incident field, which is
\begin{equation}
\boldsymbol{\mathcal{E}}_{\rm I}(\boldsymbol{\kappa}) = -\omega \mu_0 \frac{{\rm e}^{-{\rm i}(\boldsymbol{\kappa}\cdot
\mathbf{x}'' + k_{z0} z'')}}{2 k_{z0}} \Big(\mathbf{e}_\perp\otimes\mathbf{e}_\perp
+\mathbf{e}_\parallel(-k_{z0})\otimes \mathbf{e}_\parallel(-k_{z0})\Big) \cdot \mathbf{j}_{e0}.
\label{Eq:incident_field}
\end{equation}
Now, it is a straightforward exercise to calculate the reflected fields.
Substituting Eq.~(\ref{Eq:incident_field}) into Eq.~(\ref{Eq:Ref0}) gives the zeroth-order field
\begin{equation}
\boldsymbol{\mathcal{E}}_{\rm R}^{(0)}(\boldsymbol{\kappa}) = -\omega \mu_0 \frac{{\rm e}^{-{\rm i}(\boldsymbol{\kappa}\cdot
\mathbf{x}'' - k_{z0} z'')}}{2 k_{z0}} \Big( r_\perp \mathbf{e}_\perp\otimes\mathbf{e}_\perp
+ r_{\parallel} \mathbf{e}_\parallel(k_{z0})\otimes \mathbf{e}_\parallel(-k_{z0}) \Big) \cdot\mathbf{j}_{e0},
\end{equation}
where we have introduced the usual Fresnel reflection coefficients for the TE- and TM-modes, defined as
\begin{equation}
r_\perp := \frac{k_{z0}-k_z}{k_{z0}+k_z}
\end{equation}
and
\begin{equation}
r_\parallel := \frac{\varepsilon_{{\rm S}} k_{z0}-k_z}{\varepsilon_{{\rm S}}k_{z0}+k_z}.
\end{equation}
The first-order field can be calculated by substituting Eq.~(\ref{Eq:incident_field}) in Eq.~(\ref{Eq:Ref1}),
giving
\begin{equation}
\begin{split}
\boldsymbol{\mathcal{E}}_{\rm R}^{(1)}(\boldsymbol{\kappa}) &= -{\rm i}(\varepsilon_{{\rm S}}-1) \int\! \frac{{\rm d}^2 \kappa'}
{(2\pi)^2}\, F(\boldsymbol{\kappa}-\boldsymbol{\kappa}')\, \mathds{L}(\boldsymbol{\kappa'})\,
\frac{\omega \mu_0}{2 k_{z0}}\, {\rm e}^{-{\rm i} (\boldsymbol{\kappa}\cdot\mathbf{x}'' +k_{z0} z'')}\\
& \qquad\qquad\qquad\qquad \times \left(t'_\perp \mathbf{e}'_\perp \otimes \mathbf{e}'_\perp + t'_{\parallel}
\mathbf{e}'_\parallel(-k'_{z})\otimes \mathbf{e}_\parallel(-k'_{z})\right) \cdot \mathbf{j}_{e0},
\end{split}
\end{equation}
with the transmission coefficients for the TE- and the TM-modes defined as
\begin{equation}
t_\perp := \frac{2 k_{z0}}{k_{z0}+k_z}
\end{equation}
and
\begin{equation}
t_\parallel := \frac{2 k_{z0}}{\varepsilon_{{\rm S}}k_{z0}+k_z}.
\end{equation}
Due to the definition of the vector $\mathbf{e}_\parallel$ for the TM-modes
in Eq.~(\ref{Eq:unit_tm}) the TM-mode
transmission coefficient defined here does not coincide with the standard formulation
of the transmission coefficient. Finally, the reflected electric dyadic Green's function can be read off,
giving the zeroth-order expression
\begin{equation}
\mathds{G}_{\rm R}^{\rm E, 0}(\mathbf{r},\mathbf{r}'') = \int\! \frac{{\rm d}^2 \kappa}{(2\pi)^2}
\frac{{\rm i} {\rm e}^{{\rm i}(\boldsymbol{\kappa}\cdot(\mathbf{x}-\mathbf{x}'') - k_{z0}(z+z''))}}{2 k_{z0}}
\Big( r_\perp \mathbf{e}_\perp\otimes\mathbf{e}_\perp
+ r_{\parallel} \mathbf{e}_\parallel(k_{z0})\otimes \mathbf{e}_\parallel(-k_{z0}) \Big),
\label{Eq:Green_ref_0}
\end{equation}
and the first-order expression
\begin{equation}
\begin{split}
\mathds{G}_{\rm R}^{\rm E, 1}(\mathbf{r},\mathbf{r}'') &= (1- \varepsilon_{{\rm S}}) \int\!\frac{{\rm d}^2
\kappa}{(2\pi)^2} \int\! \frac{{\rm d}^2 \kappa'}{(2\pi)^2} \,F(\boldsymbol{\kappa}-\boldsymbol{\kappa}')\,
\mathds{L}(\boldsymbol{\kappa'})\,\frac{1}{2 k'_{z0}} {\rm e}^{{\rm i} (\boldsymbol{\kappa} \cdot \mathbf{x} - k_{z0} z)}
{\rm e}^{-{\rm i}(\boldsymbol{\kappa'}\cdot\mathbf{x} +k'_{z0} z)}\\
& \quad\quad \qquad \qquad \qquad \qquad \times \left(t'_\perp \mathbf{e}'_\perp \otimes \mathbf{e}'_\perp +
t'_{\parallel} \mathbf{e}'_\parallel(-k'_{z})\otimes \mathbf{e}_\parallel(-k'_{z})\right).
\end{split}
\label{Eq:Green_ref_1}
\end{equation}
Since the magnetic dyadic Green's function is linked with the electric one by means of Eq.~(\ref{Eq:Green_electric_magnetic}),
we now have all ingredients to calculate the LDOS above a structured surface up to first order.
\section{Local density of states above a structure surface}
\label{Sec4}
The zeroth- and first-order Green's functions in Eqs.~(\ref{Eq:Green_ref_0}) and (\ref{Eq:Green_ref_1})
can now be used to calculate the electric and magnetic local density of states
\begin{equation}
\begin{split}
D^{{\rm E}/{\rm H}}(\omega,\mathbf{r}) &\approx D^{{\rm E}/{\rm H}}_0(\omega,\mathbf{r}) + D^{{\rm E}/{\rm H}}_1(\omega,\mathbf{r}) \\
&:= \frac{\omega}{\pi c^2} \Im\,{\rm Tr}\,\mathds{G}^{{\rm E}/{\rm H},0}_{R}
+ h \frac{\omega}{\pi c^2} \Im\,{\rm Tr}\,\mathds{G}^{{\rm E}/{\rm H},1}_{R}
\label{Eq:LDOS_01}
\end{split}
\end{equation}
by means of the equilibrium relations in Eqs.~(\ref{Eq:equi_electric_LDOS}) and (\ref{Eq:equi_magnetic_LDOS}).
As expected, and as a confirmation of the validity of the approach,
to zeroth order we obtain the well-known expressions~\cite{JoulainEtAl03}
\begin{equation}
\label{eq:DE0}
D^{E}_0(\omega, a)=\frac{\omega}{4 \pi^2 c^2} \Im \, \int\limits_{0}^{\infty} \!
\text{d}k_x \, \frac{k_x {\rm e}^{-2\gamma a }}{2\gamma} \left( r_\perp
+\frac{2k_x^2-k_0^2}{k_0^2} r_\parallel \right)
\end{equation}
and
\begin{equation}
D^{H}_0(\omega, a)=\frac{\omega}{4 \pi^2 c^2} \Im \, \int\limits_{0}^{\infty} \!
\text{d}k_x \, \frac{k_x {\rm e}^{-2\gamma a }}{2\gamma} \left( r_\parallel
+\frac{2k_x^2-k_0^2}{k_0^2} r_\perp \right)
\end{equation}
for the LDOS at the distance $a = -z$ above a semi infinite body,
with $\gamma=\sqrt{k_x^2-k_0^2}$. The first-order contributions are
\begin{equation}
\begin{split}
D^{E}_1 (\omega,a) =& h \Im \bigg[ \frac{1-\epsilon_{\rm S}}{\omega \pi} \int\!
\frac{\text{d}^2 \kappa}{4 \pi^2} \,\int\!
\frac{\text{d}^2 \kappa'}{4 \pi^2} \,
F(\boldsymbol{\kappa}-\boldsymbol{\kappa'})\, {\rm e}^{{\rm i}(\boldsymbol{\kappa}-\boldsymbol{\kappa'})\cdot \mathbf{x}}
{\rm e}^{{\rm i}(k_{z0}+k_{z0}')a} \\
&\frac{t_\parallel'}{2 k_{z0}'}\frac{t_\parallel}{2 k_{z0}} \bigg\{ k_z
k_{z0}(\kappa'^2+k_z' k_{z0}')+k_z' k_{z0}'(\kappa^2+k_z k_{z0})\\
&-(\boldsymbol{\kappa} \cdot \boldsymbol{\kappa'})\big( k_z k_z' + \epsilon_S k_{z0} k_{z0}'
\big) + \epsilon_S \kappa^2 \kappa'^2+(\boldsymbol{\kappa} \cdot \boldsymbol{\kappa'} )^2
\bigg\} \bigg] \\
\equiv& h \Im \bigg[ \frac{1-\epsilon_{\rm S}}{\omega \pi} \int\!
\frac{\text{d}^2 \kappa}{4 \pi^2} \,\int\!
\frac{\text{d}^2 \kappa'}{4 \pi^2} \,
F(\boldsymbol{\kappa}-\boldsymbol{\kappa'})\, {\rm e}^{{\rm i}(\boldsymbol{\kappa}-\boldsymbol{\kappa'})\cdot \mathbf{x}}
I_{\rm E}(\boldsymbol{\kappa},\boldsymbol{\kappa}') \bigg]
\end{split}
\label{eq:DE1}
\end{equation}
and
\begin{equation}
\begin{split}
D^{H}_1 (\omega,a) =& h \Im \bigg[ \frac{\epsilon_{\rm S}-1}{\omega \pi} \int\!
\frac{\text{d}^2 \kappa}{4 \pi^2} \,\int\!
\frac{\text{d}^2 \kappa'}{4 \pi^2} \,
F(\boldsymbol{\kappa}-\boldsymbol{\kappa'}) \, {\rm e}^{{\rm i}(\boldsymbol{\kappa}-\boldsymbol{\kappa'})\cdot \mathbf{x}}
{\rm e}^{{\rm i}(k_{z0}+k_{z0}')a} \\
&\frac{t_\parallel'}{2 k_{z0}'}\frac{t_\parallel}{2 k_{z0}} \bigg\{ \bigg[
\frac{\kappa'^2 + k_z' k_{z0}'}{k_0^2} \big(k_{z0} k_{z0}'-\boldsymbol{\kappa} \cdot
\boldsymbol{\kappa'}\big) + k_{z0}k_z' \bigg] \big( \kappa^2 + k_{z0} k_z \big)\\
&+k_{z0}' \big(k_z-k_{z0} \big) \frac{\big(\boldsymbol{\kappa} \times \boldsymbol{\kappa'}
\big)^2}{\kappa'^2}\frac{\kappa'^2 + k_{z0}'k_z'}{k_0^2}+k_z'\big( k_z -
k_{z0} \big) \frac{\big( \boldsymbol{\kappa} \cdot \boldsymbol{\kappa'}
\big)^2}{\kappa'^2}\\
&-\epsilon_S k_0^2 \boldsymbol{\kappa} \cdot \boldsymbol{\kappa'} \bigg\} \bigg] \\
\equiv& h \Im \bigg[ \frac{\epsilon_{\rm S}-1}{\omega \pi} \int\!
\frac{\text{d}^2 \kappa}{4 \pi^2} \,\int\!
\frac{\text{d}^2 \kappa'}{4 \pi^2} \,
F(\boldsymbol{\kappa}-\boldsymbol{\kappa'}) \,{\rm e}^{{\rm i}(\boldsymbol{\kappa}-\boldsymbol{\kappa'})\cdot \mathbf{x}}
I_{\rm H}(\boldsymbol{\kappa},\boldsymbol{\kappa}') \bigg].
\end{split}
\label{eq:DH1}
\end{equation}
From these two equations~(\ref{eq:DE1}) and (\ref{eq:DH1}) it can be seen that in first order the electric
and magnetic local density of states
are given by an integral of the Fourier transform
\begin{equation}
F(\boldsymbol{\kappa} - \boldsymbol{\kappa}') = \int \!\! {\rm d} x \int\!\! {\rm d} y \, f(x,y) \,{\rm e}^{- {\rm i} \mathbf{x} \cdot (\boldsymbol{\kappa} - \boldsymbol{\kappa}')}
\end{equation}
of the profile $f(x,y)$, multiplied by the exponential factor $\exp({\rm i} (\boldsymbol{\kappa} - \boldsymbol{\kappa}') \cdot\mathbf{x} )$
and the function $I_{\rm E}(\boldsymbol{\kappa},\boldsymbol{\kappa}')$ or $I_{\rm H}(\boldsymbol{\kappa},\boldsymbol{\kappa}')$, respectively.
Due to the factor $\exp({\rm i}(k_{z0}+k_{z0}')a)$, these two functions are exponentially damped for
$\kappa$ and $\kappa'$ with modulus much greater than the inverse observation
distance $a^{-1}$ in the near field, since in the evanescent regime $k_{z0} \approx {\rm i} \kappa$ and $k_{z0}' \approx {\rm i} \kappa'$.
Therefore, the smaller the observation distance the more Fourier components
contribute to the integral. Hence, from the structure of the first-order integrals it can be expected that for
distances smaller than the characteristic width of the surface profile function the electric and magnetic local
density of states resembles the surface profile.
This behaviour is confirmed in Fig.~\ref{fig:wuerfel}, where the numerical evaluation of
the electric density of states from Eq.~(\ref{Eq:LDOS_01}) above a square pad with a height $h = 5\, {\rm nm}$
and width $w = 15 \,{\rm nm }$ on a plane surface is plotted. Here, we have used the
frequency $\omega = 10^{14}\,{\rm s}^{-1}$ and the Drude permittivity $\epsilon_{{\rm S}}$ of gold.
Obviously, at an observation distance of $a = 40\, {\rm nm}$ the values of the local density of states resemble
a two dimensional bell-shaped function, whereas at an observation distance of $a = 5.5 \,{\rm nm}$,
being much smaller than the width of the square pad, the values of the local density
mimic the underlying structure, apart from softened edges.
\begin{figure}[Hhbt]
\centering
\subfigure[Observation distance $a = 40 \, \text{nm}$]
{\epsfig{file=Fig3a.eps, width = 0.48\textwidth}
}
\subfigure[Observation distance $a = 5.5 \, \text{nm}$]
{\epsfig{file=Fig3b.eps, width = 0.48\textwidth}
}
\caption{Electric LDOS from eq.~(\ref{Eq:LDOS_01}) above a square pad with an edge length of $15 \, \text{nm}$
and a height of $h = 5 \, \text{nm}$ for two different observation distances $a$ from the base plane.
The frequency considered is $\omega = 10^{14} \text{s}^{-1}$.}
\label{fig:wuerfel}
\end{figure}
In the following, we discuss the magnitude of the contributions of the zeroth and first order of the electric and
magnetic local density of states. For that purpose, we assume an infinitely extended bar with a width $w$
on a plane surface modelled by the profile function
\begin{equation}
f(x,y) = \frac{1}{\exp(d (|x|- \frac{w}{2}))+1},
\label{Eq:profile}
\end{equation}
assuming $d = 10^9 \, \text{m}^{-1}$ and $w = 30 \, \text{nm}$. The height of the bar (see Fig.~\ref{Fig:profile})
is chosen to be $h = 5 \, \text{nm}$.
\begin{figure}[Hhbt]
\centering
\epsfig{file=Fig4.eps, width = 0.47\textwidth}
\caption{Plot of $h f(x,y)$ with the profile function defined by Eq.~(\ref{Eq:profile}) with
$d = 10^9 \, \text{m}^{-1}$, $w = 30 \, \text{nm}$, and $h = 5 \, \text{nm}$.}
\label{Fig:profile}
\end{figure}
The numerical values of the local density of states at a constant observation distance $a = 10\, {\rm nm}$
above the base plane are shown in Fig.~\ref{fig:part_consth},
using again the frequency $\omega = 10^{14} {\rm s}^{-1}$ and the Drude permittivity for
gold. Firstly, one observes that the values of the magnetic LDOS are much greater than the
values of its electric counterpart, which is typical for metals, whereas for a polar dielectric
bodies the electric LDOS usually gives greater values than the
magnetic one. Secondly, at the given observation distance we find values for the first-order contribution
to the electric LDOS, which are of the same order of magnitude
as the zeroth-order contribution. On the other hand, the values of the first-order magnetic LDOS are
significantly smaller than the zeroth-order values for all lateral positions. Thirdly, the values of
$D^{{\rm E}}_1$ and $D^{{\rm H}}_1$ give an equally good image of the bar on the plane surface,
the width of the two bell-shaped curves being approximately the same. Therefore, in this case both first-order contributions give
qualitatively similar results.
\begin{figure}[Hhbt]
\centering
\subfigure[Electrical LDOS]{
\epsfig{file=Fig5a.eps, width = 0.47\textwidth}
}
\subfigure[Magnetical LDOS]{
\epsfig{file=Fig5b.eps, width = 0.47\textwidth}
}
\caption{LDOS of an gold half space structured by a gold bar (see Fig.~\ref{Fig:profile}). The
frequency used is $\omega=10^{14}\,{\rm s}^{-1}$, and the observation distance $10 \,
\text{nm}$. Observe the different scales for the electric LDOS (a) and the magnetic LDOS (b)}
\label{fig:part_consth}
\end{figure}
In Fig.~\ref{fig:z-spekto} we plot
the ratio of the first- and zeroth-order LDOS $D_1/D_0$ for the same surface profile directly above
the bar, i.e.\ at $x = 0$, for observation distances ranging from $6\, {\rm nm}$ to $100 \,{\rm nm}$.
As expected, this ratio increases for decreasing observation distance, suggesting that for small
distances higher-order terms have to be considered. These higher-order terms can in principle
be calculated by an iterative scheme deduced by Greffet~\cite{Greffet1988}. We
suggest that it might be sufficient to consider only $D_0$ and $D_1$ as long as the ratio $D_1/D_0$
does not exceed the ten percent level. This means that the numerical results depicted in Figs.~\ref{fig:part_consth} a)
and b) refer to distances where higher-order terms should be taken into account.
Furthermore, it is evident from Fig.~\ref{fig:z-spekto} that for all distances the ratio of the leading
two contributions to the magnetic LDOS gives much smaller values than the corresponding ratio
for the electric LDOS. Therefore, it can be concluded
that the approximation committed when considering only the zeroth- and first-order terms holds
for the magnetic part for much smaller distances than for the electric part. This also means
that the underlying structure becomes important in the electric LDOS for much greater distances
than in the magnetic LDOS.
\begin{figure}[Hhbt]
\centering
\begin{minipage}[t]{0.9\textwidth}
\epsfig{file=Fig6.eps, width = 0.6\textwidth}
\caption{Dependence of the ratio of the first and zeroth order of the LDOS
on the observation distance $a$ calculated above the profile shown in Fig.~\ref{Fig:profile}
for $x = 0$ and frequency $\omega= 10^{14} \, \text{s}^{-1}$.}
\label{fig:z-spekto}
\end{minipage}
\end{figure}
So far we have determined the LDOS at a constant height above the base plane of the structured
surface. A very common operation mode in near-field microscopy is the constant-distance mode, where the
separation between the tip and the individual features of the sample is kept
constant~\cite{WishnathEtAl08}. In order to calculate the LDOS relevant for this mode, we use the observation
distance $a+h f(x,y)$ instead of $a = const.$ in Eq.~(\ref{Eq:LDOS_01}).
Fig.~\ref{fig:part_constd} shows the LDOS obtained for this
constant-distance mode above the structured half space depicted in Fig.~\ref{Fig:profile}.
\begin{figure}[Hhbt]
\centering
\begin{minipage}[t]{0.9\textwidth}
\epsfig{file=Fig7.eps, width = 0.6\textwidth}
\caption{Electric local density of states ($D^{\rm E}_0$: dash-dotted line; $D^{\rm E}_1$: dashed line;
$D^{\rm E}_0 + D^{\rm E}_1$: solid line) above the structured surface as
depicted in Fig.~\ref{Fig:profile} at a frequency of
$\omega=10^{14} \, \text{s}^{-1}$ and a constant separation of $10 \,
\text{nm}$ between observation point and surface.}
\label{fig:part_constd}
\end{minipage}
\end{figure}
In the constant-height mode (see Fig.~\ref{fig:part_consth} (a) and (b)) the first-order term
$D_1$ gives a rough image of the underlying surface structure, while the
zeroth-order term $D_0$ is constant. However, in the constant-distance
mode (see Fig.~\ref{fig:part_constd}) the qualitative
behaviour of the two contributions is more complex, due to the variation of the observation distance.
Here, the zeroth-order term $D_0$ coincides more or less with the inverse of the underlying structure,
and the first-order term $D_1$ gives
relatively large values near the edge of the bar, due to the variation in observation distance. Therefore,
at least two regimes of observation distances can be distinguished: At large distances, where $D_1/D_0$ is small, the LDOS is dominated
by the zeroth-order term, giving values which coincide approximately with the inverse of the surface
structure. At distances where at least the first-order term has to be taken into account, the variation in distance
leads to a rather complex pattern of the LDOS.
\section{Consequences for the near-field scanning thermal microscope}
\label{Sec5}
Up to this point, only the LDOS at a frequency corresponding to $300 \, \text{K}$ have been
calculated. Before we discuss some numerical results for the full near-field radiative
heat transfer between a metallic nano sphere and a structured surface, we
emphasize the restrictions of the model: Firstly, due to the dipole approximation this
model is only valid for distances $a$ significantly greater than the radius of the sphere $R$.
Secondly, for distances and sphere radii smaller than the mean free path of the conduction
electrons, nonlocal and quantum mechanical effects become important and have
to be implemented. Thirdly, due to the perturbative approach the results apply under the condition
that the height $h$ of the profile is the smallest length scale, i.e.\ $h \ll \min\{a,\lambda_{\rm th},w\}$
with the width of the surface structure $w$ and the thermal wavelength $\lambda_{\rm th} \approx \hbar \beta c$.
In particular this means that the pertubative approach is not valid for
distances $a \ll R$; in this limit the ``proximity approximation'' can be
utilised~\cite{VolokitinPersson2007}. Nonetheless, it is reasonable to explore the perturbative
predictions in some detail.
To this end, we evaluate the near-field radiative heat transfer between a gold nanoparticle
and a gold sample as given by Eq.~(\ref{Eq:DipolModel}), using the polarisabilities
of the sphere~(\ref{Eq:alpha_E}) and (\ref{Eq:alpha_H}) with the
Drude permittivities $\epsilon_{\rm P}$ and $\epsilon_{\rm S}$ for gold. As surface
profile we employ the infinitely extended bar from Eq.~(\ref{Eq:profile}) with parameters as
in Fig.~\ref{Fig:profile}. In order to relate the results for the energy flow $P$ to the numerical results for the LDOS we evaluate
Eq.~(\ref{Eq:DipolModel}) for a constant height, $a = 10\, {\rm nm}$, and a
constant distance, $a = 10\, {\rm nm} + h f(x,y)$, where
a nanoparticle radius of $R = 10 \,{\rm nm}$ has been chosen. As shown in Fig.~\ref{fig:p}, in both cases
the near-field radiative heat transfer is dominated by the magnetic contribution, this
being a typical feature for good metals like gold (at the given distance). Apart from
this the curves displayed in Fig.~\ref{fig:p} reflect the corresponding plots of the LDOS
at a frequency near the thermal frequency $\omega_{\rm th}$.
This is a consequence of Eq.~(\ref{Eq:DipolModel}), since the main contributions to the frequency
integral stem from frequencies near the thermal frequency as long as there are no resonances
in the thermally accessible frequency regime. We also implemented the ``proximity approximation''
for the given geometry numerically (with the same parameters as in Fig.~\ref{fig:p}), obtaining qualitatively
similar results for the near-field radiative heat transfer as those given by the dipole model.
\begin{figure}[Hhbt]
\centering
\subfigure[Heat transfer in the constant-height mode.]
{\epsfig{file=Fig8a.eps, width = 0.48\textwidth}
}
\subfigure[Heat transfer in the constant-distance mode.]
{\epsfig{file=Fig8b.eps, width = 0.48\textwidth}
}
\caption{Near-field radiative heat transfer between the structured sample
depicted in Fig.~\ref{Fig:profile} ($T_{\rm S} = 100 \, \text{K}$) and
a spherical probe ($T_{\rm P} = 300 \, \text{K}$) with a radius of
$10 \, \text{nm}$. The height and the separation was $10 \, \text{nm}$ in each case.}
\label{fig:p}
\end{figure}
Hence, the conclusions deduced in the last section for the local density of states can be transfered to
the near-field radiative heat transfer between a sphere and a structured surface, modelling a NSThM tip
and a structured sample. Thus, for example we conclude that the lateral resolution of a NSThM
increases for decreasing distances, as long as this microscope can be described within the dipole model.
The lateral resolution is comparably good for the electric part $P_{\rm E}$ and the magnetic part $P_{\rm H}$, with
the topology of the surface profile being well resolvable for distances much smaller than the width of the surface pattern.
On the other hand, for non-metallic materials, i.e.\ in a situation with $P_{\rm E} \gg P_{\rm H}$, the underlying structure
becomes important for much greater distances than for metallic materials, as can be concluded from Fig.~\ref{fig:z-spekto}.
Furthermore, a NSThM gives very different images of the underlying structure when either the constant-height or
the constant-distance mode is used. In particular, for distances where $D_0 \gg D_1$ the values for $P$ should
resemble the inverse surface structure in constant-distance mode, whereas for distances
where $D_1$ becomes important, the measured signal can be rather complex in that mode.
On the other hand, in constant-height mode the signal should approximately resemble the
topology of the underlying structure at distances where $D_1$ becomes important. These are strong predictions
that are amenable to immediate verification in current experiments~\cite{WishnathEtAl08}, and may help to correctly
interprete the various types of signals obtainable with a NSThM~\cite{Preprint08}.
\acknowledgments
This work was supported in part by the Deutsche Forschungsgemeinschaft through grant No.\ KI 438/8-1.
We also thank M.~Holthaus for useful comments on the text.
|
\section{introduction}
WZ Sge-type stars are the close binary systems at a latest stage
of evolution. They consist from a white dwarf (WD) accreting
matter from a late (probably brown dwarf) type secondary. These
binaries are attractive objects for search for the white dwarf
pulsations. During outburst the observed brightness variability of
WZ Sge type stars is caused mostly by the thermal and tidal disk
instability. The orbital modulation also could contribute in
high-inclined systems during this state. The probability to detect
of the WD pulsations depends on the quantity of sources of
variable radiation and on the temperature of WD at the time of
observation. The quiescent state seems to be more preferable for
this task because of reduced number of sources of variable
radiation. However it is difficulty to predict when the white
dwarf should pass its instability strip after the outburst for
each individual case.
We conducted the observations of five WD Sge-type dwarf novae
J080434.20+510349.2, SDSS J102146.44+234926.3, V1108 Her, ASAS
J0025 and WZ Sge itself in 3--7 years after their outbursts and
present the analysis of activity of these systems.
\section{Observations}
CCD observations of the five WZ Sge-type stars were carried out
using different telescopes of the Crimean astrophysical
observatory (Ukraine) and Terskol observatory (Russia). The
observations were made in $R$ spectral band or in unfiltered
light. The log of observations is given in Table 1, where the site
of observations, telescopes, date of the outburst preceding
observation and date of observations are presented.
\section{SDSS J080434.20+510349.2}
This object contains the accreting WD pulsator. The first
appearance of 12.6-min pulsations were detected in 8-9 month after
the 2006 outburst [1]. We continued a monitoring of SDSS
J080434.20+510349.2 (hereafter, SDSS J0804) in 2009. The original
data folded on the orbital period 0.0590048 d [2] are shown in
Fig. 1 (left). The orbital profile is a two-humped one, sometimes
one of humps could be more or less divided into two (JD 2454852
and JD 2454882). It is seen that in some occasions the orbital
modulation is slightly affected by another short-term variability.
Periodograms calculated for these data (using ISDA package [3])
are shown in Fig. 2 (right). Besides of peaks pointing to the
orbital period (Porb) and its half value Porb/2, the periodograms
for only two from six rows of observations (JD 244852 and JD
244882) contain peak at period 740 s (12.33 min) and it's twice
value. In the limits of width of peak it coincides with the known
period of pulsation 12.6 min. So in $\sim$ 3 year after the 2006
outburst the appearance of pulsations became irregular. It is
worth to note that the complex structure of the orbital light
curve produce additional peak on the light curve at Porb/4. More
of a linear combinations of the orbital and pulsation frequencies
appear for these data.
\begin{table}
\begin{tabular}{lrrrr}
\hline
& \tablehead{1}{r}{b}{Site}
& \tablehead{1}{r}{b}{Year of\\
outburst}
& \tablehead{1}{r}{b}{Year of \\observations}\\
\hline
SDSS J080434.20+510349.2 & CrAO (2.6-m), Terskol (2-m) & 2006 & 2009\\
V1108 Her & CrAO (2.6-m) & 2004 & 2008\\
SDSS J102146.44+234926.3 & CrAO (2.6-m) & 2006 & 2009\\
ASAS J0025 & Terskol (60 cm) & 2004 & 2007\\
WZ Sge & CrAO (2.6-m, 1.25-m) & 2001 & 2008\\
\hline
\end{tabular}
\caption{Log of observations for five WZ Sge-type stars}
\label{tab:a}
\end{table}
\begin{figure}
\includegraphics[height=0.7\textheight]{eppavlenko01.EPS}
\caption{Left: original data of SDSS J0804 folded on the orbital period of 0.0590048 day. Right:
corresponding periodograms. The periodograms for original data are designed by thin line while
these for data after orbital period subtraction, by thick line}
\end{figure}
\section{V1108 Her}
The last known measuring of the orbital period of V1108 Her was
done by Price et al. [4] in a few days after the end of the 2004
outburst. Accordingly to their observations, period was
0.056855(69) d. We obtained our observation in 2008, believing
that in 4 years after the 2004 outburst the data may not be
affected by positive superhumps. The periodogram for the data
covering 8 nights of observations is shown in Fig. 2 (left). The
strongest peak points to period 0.05672(4) d. The corresponded
fractional period excess [5] (for period of superhump 0.057780 d
[4]) is 0.0187, so the mass ratio of components is 0.09. This
confirms the previous suggestion of Price et al. [4] that V1108
Her could be the "period bouncer". The data folded on the orbital
period are shown in Fig. 2 (right). One could see that the main
brightness variations are caused by two-humped orbital variations
with unequal mean amplitudes of neighbor humps ($0^{m}.05$ and
$0^{m}.025$). The shape of the orbital light curve resembles these
of high-inclined WZ Sge-type systems.
\begin{figure}
\includegraphics[height=0.4\textheight]{eppavlenko02.EPS}
\caption{Left: periodogram of V1108 Her for 4 nights of observations in 2008.
Frequency is expressed in units of 1/day. Right: data folded on the orbital period}
\end{figure}
\section{SDSS J102146.44+234926.3, ASAS J0025 and WZ Sge}
The true orbital period of SDSS J102146.44+234926.3 is unknown.
Our observations in 2009 did not reveal expected orbital
modulation. Instead for some nights we detected the brightness
variations with 20.6 and 20 min period (Fig. 3) or with the
multiple of them. The mean amplitude for Jan 20/21, 2009 was
$0^{m}.1$. The source of this irregular periodicity is unclear.
Potentially it could be connected with instability of outer parts
of accretion disk, or with WD pulsations, or equal to the 1/4 of
expected value of orbital period.
ASAS J0025 displays a high-amplitude nightly variability with
amplitude up to $0^{m}.3$ (Fig. 4). However the data obtained in
2007 can't confirm the 0.05654 d candidates to the orbital period
suggested by Templeton et al. [6]. Instead a prominent variability
in a vicinity of suggested orbital period was observed (Fig. 5).
Such behavior resembles the variability often observing in a
low-inclined systems (for example in MV Lyr [7]) connected
probably with instability in accretion disk.
WZ Sge is eclipsed system. In 7 years after 2001 outburst the main
brightness variations are caused by orbital modulation with period
of 81-min [8]. The example of two-humped light curve is presented
in Fig. 6. The amplitude of humps varies from cycle to cycle in a
region of $0^{m}.15-0^{m}.3$. There are variations of much smaller
amplitude superimposing the orbital light curves.
\begin{figure}
\includegraphics[height=0.36\textheight]{eppavlenko03.EPS}
\caption{Periodogram of SDSS J102146.44+234926.3 for different data in 2009}
\end{figure}
\begin{figure}
\includegraphics[height=0.2\textheight]{eppavlenko04.EPS}
\caption{The example of nightly light curve of ASAS J0025 in 2007}
\end{figure}
\begin{figure}
\includegraphics[height=0.25\textheight]{eppavlenko05.EPS}
\caption{The periodogram of ASAS J0025 for 4 nights in 2007}
\end{figure}
\begin{figure}
\includegraphics[height=0.3\textheight]{eppavlenko06.eps}
\caption{Example of the WZ Sge light curve in 2008}
\end{figure}
\section{Conclusion}
We considered the behavior of five WZ Sge-type stars in several
years after their outburst. The main brightness variations of the
high-inclined systems (WZ Sge, SDSS J0804, V1108 Her) is caused by
the orbital modulation. In SDSS J0804 in some occasions the WD
pulsations superimpose the orbital light curves. SDSS
J102146.44+234926.3 and ASAS J0025 display brightness variability
which nature is unclear.
\begin{theacknowledgments}
E. Pavlenko and A. Golovin are grateful to the LOC for possibility to
participate in the conference.
This work was partially supported by grant 28.2/081 of
Ukrainian Fund of Fundamental Researches.
\end{theacknowledgments}
\bibliographystyle{aipproc}
|
\section{Introduction}
By a surface we always mean a complete, compact and connected $2$-dimen\-si\-o\-nal riemannian manifold without boundary.
Unless explicitly stated otherwise, all our graphs will be connected, undirected,
and may have multiple edges and loops but not vertices of degree two.
The notion of cut locus was introduced by H. Poincar\'e \cite{p} in 1905, and gain since then an important place in global riemannian geometry.
The {\it cut locus $C(x)$ of} the point $x$ in the riemannian manifold $M$ is the set of all extremities (different from $x$) of maximal (with respect to inclusion) shortest paths starting at $x$; for basic properties and equivalent definitions refer, for example, to \cite{ko} or \cite{sa}.
S. B. Myers \cite{M} established that the cut locus of a real analytic riemannian surface is (homeomorphic to) a graph,
and a partial converse to his result was proven in \cite{iv2}, namely that every (metric) connected graph can be realized as a cut locus; i.e., there exist a riemannian surface $S_G=(S_G,h)$ and a point $x \in S_G$ such that $C(x)$ is isometric to $G$.
If moreover $G$ is cyclic of order $k$, then it can be realized on a surface of constant curvature \cite{iv2}.
In \cite{iv1} we introduced the notion of cut locus structure on a graph, and discussed its basic properties, while in \cite{iv3} we proposed upper bounds on the number of such structures.
The generic behaviour of cut locus structures is also presented in \cite{iv2}.
In this paper, we are concerned about the orientability of the surfaces $S_G$ realizing the graph $G$ as a cut locus.
If $G$ has an odd number of generating cycles then any surface realizing $G$ is non-orientable.
If the number of generating cycles is even then one cannot generally distinguish, by simply looking to the graph $G$, whether $S_G$ is orientable or not.
In other words, seen as a graph, {\sl the cut locus does not encode the orientablity of the ambient space}; this is our main motivation to introduce ``cut locus structures'' on graphs.
In order to characterize the orientability of a surface $S_G$ realizing $G$ as a cut locus,
we codify in a cut locus structure any small neighbourhood of the cyclic part $G$ of $C(x)$ in $S_G$.
In Section \ref{CL-str} we briefly present the notion of cut locus structure,
in Section \ref{crit_orient} we characterize those cut locus structures living on orientable surfaces, while
and in Section \ref{or_graphs} we give several criteria to provide {\it orientably realizable} graphs (i.e., graphs having at least one orientable realization).
For example, we obtain (Corollary \ref{Petersen}) that the Petersen graph has at least two orientable realizations.
As an application of our results in $\S$\ref{crit_orient}, in Section \ref{S_4cy} we classify all
distinct orientable cut locus structures on graphs with four generating cycles.
In particular, this enables us to point out such a graph which is not orientably realizable.
\bigskip
At the end of this section we recall a few definitions and facts about graphs, to fix notation.
We shall denote by $m(G)$ the number of edges in the graph $G$, and by
$n(G)$ the number of its vertices.
An edge in $G$ is called {\it external} if it is incident to (least) one vertex of degree one, and is called a {\it bridge} if its removal disconnects $G$.
Denote by $B$ the set of all {\it bridges} in the graph $G$.
Each non-vertex component of $G \setminus B$ is called a $2$-{\it connected component} of $G$.
A $k${\it -graph} is a graph all of whose vertices have degree $k$.
In particular, a $3$-graph will also be called {\it cubic}.
An {\it edge contraction} in the graph $G$ is an operation which removes an edge from $G$ while simultaneously merging together the two vertices it used to connect to a new vertex; all other edges incident to either of the two vertices become incident to the new vertex.
Consider the graph $G$ as a simplicial complex;
the {\it cyclic part} of $G$ is the minimal (with respect to inclusion) subgraph $G^{cp}$ of $G$,
to which $G$ is contractible; i.e., the minimal subgraph of $G$ obtained by repeatedly contracting external edges,
and for each vertex remaining of degree two (if any) merging its incident edges.
A graph is called {\it cyclic} if it is equal to its cyclic part.
The power set ${\cal E}$ of $E$ becomes a $Z_2$-vector space over the two-element field $Z_2$,
if endowed with the symmetric difference $\ast$ as addition, and it is called the {\it edge space} of $G$.
The {\it cycle space} is the subspace ${\cal Q}$ of ${\cal E}$ generated by (the edge sets of) all the simple cycles of $G$.
If $G$ is seen as a simplicial complex, ${\cal Q}$ is the space of $1$-cycles of $G$ with mod $2$ coefficients.
The symmetric difference $\ast$ of two simple cycles is either a simple cycle or a union of edge-disjoint simple cycles.
The dimension $q=q(G)$ of the cycle space of the graph $G$ is given by $q(G)=m(G)-n(G)+1$.
\section{CL-structures}
\label{CL-str}
In this section we briefly present the notion of cut locus structure, that we introduced in \cite{iv1}.
\begin{df}
A $G$-{\sl patch} on the graph $G$ is a topological surface $P_G$ with boundary, containing (a graph isometric to) $G$ and contractible to $G$.
A $G$-{\sl strip} is a $G$-patch whose boundary is topologically a circle.
A {\sl cut locus structure} (shortly, a {\sl CL-structure}) on the graph $G$ is a strip on the cyclic part $G^{cp}$ of $G$.
A CL-structure on $G$ is {\sl orientable} if the corresponding $G^{cp}$-strip is an orientable surface.
\end{df}
\begin{df}
An {\sl elementary decomposition} of a $G$-patch $P_G$ is a decomposition of $P_G$ into {\sl elementary strips} such that:
\\- each edge-strip corresponds to precisely one edge of $G$;
\\- each point-strip corresponds to precisely one vertex of $G$.
\end{df}
Denote by ${\cal P}$ and ${\cal A}$ the set of all elementary point-strips, respectively edge-strips,
of a CL-structur ${\cal C}$e on the graph $G$.
\begin{df}
Consider an elementary decomposition of the $G$-strip $P_G$ such that each elementary strip has a distinguished face,
labeled $\bar 0$. The face opposite to the distinguished face will be labeled $\bar 1$.
Here, $\bar 0$ and $\bar 1$ are the elements of the $2$-element group $(Z_2, \oplus)$.
To each pair $(v,e) \in V \times E$ consisting of a vertex $v$ and an edge $e$ incident to $v$,
we associate the $Z_2$-sum $\bar s (v,e)$ of the labels of the elementary strips $\nu \in {\cal P}$,
$\varepsilon \in {\cal A}$ associated to $v$ and $e$; i.e.,
$\bar s (v,e) = \bar 0$ if the distinguished faces of $\nu$ and $\varepsilon$ agree to each other, and $\bar 1$ otherwise.
Therefore, to any cut locus structure ${\cal C}$ we can associate a function $s_{\cal C} :E \to \{\bar 0,\bar 1\}$,
\begin{eqnarray}
\label{companion}
s_{\cal C}(e)= \bar s (v,e) \oplus \bar s (v',e),
\end{eqnarray}
where $v$ and $v'$ are the vertices of the edge $e \in E$.
We call the function $s_{\cal C}$ defined by (\ref{companion}) the {\sl companion function} of ${\cal C}$.
An edge-strip $P_e$ (or simply an edge $e$) in a CL-structure ${\cal C}$ is called {\sl switched} if $s_{\cal C} (e)=\bar 1$.
\end{df}
\begin{df}
\label{diff_CL}
The {\sl CL-structures} ${\cal C}$, ${\cal C}'$ on $G$ are called {\sl equivalent}
if their characteristic functions are equivalent on every $2$-connected component $H$ of $G$;
i.e., on every $H$ either $s_{\cal C} = s_{{\cal C}'}$, or $s_{\cal C} =\bar 1 \oplus s_{{\cal C}'}$.
\end{df}
We shall use the following way to planary represent a CL-structure ${\cal C}$ on the $3$-graph $G$ \cite{iv1}:
we represent in the plane each vertex-strip such that its distinguished face is ``up'',
and afterward connects the vertex-strips by edge-strips.
To schematically represent this, we shall overwrite an ``{\rm x}'' to the drawn image of an edge if its strip is switched,
and an ``{\rm =}'' to the drawn image of a edge if its strip is not-switched.
If, moreover, $G$ is planar then one can use any planar representation of $G$
to schematically represent any $G$-strip.
\bigskip
We explain now the relationship between patches and cut locus realizations of graphs.
Assume that the cut locus $C(x)$ of the point $x$ in the surface $S$ is isometric to the graph $G$.
Then, cutting off the surface an open intrinsic disc of radius smaller than the injectivity radius at $x$,
provides a strip on $G$, and consequently on $G^{cp}$, called
{\it the cut locus natural structure of} $x$, and denoted by $CLNS(x)$.
The converse is established by the following result, allowing us to consider strips whenever we think about cut locus realizations of graphs.
\begin{thm}
\label{glue} {\rm \cite{iv2}}
For every graph $G$ there exists at least one $G$-strip, and
each $G$-strip provides a realization of $G$ as a cut locus.
\end{thm}
We shall implicitly use the following simple result, easily obtained from the above considerations.
\begin{lm}
\label{cy-sw}
Let $G$ be a $3$-graph. In any planar representation of a $G$-strip,
each cycle-patch contains at least one switched edge-strip.
\end{lm}
\section{Orientable cut locus structures}
\label{crit_orient}
We are concerned next about the orientability of the surfaces $S_G$ realizing the graph $G$ as a cut locus.
If $G$ has an odd number of generating cycles then any surface realizing $G$ is non-orientable,
because any cut locus on an orientable surface of genus $g$ has $2g$ generating cycles.
If the number of generating cycles is even then one cannot generally distinguish,
by simply looking to $G$, whether $S_G$ is orientable or not, see Example \ref{Non_orient} or Theorem \ref{4}.
\bigskip
The following result expresses formally the simple fact that a circle-patch is an orientable surface if and only if
it has an even number of switches.
Together with Theorem \ref{or_cy}, it will be repeatedly applied in Section \ref{S_4cy}.
\begin{lm}
\label{sum_e}
A patch $P$ over a cycle $C$ is an orientable surface if and only if
\begin{equation}
\label{oplus_cycle_patch}
\oplus_{e \in E(C)} s(e)= \bar 0 .
\end{equation}
\end{lm}
\begin{lm}
\label{prod_or_cy}
The product $\ast$ of cycles defines a natural operation for the cycle-patches.
In particular, if the cycle $C$ is the product of the cycles $C_1, ..., C_k$,
$C=C_1 \ast ...\ast C_k$, each of which given with an orientable patch, then the induced $C$-patch is orientable.
\end{lm}
\noindent{\sl Proof:}
We prove the result by induction over $k$.
Assume first $k=2$, and let $\varepsilon_1$ be the sum of switches in $C_1 \setminus C_2$, $\varepsilon_2$ be the sum of switches in $C_2 \setminus C_1$, and $\varepsilon_{12}$ the sum of switches in $C_1 \cap C_2$.
By Lemma \ref{sum_e}, the $C_1$-patch is orientable if and only if
\begin{equation}
\label{C1}
\varepsilon_1 \oplus \varepsilon_{12} = \bar 0,
\end{equation}
the $C_2$-patch is orientable if and only if
\begin{equation}
\label{C2}
\varepsilon_2 \oplus \varepsilon_{12} = \bar 0,
\end{equation}
and respectively the $C=C_1 \ast C_2$-patch is orientable if and only if
\begin{equation}
\label{C1C2}
\varepsilon_1 \oplus \varepsilon_2 = \bar 0.
\end{equation}
But the equation (\ref{C1C2}) follows by simply adding (\ref{C1}) and (\ref{C2}).
Assume now that the statement is true for $k$ cycle patches; in order to prove it for $k+1$ cycle patches, simply put
$C'_1=C_1 \ast...\ast C_k$, $C'_2=C_{k+1}$, and apply the case $k=2$ to $C'_1$ and $C'_2$.
\hfill $\Box$
\bigskip
Recall that for any surface $S$ and any point $x$ in $S$, $C(x)$, if not a single point, is a local tree (i.e., each of its points $z$ has a neighbourhood $V$ in $S$ such that the component $K_z(V)$ of $z$ in $C(x)\cap V$ is a tree).
A {\it tree} is a set $T$ any two points of which can be joined
by a unique Jordan arc included in $T$.
Eventhough a cut locus $C(x)$ may be quite large a set
(see \cite{GS} or \cite{H2} for examples of non triangulable cut loci),
its {\it cyclic part} is a cyclic graph, see for example \cite{Itoh-Zamfirescu}.
The {\it tangential cut locus} of the point $x \in S$ is the boundary of
the maximal (with respect to inclusion) domain in the tangent space $T_xS$ to $S$ at $x$, to which the exponential map at $x$ is a diffeomorphism.
The last part of the following preliminary result has some interest in its own.
\begin{lm}
\label{n-or_crit}
A surface $S$ is non-orientable if and only if for any point $x$ in $S$ there exists an edge $e$ of $C(x)^{cp}$ whose two images in the tangential cut locus have the same orientation. Moreover, such an edge $e$ cannot be a bridge of $C(x)^{cp}$.
\end{lm}
\noindent{\sl Proof:}
The ``if and only if'' part is clear.
For the last part, fix some point $x$ in $S$ and assume $e=12$ is a bridge in $G=C(x)^{cp}$, the removal of which yields two subgraphs $G_1$, $G_2$ of $G$, with $i$ a vertex of $G_i$, $i=1,2$.
Let $P_i$ be the patch induced by $CLNS(x)$ along $G_i$, $i=1,2$.
Now regard the boundary of $CLNS(x)$ as a directed curve $O$ homeomorphic to a circle.
$O$ enters $P_1$ at $1$, covers twice all edges in $G_1$, and
exits $P_1$ again at $1$; afterwards it goes along $e$ to $2$,
enters $P_2$ at $2$, covers twice all edges in $G_2$, and
exits $P_2$ again at $2$. Therefore, with $O_i=O \cap P_i$, $i=1,2$,
the order along $O$ is $e=21$, $1$, $O_1$, $1$, $e=12$, $2$, $O_2$, $2$.
Since this is also the corresponding order in the tangential cut locus, the proof is complete.
\hfill $\Box$
\bigskip
One can roughly figure out the following statement by the fact that a non-orientable surface has only one ``face''.
\begin{thm}
\label{crit}
The surface $S_G$ realizing the connected graph $G$ as a cut locus of the point $x$ in $S$ is non-orientable if and only if there exists a cycle $C$ of $G$ and a non-orientable $C$-patch induced by $CLNS(x)$.
\end{thm}
\noindent{\sl Proof:}
By Lemma \ref{n-or_crit}, the surface $S_G$ is non-orientable if and only if there exists an edge $e$ in $G$ whose two images in the tangential cut locus of $x$ have the same orientation.
Again by Lemma \ref{n-or_crit}, such an edge can always be included in a cycle, whose induced patch is consequently non-orientable.
\hfill $\Box$
\begin{co}
\label{loop_at_3}
If the cyclic part of $G$ has a loop at a degree three vertex then $S_G$ is non-orientable.
\end{co}
\noindent{\sl Proof:} All CL-structures on $G$ induce the same patch along a loop at a degree three vertex of $G$,
which is non-orientable by Lemma \ref{sum_e}, and therefore $S_G$ is non-orientable by Theorem \ref{crit}.
\hfill $\Box$
\begin{ex}
\label{loop_at_4}
There exist orientable realizations of graphs having loops at vertices of degree four (or more).
To see this, consider a flat torus $T$ of square fundamental domain $D$.
Denote by $x$ the point corresponding to the vertices of $D$, and consider $CLNS(x)$.
The cyclic part of $C(x)$ consists of two loops at a degree four vertex and $T$ is orientable,
see Figure 1.
\end{ex}
\begin{figure*}
\label{Orient_Cy_deg_4}
\centering
\includegraphics[width=0.5\textwidth]{Orient_Cy_deg_4.eps}
\caption{CLNS on a flat torus of square fundamental domain.}
\end{figure*}
Theorem \ref{crit} can alternatively be stated as follows.
\begin{thm}
\label{or_cy}
The surface $S_G$ realizing the connected graph $G$ as a cut locus is orientable if and only if there exists a system of generating cycles for $G^{cp}$, each of which has an orientable patch in $S_G$.
\end{thm}
\noindent{\sl Proof:}
If there exist an orientable surface $S_G$ realizing $G$ as a cut locus then
every cycle of $G$ has an orientable patch in $S_G$, by Theorem \ref{crit}.
Conversely, assume there exists a system of generating cycles for $G^{cp}$, each of which has an orientable patch in the surface $S_G$ realizing $G$ as a cut locus. Then, by Lemma \ref{prod_or_cy}, each cycle in $S_G$ is orientable.
\hfill $\Box$
\begin{ex}
\label{Non_orient}
By the use of Theorems \ref{crit} and \ref{or_cy} one can easily see that
the CL-structure in Figure 2 (a) is orientable, while the one in Figure 2 (b) is not.
To extend this example to a $3$-graph with $2g$ generating cycles,
simply connect by an edge-strip the non-orientable CL-structure in Figure 2 (b)
to the orientable CL-structure of the graph with $2g-2$ in Figure 2 (c), see Figure 2 (d).
\end{ex}
\begin{figure*}
\label{No_orient}
\centering
\includegraphics[width=0.6\textwidth]{Fig_n_orient.eps}
\caption{Cut locus structures: orientable (a) and (c), and non-orientable (b) and (d).}
\end{figure*}
The following simple remark is called a corollary because of it importance for the next section.
\begin{co}
\label{2v}
Let $S_G$ be an orientable realization of the $3$-graph $G$, reprezen\-ted planary.
Then, for every cycle $C$ of $G$, if $C$ consists of two edges then both of them are switched,
and if $C$ consists of three edges then exactly two of them are switched.
\end{co}
\section{Orientable realizations of small graphs}
\label{S_4cy}
As an application of our previous results, we present in this section all
distinct, orientable cut locus structures on $3$-graphs with four generating cycles;
the others CL-structures realized as cut loci on surfaces of genus $2$
can be obtained from those on $3$-graphs with $4$ generating cycles by contracting edge-strips, see \cite{iv1}.
The following statement can be obtained by straightforward inductive constructions.
\begin{lm}
\label{g43}
There are only $7$ (up to isomorphisms) $3$-graphs with $4$ generating cycles and no loops;
they are listed in the Figure \ref{4cy}.
\end{lm}
\begin{figure*}
\centering
\includegraphics[width=0.8\textwidth]{Fig_4cy.eps}
\caption{All $3$-graphs with $4$ generating cycles and no loops.}
\label{4cy}
\end{figure*}
\begin{thm}
\label{4}
The list of all orientable cut locus structures on $3$-graphs with $4$ generating cycles
is presented in the Figures 4 -- 13.
\end{thm}
\noindent{\sl Proof:}
By Theorem \ref{or_cy} and Lemma \ref{sum_e}, it suffices to label each edge of our graphs with $\bar 0$ and $\bar 1$ in such a manner that for each simple generating cycle $C$, $\oplus_{e \in E(C)} s(e)= \bar 0$.
The proofs for the planar graphs in Figure \ref{4cy} (a)-(f) are similar,
as they all employ the planar representation of any CL-structure on these graphs (see Lemma \ref{cy-sw}).
For such structures, by Corollary \ref{2v}, both edges of any cycle consisting of two edges will be switched,
and precisely two edges of any cycle consisting of three edges will be switched.
The graph $G_a$ in Figure \ref{4cy} (a) is symmetric with respect to (the mid-point of) its bridge.
Both cycles of two edges have their edges switched, while the two cycles of three edges have one edge not-switched.
Since the bridge cannot modify a CL-structure (Definition \ref{diff_CL}),
we obtain the unique result illustrated in Figure \ref{4cy-i}.
\begin{figure*}
\label{4cy-i}
\centering
\includegraphics[width=0.50\textwidth]{Fig_4cy_i.eps}
\caption{Unique orientable cut locus structure for the graph in Figure \ref{4cy} (a).}
\end{figure*}
Assume the graph $G_b$ in Figure \ref{4cy} (b) has at least one orientable CL-structure. Then, all edges of the three cycles of two edges in $G_b$ are labeled $\bar 1$. But these labels contradict the existence of a $G_b$-strip, because they force a $G_b$-patch to have three boundary components, see Figure \ref{4cy-ii}.
\begin{figure*}
\label{4cy-ii}
\centering
\includegraphics[width=0.4\textwidth]{Fig_4cy_ii_imp.eps}
\caption{No orientable cut locus structure for the graph in Figure \ref{4cy} (b).}
\end{figure*}
For the graph $G_c$ in Figure \ref{4cy} (c), we first label by $\bar 1$ all edges of its two cycles of two edges, see Figure \ref{4cy-iii} (a). Figure \ref{4cy-iii} (b)-(c) shows the next step of our labeling.
Starting from Figure \ref{4cy-iii} (b), we have three possibilities to label the remaining edges according to Lemma \ref{sum_e}, shown in Figure \ref{4cy-iii} (d)-(e)-(f).
Starting from Figure \ref{4cy-iii} (c) and taking into account the symmetries of $G_c$, we have another two possibilities to label the remaining edges according, shown in Figure \ref{4cy-iii} (g)-(h).
By Theorem \ref{or_cy} and Lemma \ref{sum_e}, all obtained CL-structures (Figure \ref{4cy-iii} (d)-(h)) are orientable.
\begin{figure*}
\label{4cy-iii}
\centering
\includegraphics[width=1.0\textwidth]{Fig_4cy_iii.eps}
\caption{Five orientable cut locus structures for the graph in Figure \ref{4cy} (c).}
\end{figure*}
For the graph $G_d$ in Figure \ref{4cy} (d), we apply first Corollary \ref{2v} to obtain Figure \ref{4cy-iv} (a). Then we consider the cases in Figure \ref{4cy-iv} (b), (c), and (d), according to the labeling of cycles with three edges.
The case in Figure \ref{4cy-iv} (b) provides a patch which is not a strip, impossible.
For Figure \ref{4cy-iv} (c) we have two subcases to treat, illustrated in Figure \ref{4cy-iv} (e) and (f), and (f) further ramificates to (h) and (i).
Excluding the subcases of (c), (d) produces one more CL-structure, see Figure \ref{4cy-iv} (g).
\begin{figure*}
\label{4cy-iv}
\centering
\includegraphics[width=1.0\textwidth]{Fig_4cy_iv_imp.eps}
\caption{Four orientable cut locus structures for the graph in Figure \ref{4cy} (d).}
\end{figure*}
The symmetries of the graph $G_e$ in Figure \ref{4cy} (e) help to reduce the number of considered subcases.
At the beginning we have to take into account two cases, illustrated in Figure \ref{4cy-v} (a) and (b),
according to the edges of the first cycle of three edges.
Labeling the second cycle of three edges in Figure \ref{4cy-v} (a) produces the subcases (c), (d), and (e),
of which (c) is not a strip, while (d) and (e) provide each two CL-structures (Figure \ref{4cy-v} (g)-(h) and (i)-(j)).
Excluding the previously treated subcase (e), from (b) we obtain (f) and further (k) and (l), both of which are not strips.
\begin{figure*}
\label{4cy-v}
\centering
\includegraphics[width=0.62\textwidth]{Fig_4cy_v_imp.eps}
\caption{Four orientable cut locus structures for the graph in Figure \ref{4cy} (e).}
\end{figure*}
The graph $G_f$ in Figure \ref{4cy} (f) has two cycles of two edges, see Figure \ref{4cy-vi} (a).
There we have to treat the two subcases (b) and (c), (b) yielding (d) and (e), and (c) yielding (f).
Each of (d), (e), (f) produces two CL-structures, see Figure \ref{4cy} (g)-(h), (i)-(j), and (k)-(l).
\begin{figure*}
\label{4cy-vi}
\centering
\includegraphics[width=1.0\textwidth]{Fig_4cy_vi.eps}
\caption{Six orientable cut locus structures for the graph in Figure \ref{4cy} (f).}
\end{figure*}
The graph $G_g$ in Figure \ref{4cy} (g) is not planar, and the analysis is a little more complicated.
Lebeling the four edge cycle $abcd$ on the outer circle in Figure \ref{4cy-vii_1} (a) produces
five cases. Two of them, (c) and (f) in Figure \ref{4cy-vii_1}, are identical up to the central symmetry of $G_g$ with respect to the mid-point of the edge $ef$, and one other ((e) in Figure \ref{4cy-vii_1}) more
than one boundary component, impossible, hence there remain only (b), (c) and (d) in Figure \ref{4cy-vii_1}.
\begin{figure*}
\label{4cy-vii_1}
\centering
\includegraphics[width=0.84\textwidth]{Fig_4cy_vii_1.eps}
\caption{Finding orientable cut locus structures for the graph in Figure \ref{4cy} (g); $3$ cases to further consider: (b), (c), and (d).}
\end{figure*}
The subcases obtained from the case (b) are illustrated in Figure \ref{4cy-vii_2},
all of them leading to several boundary components.
\begin{figure*}
\label{4cy-vii_2}
\centering
\includegraphics[width=0.95\textwidth]{Fig_4cy_vii_2.eps}
\caption{No orientable cut locus structure for Figure \ref{4cy-vii_1} (b).}
\end{figure*}
The subcases obtained from the case (c) are illustrated in Figure \ref{4cy-vii_3},
the first two of them, (m) and (n), leading to several boundary components. The only orientable CL-structure
obtained in this case is represented in Figure \ref{4cy-vii_3} (o).
\begin{figure*}
\label{4cy-vii_3}
\centering
\includegraphics[width=1.0\textwidth]{Fig_4cy_vii_3.eps}
\caption{Two orientable cut locus structures for Figure \ref{4cy-vii_1} (c).}
\end{figure*}
The subcases obtained from the case (d) are illustrated in Figure \ref{4cy-vii_4},
the first two of them, (r) and (s), leading to several boundary components.
The subcase (t) also yields a contradiction, producing either several boundary components,
or a non-orientable cycle-patch, according to the labeling of the edge $de$.
\hfill $\Box$
\begin{figure*}
\label{4cy-vii_4}
\centering
\includegraphics[width=1.0\textwidth]{Fig_4cy_vii_4.eps}
\caption{No orientable cut locus structure for Figure \ref{4cy-vii_1} (d).}
\end{figure*}
\bigskip
Consider a CL-structure ${\cal C}$ on the graph $G$, a riemannian surface $(S,g)$ and a point $x \in S$.
${\cal C}$ is called {\sl stable} with respect to $x \in S$ if (i) $CLNS(x)={\cal C}$, and (ii) there exists a neighbourhood of $x$ in $S$, for all points $y$ of which holds $CLNS(y)={\cal C}$.
The CL-structure ${\cal C}$ is called {\sl stable} if it is stable on all surfaces where it can be realized as a CLNS \cite{iv2}.
We proved in \cite{iv2} that a CL-structure on the graph $G$ is stable if and only if $G$ is cubic; this and Theorem \ref{4} directly imply the following.
\begin{co}
Up to graph homeomorphisms and CL-structures equi\-va\-lence, there are 22 stable and orientable CL-structures on surfacees of genus $2$.
\end{co}
\section{Orientably realizable graphs}
\label{or_graphs}
We noticed that the orientability of a surface $S_G$ realizing the graph $G$ as a cut locus
is implied not by the properties of $G$, but by those of the cut locus structure of $G$, see Figure 2.
Nevertheless, some graphs have no orientable realization, while others have several, see Section \ref{S_4cy}.
We present next some classes of such graphs.
\begin{df}
A graph is called {\sl orientably realizable} if there exists at least one orientable riemannian surface
realizing it as a cut locus.
\end{df}
A first --obvious-- obstruction for a graph being orientably realizable is provided by the odd number of generating cycles.
Another obstruction is the existence of loops at degree three vertices, see Corollary \ref{loop_at_3}.
Theorem \ref{4} shows, in particular, that these two obstructuctions are not sufficient to characterize
the orientably realizable graphs, see Figure 5.
In this section we give a few criteria for orientability.
\begin{co}
\label{tree}
Assume there exist orientably realizable subgraphs $G_1$,..., $G_m$ in the graph $G$ whose union is $G$,
such that the intersection of any two of them has at most one point.
If the induced incidence graph of $\{ G_1,..., G_m \}$ is a tree then $G$ is orientably realizable.
\end{co}
\noindent{\sl Proof:} Theorem \ref{or_cy} and induction over $m$.
\hfill $\Box$
\begin{ex}
Corollary \ref{tree} is not necessarily true if the induced incidence graph of $\{ G_1,..., G_m \}$ in $G$ is not a tree,
as the example of a triangle ($m=3$, each $G_i$ coincides with an edge) shows.
\end{ex}
\begin{co}
\label{conn}
Assume there exist orientably realizable disjoint subgraphs $G_1$, $G_2$ of the graph $G$ such that
$G \setminus (G_1 \cup G_2)$ is the disjoint union of $k$ edges, each of which connects $G_1$ to $G_2$.
If $k \in \{1, 3\}$ then $G$ is orientably realizable.
\end{co}
\noindent{\sl Proof:}
The case $k=1$ follows easily from Corollary \ref{tree}, by considering the connecting edge as a third graph.
For $k=3$, put $\{ e_1, e_2, e_3 \}=G \setminus (G_1 \cup G_2)$.
We may assume that the edges $e_1, e_2, e_3$ determine two generating cycles of $G$, say $C$ containing $e_1, e_2$ and $C'$ containing $e_2, e_3$.
Denote by $\varepsilon$ the (mod 2) sum of the switched edges of $C$ in the $G_1$-strip $P_{G_1}$ and the $G_2$-strip $P_{G_2}$; define similarly $\varepsilon'$ for $C'$.
There are three cases to analize.
{\sl i)} If $\varepsilon = \varepsilon' =0$ then join $P_{G_1}$ to $P_{G_2}$ by $3$ switched edge-strips.
{\sl ii)} If $\varepsilon =0$ and $\varepsilon' =1$ (or vice versa) then join $P_{G_1}$ to $P_{G_2}$ by switched $e_1$ and $e_2$-strips and a non-switched $e_3$-strip.
{\sl iii)} If $\varepsilon = \varepsilon' =1$ then join $P_{G_1}$ to $P_{G_2}$ by switched $e_1$- and $e_3$-strips and a non-switched $e_2$-strip.
It is straightforward to check that the result is, in all cases, a $G$-strip,
which is orientable by Theorem \ref{or_cy}.
\hfill $\Box$
\begin{ex}
Corollary \ref{conn} is not necessarily true if the subgraphs $G_1, G_2$ are not orientably realizable.
To see this, consider the graph $G$ composed of two cycles $G_1, G_2$ joined by an edge ($k=1$);
by Lemma \ref{cy-sw} and Corollary \ref{2v}, $G$ is not orientably realizable.
\end{ex}
\begin{ex}
Corollary \ref{conn} is not necessarily true for $k=2$.
With the notation in its proof,
if $\varepsilon = \varepsilon' =0$ then, imposing to have an orientable cycle along (the edges corresponding to) $\varepsilon$, $e_1$, (the edges corresponding to) $\varepsilon'$, and $e_2$,
we actually get a $G$-patch that is not a strip, see Figure \ref{k=2}.
\end{ex}
\begin{figure*}
\centering
\includegraphics[width=0.95\textwidth]{Fig_k=2.eps}
\caption{Corollary \ref{conn} is not true for $k=2$.}
\label{k=2}
\end{figure*}
\begin{co}
\label{2_trees}
Assume there exist disjoint subtrees $G_1$, $G_2$ of the graph $G$ such that
$G \setminus (G_1 \cup G_2)$ is the disjoint union of $k$ edges, each of which connects $G_1$ to $G_2$.
If $k$ is odd then $G$ is orientably realizable.
\end{co}
\noindent{\sl Proof:}
The result follows from Theorem \ref{or_cy} and the existence of a $G$-strip obtained by joining the $G_1$-strip to the $G_2$-strip by $k$ switched edge-strips.
\hfill $\Box$
\begin{co}
\label{Petersen}
The Petersen graph is orientably realizable.
\end{co}
\noindent{\sl Proof:}
Deleting the three edges starting from a vertex of the Petersen graph,
we obtain two components: a point and (a graph isomorphic to) the graph in Figure \ref{4cy} (g).
The orientable realizability for the point is trivial, while for the second component two orientable realisations are provided by Theorem \ref{4} (see Figure 12 (m) and (o)). The conclusion follows now by Corollary \ref{conn}.
\hfill $\Box$
\begin{co}
\label{N-or_crit-2}
Assume the graph $G$ can be represented as the union of two disjoint subgraphs, joined by a bridge.
If one of the subgraphs is not orientably realizable
--in particular if it is the graph in Figure 5-- then neither is $G$.
\end{co}
\noindent {\bf Acknowledgement } C. V\^\i lcu was partially supported by the grant PN II Idei 1187 of the Romanian Government.
|
\section{Introduction}
In recent years it became clear that understanding the orbital motion of partons, and in
particular the role of partonic initial and final state interactions of quarks, is crucial
for the construction of a more complete picture of the nucleon in terms of elementary quarks
and gluons.
Parton distribution functions have been generalized to contain information not only on the
longitudinal momentum but also on the transverse momentum distributions of partons in a fast moving hadron.
Intense investigation of Transverse Momentum Dependent (TMD) distributions of partons both from the
experimental and theoretical sides, indicate that QCD-dynamics inside hadrons is
much richer than what can be learned from collinear parton distributions.
Two fundamental mechanisms have been identified leading to single-spin asymmetries
(SSAs) in hard processes:
the Sivers mechanism \cite{Sivers:1990fh,Anselmino:1998yz,Brodsky:2002cx,Collins:2002kn,Ji:2002aa}, which generates
an azimuthal asymmetry in the distribution of quarks in the nucleon due to their orbital motion,
and the Collins
mechanism \cite{Collins:2002kn,Mulders:1995dh}, which generates
an asymmetry during the hadronization of quarks.
TMD distributions also contain unique information on the role of initial and final state
interactions of active partons in hard scattering processes~\cite{Brodsky:2002cx,Collins:2002kn,Ji:2002aa}.
Semi-inclusive deep-inelastic scattering (SIDIS) has emerged as a powerful tool to probe
nucleon structure and provide direct access to TMDs through measurements of spin and azimuthal asymmetries.
In particular, SSAs, arising from correlations of the transverse parton momentum and
the transverse spin of the parton or from the initial or final state hadron, provide unprecedented information
about spin-orbit correlations.
QCD factorization for
semi-inclusive deep-inelastic scattering at low transverse momentum in the
current-fragmentation region has been established in references~\cite{Ji:2004wu,
Collins:2004nx, Bacchetta:2008xw}. This new framework provides a rigorous basis to study the
TMD parton distributions from SIDIS data using different spin-dependent and
independent observables. The analyses of the TMDs strongly depends also on knowledge of the fragmentation functions ~\cite{deFlorian:2009vb,Amrath:2005gv,Bacchetta:2007wc,Matevosyan:2010hh}.
Many experiments are currently trying to pin down various
TMD effects through 1) semi-inclusive deep-inelastic scattering (
HERMES at DESY~\cite{Airapetian:1999tv,Airapetian:2001eg,Airapetian:2004tw,Airapetian:2006rx},
COMPASS
at CERN~\cite{Alexakhin:2005iw}, CLAS and Hall-A at Jefferson
Lab~\cite{Avakian:2003pk,Avakian:2005ps,Avakian:2010ae}),
2) polarized proton-proton collisions (PHENIX, STAR and BRAHMS at RHIC
\cite{Adams:2003fx,Chiu:2007zy,Arsene:2008mi})
and
3) electron-positron annihilation (Belle at KEK~\cite{Abe:2005zx}).
The cross section for single pion production
by longitudinally polarized electrons
scatteried from unpolarized protons may be written in terms of a
set of structure functions
\cite{Mulders:1995dh,Levelt:1994np,Bacchetta:2006tn}.
The helicity
dependent part ($\sigma_{LU}$)
arises from the anti-symmetric part of the hadronic tensor:
\begin{eqnarray}
\label{HLT}
\frac{d\sigma_{LU}}{dx_B dy,dz d^2P_T d\phi_h} \propto \lambda_e \times \sqrt{y^2+\gamma^2} \sqrt{1-y-\frac{1}{4}\gamma^2} \sin \phi_h\,\,{F^{\sin \phi_h}_{LU}}.
\end{eqnarray}
\noindent The subscripts on $\sigma_{LU}$ specify the beam and target
polarizations, respectively ($L$ stands for longitudinally
polarized and $U$ for unpolarized).
The azimuthal angle $\phi_h$ is the angle between leptonic and hadronic
plane according to the Trento convention \cite{Bacchetta:2004jz}.
The kinematic variables $x_B$, $y$, and $z$ are defined as:
$
x_B = Q^2/{2(P_1q)}, \,\, y={(P_1q)/(P_1k_1)}, \,\, z={(P_1P)/(P_1q)},
$
where $Q^2=-q^2=-(k_1-k_2)^2$ is the four-momentum
of the virtual photon, $k_1$ ($k_2$) is the four-momentum of the incoming (scattered) lepton,
$P_1$ and $P$ are the four momenta of the target nucleon and the observed final-state
hadron (respectively), $\gamma^2=4M^2x_B^2y^2/Q^2$, $M$ is the nucleon mass, and $\lambda_e$ is the electron beam helicity.
The structure function $F^{\sin \phi_h}_{LU}$ arises due to the
interference of the longitudinal and transverse photon contributions.
In the partonic description
of SIDIS,
assuming factorization holds, contributions to structure functions can be written
as convolutions of parton distribution and fragmentation
functions dependent on the scaling variables $x_B$ and $z$,
respectively~\cite{Mulders:1995dh,Ji:2004wu}.
The beam-spin asymmetries
in single-pion production off an unpolarized target are higher-twist by nature
\cite{Afanasev:2003ze,Yuan:2003gu}.
Higher-twist observables are important for understanding the
long-range quark-gluon dynamics and
may also be accessible as leading contributions through the measurements of
certain asymmetries
\cite{Levelt:1994np,Jaffe:1991ra,Tangerman:1994eh,Kotzinian:1994dv},
in particular the beam SSAs.
Recently, higher-twist effects in SIDIS were interpreted in terms of
average transverse forces acting on the active quark at the instant after
being struck by the virtual photon \cite{Burkardt:2008vd}.
Different contributions to
the beam SSAs discussed so far provide information on leading and
sub-leading parton
distribution and fragmentation functions, related both to Collins and
Sivers production
mechanisms.
Sizable beam SSAs were predicted \cite{Yuan:2003gu} based on spin-orbit correlations
as the dynamical origin.
Within this framework, the asymmetry generated at the
distribution level,
is given by either the convolution of the T-odd parton distribution
$h_1^{\perp}$ with the twist-3 fragmentation function $E$
\cite{Jaffe:1991ra}, or the convolution of the twist-3 T-odd distribution
function $g^\perp$ with the unpolarized fragmentation function $D_1$\cite{Metz:2004je}.
\section{The $\pi^0$ beam spin asymmetry }
In this section we present measurements from the E01-113 CLAS dataset of
beam-spin asymmetries in the electroproduction of
neutral pions in deep-inelastic scattering
using the 5.776 GeV
electron beam and the CEBAF Large Acceptance
Spectrometer (CLAS) \cite{Mecking:2003zu} at Jefferson Laboratory.
Longitudinally polarized electrons are scattered off
a liquid-hydrogen target.
The beam polarization, frequently measured with a
M{\o}ller polarimeter, was on average $0.80$
with a fractional systematic uncertainty of 3\%.
The beam helicity was flipped every 30 ms to minimize systematic instrumental effects.
Scattered electrons were detected in CLAS.
Electron candidates were selected by a hardware trigger using a
coincidence between
the gas Cherenkov
counters and the lead-scintillator electromagnetic calorimeters (EC).
Neutral pion events were identified by calculating the invariant mass of two photons
detected with the CLAS electromagnetic calorimeter and the Inner Calorimeter (IC) \cite{girod:2007jq}.
In each kinematic bin, $\pi^0$ events are selected by a gaussian plus linear fit (see figure \ref{pi0bgkfit}).
The combinatorial background is subtracted in each bin from the number of events inside $3 \sigma$ using the
linear component of the fit.
\begin{figure}[h]
\begin{minipage}{18pc}
\includegraphics[width=18pc]{pi0bgkfit.eps}
\caption{\label{pi0bgkfit}Invariant mass spectrum of the $\gamma \gamma$ system $M(\gamma \gamma)$ in an arbitrarly chosen $x_B$, $P_T$, $z$ and $\phi _{h}$-bin, fitted by gausian plus linear polinomial. Vertical black lines indicates $\pm 3\sigma$ from the mean.}
\end{minipage}\hspace{2pc}%
\begin{minipage}{18pc}
\includegraphics[width=18pc]{exFITsin12bin.eps}
\caption{\label{exFITsin12bin}Example of a $p_0 \cdot \sin(\phi_h)$ fit for $0.4<z<0.7$, $0.1<x_B<0.2$ and $0.2$ GeV/c$ <P_T<0.4$ GeV/c. Only statistical error bars are shown.}
\end{minipage}
\end{figure}
Deep-inelastic scattering events are selected by requiring $Q^2>1$ GeV$^2$ and $W^2>4$ GeV$^2$,
where $W$ is the invariant mass of the hadronic final state.
Events with low missing mass of the $e\pi^0$ system ($M_X<1.5$ GeV) were discarded to exclude
contributions from exclusive processes.
The minimum value of the transverse $\pi^0$ momentum, $P_T>0.05$ GeV, ensures that the azimuthal
angle $\phi_h$ is well defined.
The total number of selected electron-$\pi^0$ coincidences
is $\approx 3.0\times 10^6$.
The beam-spin asymmetry has been calculated for each kinematic bin as:
\begin{equation}
A_{LU} (\phi_{h})=\frac{1}{P} \frac{N^{+}_{\pi^0}(\phi_{h}) - N^{-}_{\pi^0}(\phi_{h})}{N^{+}_{\pi^0}(\phi_{h}) + N^{-}_{\pi^0}(\phi_{h}) }
\label{alueq0}
\end{equation}
where $P=0.794 \pm 0.024$ is the absolute beam polarization this data set and $N^{+}_{\pi^0}$ and $ N^{-}_{\pi^0}$ are the number of $\pi^0$'s with positive and negative beam helicity, respectively.
Asymmetry moments are extracted by
fitting the $\phi_h$-distribution of $A_{LU}$ in each $x_B$ and $P_T$ bin with the theoretically motivated
function $p_0 \cdot \sin(\phi_h)$.
An example of this fit is shown in figure~\ref{exFITsin12bin} for an arbitrarly chosen kinematic bin.
\begin{figure*}[ht]
\begin{center}
\includegraphics[width=0.99\textwidth]{ALU_vs_PTXB_sysHPrel.eps}
\end{center}
\caption{$A^{\sin \phi_h}_{LU}$ as function of $P_T$ for different $x_B$ ranges
and integrated over $0.4<z<0.7$. The error bars correspond to statistical and the bands to systematic uncertainties.
An additional 3\% uncertainty arises from the beam polarization measurement and another 3\% uncertainty
from radiative effects.}
\label{ALU_vs_PTXB_sysH}
\end{figure*}
In figure~\ref{ALU_vs_PTXB_sysH}, the
extracted $A_{LU}^{sin\phi}$ moment is presented as a function of $P_T$ for different $x_B$ ranges.
Systematic uncertainties, presented by the bands, include the uncertainties due to the background
subtraction, the event selection and possible contributions of higher harmonics in the extraction of the moments. These contributions are added in quadrature.
An additional 3\% uncertainty due to the beam polarization measurement and another 3\% uncertainty from radiative effects should be added to the systematic uncertainties.
The $A^{\sin \phi_h}_{LU}$ moment increases at low $P_T$ and reaches a plateau at values of about 0.3 GeV.
There is an indication that the decrease of $A^{\sin \phi_h}_{LU}$ at large $P_T$,
expected from perturbative QCD, starts already at $P_T \sim$ 0.6 GeV.
The surprising characteristics of favored and unfavored Collins functions, being of
roughly equal magnitude but having opposite signs, as
indicated by latest measurements at HERMES \cite{Airapetian:1999tv,Airapetian:2004tw},
COMPASS \cite{Alexakhin:2005iw} and Belle \cite{Abe:2005zx},
puts the $\pi^0$ in a unique position in SSA studies.
Since the $\pi^0$ fragmentation function (FF) is the sum of $\pi^+$ and $\pi^-$ FFs, its
favored and unfavored contributions will be roughly equal and, in the case of the Collins FF, will cancel
each other to a large extent.
Contributions to the beam-SSA related to spin-orbit correlations can thus be studied without a significant
background from the Collins mechanism.
The measured beam-spin asymmetry amplitude for $\pi^0$ appears to be comparable with the $\pi ^+$ asymmetry
from a former CLAS data set \cite{HAdubna} both in magnitude and sign, as shown in figure~\ref{CLASaul},
indicating that contributions from the Collins mechanism cannot be the dominant ones.
\begin{figure}
\begin{center}
\includegraphics[width=0.5\textwidth]{comppippi0e16_pt03Prel.eps}
\end{center}
\caption{The $\pi^0$ beam-spin asymmetry amplitude $A^{\sin \phi_h}_{LU}$
as function of $x_B$ compared to that for $\pi^+$ from an earlier CLAS measurement~\cite{HAdubna}.
Uncertainties of the $\pi^0$ measurement are as in figure~\ref{ALU_vs_PTXB_sysH}.
For both data sets $<P_T>\approx 0.38$ and $0.4<z<0.7$.}
\label{CLASaul}
\end{figure}
A similar measurement has been performed by the HERMES collaboration at a higher beam
energy of 27.6 GeV~\cite{Airapetian:2006rx}.
After taking into account the kinematic factors in the expression
of the beam-helicity dependent and unpolarized terms (\cite{Bacchetta:2006tn})
\begin{equation}
f(y)=\frac{y \sqrt{1-y}}{1-y+y^{2}/2} ,
\label{fy}
\end{equation}
CLAS and HERMES measurements are found to
be consistent as shown in figures~\ref{xbCH} and~\ref{ptCH}, indicating that at energies as low as 4-6 GeV the
behavior of beam-spin asymmetries is similar to higher energy measurements. The CLAS data provide
significant improvements in precision of beam SSA measurements in the kinematic region where the two data
sets overlap, and extend the measurements to the large $x_B$ region not accessible by HERMES.
\begin{figure}[h]
\begin{minipage}{18pc}
\includegraphics[width=18pc]{compHERMxBPrel.eps}
\caption{\label{xbCH}$A^{\sin \phi_h}_{LU}$ multiplied by the kinematic factor $<Q>/f(y)$ as a function of $x_B$ from CLAS and HERMES \cite{Airapetian:2006rx}. The $0.4$ GeV/c$ <P_T<0.6$ GeV/c range for the CLAS data was used to compare with HERMES, since this is the closest kinematic range.}
\end{minipage}\hspace{2pc}%
\begin{minipage}{18pc}
\includegraphics[width=18pc]{compHERMPrel.eps}
\caption{\label{ptCH} $A^{\sin \phi_h}_{LU}$ multiplied by the kinematic factor $<Q>/f(y)$ as function of $P_T$ from CLAS and HERMES \cite{Airapetian:2006rx} (the same as in figure~\ref{xbCH}). The $0.1<x_B<0.2$ range for the CLAS data was used to compare with, since this is the closest kinematic range.}
\end{minipage}
\end{figure}
\subsection{Contamination from two-hadron production}
In this section a
study of the beam spin asymmetry of single-pion production from exclusive and non-exclusive vector meson production is presented.
In order to investigate a possible contamination of the presented results
by pions originating from the decay of exclusively produced vector mesons,
as advocated in~\cite{Airapetian:2006rx}, events with at least one $\pi^0$ and $\pi^+$ were selected from the same dataset.
Exclusivity was ensured by requiring the missing mass $M_x$ of the
$ep\rightarrow e'\pi^0 \pi^+ X$ system to be within
0.8 GeV $< M_X(ep\rightarrow e'\pi^0 \pi^+ X)<1.1$ GeV.
The invariant mass distribution of this exclusive two pion sample
is presented in the upper panel of figure~\ref{exrhop} where the exclusive
$\rho^+$ peak is clearly visible along with non-resonant exclusive
two-pion production.
The vertical lines indicate invariant mass ranges for which the
beam spin asymmetry amplitude $A_{LU}^{\sin\phi_h}$ was extracted for each of the
two exclusively produced pions following the extraction method
as described before.
These amplitudes are presented for $\pi^0$ and $\pi^+$ in the second top
panel of figure~\ref{exrhop} as labeled therein,
along with the average values of $<P_T>$ and $<z>$ for the invariant mass
ranges.
Figure~\ref{sirhop} is the same as figure~\ref{exrhop} except for two-hadron events with $M_x(ep\rightarrow e'\pi^0 \pi^+ x)>1.5$ GeV (non exclusive).
\begin{figure}[h]
\begin{minipage}{18pc}
\includegraphics[width=18pc]{EXrhoP_1.eps}
\caption{\label{exrhop} From top to bottom: invariant mass $M_{\pi^0\pi^+}$ of two hadrons, $A^{\sin \phi_h}_{LU}$ , and average $<P_T>$ and $<z>$ as function of $M_{\pi^0\pi^+}$ for $\pi^0$ and $\pi^+$ in exclusive two-pion production.}
\end{minipage}\hspace{2pc}%
\begin{minipage}{18pc}
\includegraphics[width=18pc]{SIDrhoP_1.eps}
\caption{\label{sirhop} From top to bottom: Invariant mass $M_{\pi^0\pi^+}$ of two hadrons, $A^{\sin \phi_h}_{LU}$ , and average $<P_T>$ and $z$ as function of $M_{\pi^0\pi^+}$ for $\pi^0$ and $\pi^+$ in non-exclusive production ($M_x(ep\rightarrow e'\pi^0 \pi^+ x)>1.5$ GeV).}
\end{minipage}
\end{figure}
The $A^{\sin \phi_h}_{LU}$ values for $\pi^0$($\pi^+$) increase (decrease) with increas invariant mass, and reach their maximum absolute value at the highest invariant mass.
In contrast to a similar study for $\rho^0$ presented in \cite{Airapetian:2006rx}, the asymmetres for $\pi^0$ and $\pi^+$ have minimal values in the invariant mass range where $\rho^+$ events dominant, and furthermore the highest bin of invariant mass, the asymmetries for $\pi^0$ and $\pi^+$ have different signs.
It should be noted also, that in these studies the momentum distributions and the acceptance for the two pions are not similar, while in \cite{Airapetian:2006rx} the two pions have similar momentum distributions and acceptance. $A^{\sin \phi_h}_{LU}$ of the SIDIS $\pi^0$ and $\pi^+$ in the same kinematic bin have the same sign and their amplitudes are roughly equal (figure~\ref{CLASaul}). For the case of two-hadron production (figure~\ref{sirhop}), in all invariant mass bins (except the first) $A^{\sin \phi_h}_{LU}$ for $\pi^0$ and $\pi^+$ have different signs, which is another reason for careful studies of two hadron production.
Comparison of the measured beam SSAs for pions originating from electroproduction of two-pions in
exclusive and non-exclusive processes, suggests that the SSAs depend mainly on the kinematics of
the pion, in particular on $z$ and $P_T$, and are comparable for pions from different samples within
the same kinematical bin.
\section{Conclusions}
We have presented measurements of kinematic dependences
of the beam-spin asymmetry in semi-inclusive $\pi ^0$ electroproduction from the E01-113 CLAS dataset.
The $\sin \phi_h$ amplitude is extracted as a function of $x_B$ and the transverse pion
momentum $P_T$, integrating over the $z$-range $0.4<z<0.7$.
The asymmetry shows no significant $x_B$-dependence over the measured range.
The $P_T$ dependence of $A^{\sin \phi_h}_{LU}$ is consistent with an increase at low values of $P_T$
which reaches a plateau for values $P_T >$ 0.3 GeV.
There is an indication that the decrease of $A^{\sin \phi_h}_{LU}$ at large $P_T$, expected from perturbative QCD,
already starts at $P_T \sim$ 0.6 GeV.
The observed asymmetry amplitudes for $\pi ^0$ indicate that the major contribution pion beam SSAs
may be due to spin-orbit correlations.
The results obtained are compared with published HERMES data~\cite{Airapetian:2006rx},
providing significant improvement in precision and an important
input for studies of higher twist effects.
Despite the fact that the partonic formalism is much better suited
for higher energy reactions, there is a reasonable agreement in size and behavior
of beam SSAs measured over a wide energy range.
Preliminary results on two-hadron production show interesting trends: opposite signs of asymmetries for $\pi^0$ and $\pi^+$, the highest asymmetry for the highest invariant mass range. The beam spin asymmetry from $\pi^0$ and $\pi^+$ two-hadron production has been studied for the first time as a function of the invariant mass of two pions.
The $A^{\sin \phi_h}_{LU}$ amplitude of $\pi^0$ from exclusive and non-exclusive two-pion production
exhibits a similar magnitude and similar kinematic dependences as for semi-inclusive $\pi^0$ production. Therefore, a possible contribution of this
exclusive channel to our semi-inclusive result would not alter the presented
asymmetry amplitudes.
\
\
\noindent {\bf Acknowledgement}
\noindent We thank A. Metz and F.Yuan, A. Bacchetta, S. Kuhn, A. Kotzinian, A. Prokudin, L. Gamberg and A. Afanasev for useful and stimulating discussions.
We would like to acknowledge the outstanding efforts of the staff of the
Accelerator and the Physics Divisions at JLab that made this experiment possible.
This work was supported in part by the U.S. Department of Energy
and the National Science Foundation,
the Italian Istituto Nazionale di Fisica Nucleare, the
French Centre National de la Recherche Scientifique,
the French Commissariat \`{a} l'Energie Atomique,
The Southeastern Universities Research Association (SURA) operates the
Thomas Jefferson National Accelerator Facility for the United States
Department of Energy under contract DE-AC05-84ER40150.
\section*{References}
\bibliographystyle{iopart-num}
|
\section{Introduction}
Stars more massive than 8~M$_\odot$ explode as core collapse
supernovae, with kinetic explosion energies of the ejected material
on the order of $10^{51}$~erg.
The remnants, the protoneutron stars (PNSs), are initially hot and
lepton-rich and cool via deleptonization during the first 30~seconds
after the onset of the explosion.
Above a certain progenitor mass threshold on the order of 40~M$_\odot$,
which is an active subject of research, stars will no longer explode.
Such models will collapse and from black holes.
The critical mass for a PNS to collapse and from a black hole
is given by the equation of state (EoS).
The commonly used EoS in core collapse supernova simulations
are based on pure hadronic descriptions,
e.g. the compressible liquid-drop model with surface effects
\cite{Lattimer:1991nc}
and relativistic mean field theory including the Thomas-Fermi
approximation for heavy nuclei \cite{Shen:1998by}.
The conditions that are obtained in PNS interiors during the simulation,
are several times nuclear matter density,
temperatures on the order of several tens of MeV and a low
proton-to-baryon ratio given by the electron fraction
of\footnote{The proton-to-baryon ratio, $Y_p=n_p/n_B$,
is equal to the electron fraction,
$Y_e:=Y_{e^-}-Y_{e^+} $,
in the absence of muons.} $Y_e=Y_p\leq0.3$.
Under such conditions, the quark-hadron phase transition is not unlikely to
take place and the assumption of pure hadronic matter becomes
questionable.
Ab initio calculations of the phase diagram and EoS of quantum
chromodynamics (QCD) as the fundamental theory of strongly interacting
matter come from simulations of this gauge theory on the lattice but are
still restricted to low baryon densities (chemical potentials).
In this region of the phase diagram they predict a crossover transition
with a pseudocritical temperature for chiral symmetry restoration and
deconfinement at $T_c\simeq 150 - 170$ MeV
\cite{Borsanyi:2010bp,Bazavov:2010bx}.
A critical endpoint for first order transitions is conjectured but lies,
if it exists at all, outside the region presently accessible by lattice QCD.
An interesting conjecture supported by a statistical model analysis of
hadron production in heavy-ion collisions and by the large-N$_c$ limit of
QCD suggests the existence of a triple point in the phase diagram
\cite{Andronic:2009gj} due to a third, ``quarkyonic'' phase at temperatures
$T<T_c$ and high baryon densities \cite{McLerran:2007qj}.
This state of matter might become accessible in experiments at the planned
third generation of heavy-ion collision facilities FAIR in Darmstadt
(Germany) and NICA in Dubna (Russia) \cite{Blaschke:2010ka}
which thus allow systematic laboratory studies of conditions like in supernova
collapse and protoneutron star evolution \cite{Klahn:2011au}.
In order to investigate the appearance of quark matter in core collapse
supernova models, the implementation of a quark-hadron hybrid EoS is required.
It must be valid for a large range of densities $n_B$, temperatures $T$
and proton-to-baryon ratios.
At large baryon densities, where lattice QCD cannot be applied due
to current conceptional limitations, phenomenological models are commonly used.
In the present study we will discuss hybrid EoS which employ
quark matter models that are representative examples from
two wide classes: bag models and NJL-type models.
The popular and simple thermodynamical bag model is inspired by the success
of the vacuum MIT bag model for the hadron spectrum \cite{DeTar:1983rw}.
It describes quarks as non-interacting fermions of a constant mass confined
by an external ``bag'' pressure $B$.
NJL-type models are constructed to obey basic symmetries of the QCD Lagrangian
like the chiral symmetry of light quarks, and to describe its dynamical
breaking which results in medium-dependent masses (see \cite{Buballa:2003qv}
and references therein).
The inclusion of diquark interaction channels leads to a rich phase structure
at low temperatures and high densities with color superconductivity (diquark
condensation) in two-flavor (2SC) and three-flavor (CFL) quark matter
(see, e.g., \cite{Blaschke:2005uj} for the phase structure and
\cite{Sandin:2007zr} for the relevance to protoneutron stars;
detailed information concerning color superconductivity is
found in review articles \cite{Buballa:2003qv,Alford:2007xm}).
These models have no confining interaction and would therefore lead to the
unphysical dominance of thermal quark excitations already at temperatures
well below $T_c$. Their coupling to the Polyakov loop potential is essential
to suppress the unphysical degrees of freedom and it can be adjusted
to fit the behavior of lattice QCD thermodynamics at low densities
\cite{Roessner:2006xn}. Extending this effective model to high densities
leads to the class of PNJL models of which we apply one for this study.
Different assumptions made for the description of quark matter
lead to different critical conditions for the onset of deconfinement,
which are given in terms of a critical density that depends on the
temperature and the proton-to-baryon ratio.
The resulting different phase diagrams may lead to a different
evolutionary behavior for core collapse supernovae.
The central supernova conditions start from low densities
($6\times10^{-6}$~fm$^{-3}$/$10^{10}$~g cm$^{-3}$)
and temperatures (0.6~MeV),
where pure hadronic matter dominates,
and approach densities above nuclear saturation density
(0.16~fm$^{-1}$/$2.7\times10^{14}$~g cm$^{-3}$)
at temperatures on the order to tens of MeV and $Y_p\simeq0.2-0.3$.
The nature of the QCD transition is not precisely understood on a microscopic
level. Therefore, one usually splices independent nuclear and quark matter
EoS by constructing the phase transition using more or less appropriate
conditions for the phase equilibrium.
Popular examples are the Maxwell, Gibbs or Glendenning
\cite{Glendenning:1992vb} constructions, whereby usually the Maxwell
construction leads to the smallest region between critical densities for the
onset and the end of the mixed phase.
We will disregard finite size effects (pasta structures) and also
non-equilibrium effects due to the nucleation of the new phase with the
justification that weak processes establishing the chemical equilibrium
are fast compared to the typical timescales encountered in supernova
simulations \cite{Mintz:2009ay}.
We will compare the evolution of a representative core collapse supernova
simulation in the two different phase diagrams based on the bag model
and the PNJL model.
The core collapse supernova model is based on general relativistic
radiation hydrodynamics and three flavor Boltzmann neutrino
transport
\cite{Mezzacappa:1993gm,Mezzacappa:1993gn,Mezzacappa:1993,Liebendoerfer:2000cq,Liebendoerfer:2000fw,Liebendoerfer:2002xn}
The manuscript is organized as follows: In Sect. 2 we introduce the standard
core collapse supernova phenomenology.
In Sect. 3 we briefly introduce the two quark-hadron hybrid EoS,
the quark bag model and the PNJL model.
The evolution of a representative 15~M$_\odot$ core collapse supernova
model in the phase diagram is illustrated in Sect. 4 by comparing the quark bag
and PNJL models.
We close with a summary in Sect. 5.
\section{The standard scenario of core collapse supernovae}
Si-burning produces Fe-cores at the final phase of stellar evolution
of massive stars.
These Fe-cores start to contract due to the photodisintegration of heavy
elements and electron captures.
The latter reduces the dominant pressure of the degenerate electron gas.
During the collapse, density and temperature rise and hence
electron captures, which deleptonize the central core, increase.
The collapse accelerates until neutrino trapping densities,
on the order of $\rho\simeq10^{11}$--$10^{13}$~g/cm$^3$,
are obtained after which the collapse proceeds adiabatically.
At nuclear matter density, the repulsive nuclear interaction stiffens
the EoS significantly and the collapse halts.
The core bounces back, where a sound wave forms,
which steepens into a shock wave.
The shock wave propagates outwards where the dissociation of infalling
heavy nuclei causes an energy loss of about 8~MeV per baryon.
Furthermore, during the shock propagation across the neutrinospheres,
which are the neutrino energy and flavor dependent spheres of last scattering,
additional electron captures release a burst of $\nu_e$
that carries away 4--$5\times10^{53}$~erg/s
on a short timescale of 5--20~ms post bounce
(depending on the progenitor model and the EoS).
Both sources of energy loss turn the dynamic shock into a
standing accretion shock, already at about 5~ms post bounce.
The post bounce evolution is determined by mass accretion, due to the
continuously infalling material from the outer layers of the Fe-core as
well as the surrounding Si-layer (depending on the progenitor model).
On a timescale on the order of 100~ms up to seconds,
the central density and temperature increase continuously.
In order to achieve an explosion, energy needs to be deposit
behind the standing accretion shock which subsequently
revives the standing accretion shock.
Several mechanisms have been suggested, including
the magnetically-driven
\cite{LeBlanc:1970kg,Moiseenko:2007,Takiwaki:2009},
the acoustic \cite{Burrows:2006}
and the neutrino-driven \cite{Bethe:1984ux}.
The standard scenario, delayed explosions due to neutrino heating,
have been shown to work in spherical symmetry
\cite{Kitaura:2005bt,Fischer:2009af}
for the low mass 8.8~M$_\odot$ ONeMg-core
\cite{Nomoto:1983,Nomoto:1984,Nomoto:1987}.
For more massive progenitors, multi-dimensional phenomena such as
rotation and the development of fluid instabilities are required and
help to increase the neutrino heating efficiency
\cite{Miller:1993,Herant:1994dd,Burrows:1995ww,Janka:1996}.
Such models are also required to aid the understanding
of aspherical explosions
\cite{Bruenn:2009,Marek:2009}.
\section{The quark bag and PNJL hybrid models}
We discuss two different quark matter descriptions for use in astrophysical
applications as well as their differences in the resulting EoS.
The quark bag model for three flavor quark matter is described in detail
in the Refs.~\cite{Sagert:2008ka, Fischer:2010zzb, Fischer:2010wp}.
Bag constants in the range of $B^{1/4}=$155--165~MeV and
corrections from the strong interaction were adopted
and a fixed strange quark mass of 100~MeV is applied
in these simulations which we want to contrast here with first results
for a PNJL model.
A serious drawback of the bag model parametrization employed here is that
cannot describe the mass of the recently observed 1.97~M$_\odot$
neutron star \cite{Demorest:2010bx}.
This could be cured by extending the bag model and accounting for leading order
QCD corrections and diquark condensation \cite{Ozel:2010bz,Weissenborn:2011qu}
but results in an early onset of quark matter only
for considerably stiff nuclear EoS.
The three flavor quark matter PNJL model is based on the description in
Ref.~\cite{Blaschke:2005uj,Sandin:2007zr} with the Polyakov loop extension
according to \cite{Roessner:2006xn}. An additional isoscalar vector meson
interaction leads to a stiffening of the quark matter equation of state and
allows to describe hybrid stars with a mass of 2~M$_\odot$ \cite{Klahn:2006iw}.
The diquark and vector meson coupling is set to $\eta_D=1.02$ and
$\eta_V=0.25$, respectively.
Strangeness on the quark side in the PNJL model occurs at higher densities
than the onset of deconfinement, due to the dynamical quark masses involved.
The resulting phase diagram is discussed in Ref.~\cite{Blaschke:2010ka}.
The Figs.~\ref{fig-phasediagram-bag} and \ref{fig-phasediagram-pnjl}
compare the phase diagrams for the quark bag model and the PNJL
model for different proton-to-baryon ratios $Y_p$ relevant for
supernova matter.
There are
different critical densities for the onset of deconfinement
comparing the quark bag and PNJL models for equal $Y_p$.
E.g. for $Y_p=0.3$ and $T=0$, the critical densities are
0.159~fm$^{-1}$ for the bag model in comparison to
0.214~fm$^{-1}$
for the PNJL model
(see the solid lines in the
Figs.~\ref{fig-phasediagram-bag} and \ref{fig-phasediagram-pnjl}).
For the bag model, the early onset of deconfinement close to
normal nuclear matter density
illustrates the strong isospin dependency of the critical density.
For instance, at higher $Y_p=0.5$ the critical density shifts to
0.321~fm$^{-1}$ for $T=0$.
In general, small transition densities result from the application
of the Gibbs conditions for the phase transition with the corresponding
extended mixed phase and from the existence of the s-quark flavor which
results in a larger number of degrees of freedom when compared to nuclear
matter.
For the PNJL model, the transition to quark matter,
shown in Fig.~\ref{fig-phasediagram-pnjl},
is based on the Maxwell construction.
The critical densities are generally larger and the mixed phase is very
narrow in density compared to the quark bag model.
The PNJL model also shows the typical isospin dependency.
The critical density reduces at decreasing $Y_p$, which is however
much weaker than for the quark bag model.
Both models differ also significantly in the maximum mass of neutron
stars, 1.50~M$_\odot$ for the bag model with $B^{1/4}=165$~MeV
and 1.97~M$_\odot$ for the PNJL model.
Furthermore, the temperature dependence of the critical density
is different in both quark matter descriptions.
This is of particular relevance for astrophysical applications that
explore a possible quark-hadron phase transition.
In the quark bag model, the critical density reduces continuously at
higher temperatures.
It reaches about 0.1~fm$^{-1}$ at $T=50$~MeV for $Y_p=0.3$.
For the PNJL model, the critical density rises with increasing temperature
for any $Y_p$.
This behavior is a result of the neglect of any modification of the
Polyakov loop potential on the quark density and should be improved
as soon as reasonable constraints for such a procedure can be defined,
see \cite{Blaschke:2010vj} for a first step.
The resulting EoS are illustrated in Figs.~\ref{fig-pressure-bag}
and \ref{fig-pressure-pnjl}, showing the pressure-density curves
for different entropies per baryon and fixed $Y_p=0.3$.
The largely extended mixed phase region in the phase diagram
for the bag model corresponds to a soft EoS.
This is illustrated via the pressure-density curves in
Fig.~\ref{fig-pressure-bag},
where the adiabatic index differs initially only little between
the onset of deconfinement and the pure hadronic phase.
Towards the end of the mixed phase where the pure quark phase
sets in, the adiabatic index is largely reduced for any entropy per
baryon and hence the EoS is significantly softer compared to the
hadronic case.
In the pure quark phase, the adiabatic index increases again
significantly, which makes the pure quark phase stiffer than the
mixed phase.
The sharp transition between the mixed and the pure quark phases,
gives rise to a strong effect for the hydrodynamics evolution
along isentropes.
The PNJL hybrid EoS shows a different pressure behavior
(see Fig.~\ref{fig-pressure-pnjl}) during the quark-hadron phase
transition.
The largest change of the adiabatic index is found at the onset
of deconfinement.
Close to the onset of pure quark matter, the adiabatic index
changes only little.
These two aspects are a general feature of the chosen construction
of the PNJL hybrid EoS, for which a smooth transition can be expected
in dynamical simulations.
\section{Core collapse supernova evolution in the phase diagram}
The central temperature and density evolution of the 15~M$_\odot$
core collapse supernova simulation is shown via the dashed lines in
the Figs.~\ref{fig-phasediagram-bag} and \ref{fig-phasediagram-pnjl},
for the first second post bounce, during which the explosion
is expected to take place for such massive Fe-core progenitors
\cite{Marek:2009}.
This phase of the supernova is determined by mass accretion
($\sim0.1$~M$_\odot$/s) onto the central PNS, which consequently
contracts on a timescale between 100~ms up to seconds.
During the PNS contraction, the central density and temperature
rise continuously\footnote{The highest temperatures $\sim30-60$~MeV
are not obtained at the center of the PNS but at slightly lower densities;
it corresponds to the region where the bounce shock formed initially.}.
During this evolution, the central electron fraction reduced from
$Y_p\simeq0.3$ to $Y_p\simeq0.25$.
The simulation is based on a pure hadronic description of matter
\cite{Shen:1998by}.
The current supernova models explore only hadronic EoS.
It is shown here to illustrate the possibility and the conditions
relevant in order to obtain quark matter at supernova cores.
The differences between the two hybrid EoS become clear.
Using the quark bag model, it is possible to obtain quark matter.
The early post bounce evolution of the central mass trajectories considered
for the 15~M$_\odot$ progenitor model, enter deeply into the mixed phase.
The same holds true for the non-central part of the PNS, where
up to 0.5~M$_\odot$ can reach the mixed phase within the
first 500~ms post bounce evolution
(depending on the progenitor model \cite{Fischer:2010wp}).
Within the PNJL model, the central densities obtained for this particular
15~M$_\odot$ core collapse supernova simulation are not sufficiently
high enough to enter the mixed phase.
It would require an additional rise of the central density by a factor of
two or more.
For such quark-hadron hybrid EoS quark matter will not be reached
in core collapse supernovae of massive stars up to 15~M$_\odot$
as discussed here, during the expected explosion phase of
0.5--1~seconds post bounce.
In order to investigate the EoS softening effects from the presence of
quark matter in the PNS interior and possible consequences for the
dynamical evolution (and possible observations), we apply the quark
bag model hybrid EoS (using $B^{1/4}=155$~MeV, $\alpha_S=0.3$)
to a core collapse supernova evolution (again the 15~M$_\odot$ model).
The largely reduced adiabatic index at the end of the mixed phase
causes a gravitational collapse of the PNS, spatially separated into
central sub-sonic collapse and outer super-sonic collapse.
During the collapse density and temperature rise
(see Fig.~\ref{fig-hydro-h15y}),
which in turn favors pure quark matter over hadronic matter.
In the pure quark phase, where the adiabatic index is increased,
the collapse halts and a strong hydrodynamic shock front forms.
This was first observed in the context of a first order deconfinement
phase transition in ref.~\cite{Takahara:1988}, in relation to the statistically
insignificant multi-peaked neutrino signal form SN1987A \cite{Hirata:1988},
and later discussed in more detail in the refs.~\cite{Gentile:1993, Grigorian:1999}.
The shock wave appears initially as a pure accretion front, which is
illustrated in Fig.~\ref{fig-hydro-h15y} at 310.4668~ms post bounce.
The shock propagates outwards in radius towards the
PNS surface, driven by the thermal pressure of the deconfined
quarks and neutrino heating behind the shock.
The latter aspect is related to local heating rates due to electron
(anti)neutrino absorptions, which increase at the quark-hadron
phase boundary where the infalling hadronic material converts into
quark matter and density and temperature rise by several orders of
magnitude.
At the PNS surface, where the density decreases over several orders
of magnitude, the accretion front accelerates and positive velocities
are obtained
(see Fig.~\ref{fig-hydro-h15y} at 310.5135~ms post bounce).
This moment determines the onset of explosion, even in core collapse
supernova models where otherwise (i.e. without the quark-hadron phase
transition) no explosions could have been obtained.
The shock expands and continues to accelerate where matter velocities
on the order of $10^{5}$~km/s are obtained.
The shock propagation across the neutrinospheres releases an additional
burst of neutrinos, illustrated in Fig.~\ref{fig-lumin}.
This second burst rises in all neutrino flavors.
It differs from the reference case discussed in ref.~\cite{Fischer:2010wp},
where the second burst was dominated by $\bar{\nu}_e$ and
($\nu_{\mu/\tau}$, $\bar{\nu}_{\mu/\tau}$) due to the lower density
of the PNS envelope of the less massive progenitor chosen and
the different quark bag EoS parameters.
The rise of the burst is due to the presence of numerous electron-positron
pairs that allow for the production of $\nu_e$ and $\bar{\nu}_e$ via the
charged current reactions as well as
($\nu_{\mu/\tau}$, $\bar{\nu}_{\mu/\tau}$) via pair processes.
It differs from the deleptonization burst at core bounce, which is only
in $\nu_e$ emitted via a large number of electron captures.
This second burst, if occurring, is a strong observational indication for
the reconfiguration of the high density domain in core collapse supernova
evolution.
Coming back to the PNJL model in core collapse supernova simulations,
preliminary results of long-term non-exploding models in the progenitor
mass range between 15--25~M$_\odot$,
show a smooth transition to quark matter at the PNS interior
where only the mixed phase could be reached.
Very little mass ($\sim0.05$~M$_\odot$) is converted into the mixed phase
and the quark fraction rises only on timescales on the order of seconds,
which corresponds to the PNS contraction timescale induced via
mass accretion.
However, massive star explosions are expected to take place during
the first 0.5--1~seconds post bounce.
Hence, a direct correlation to the explosion mechanism of massive
stars as well as direct observables that relate to the explosion phase,
cannot be expected for such a hybrid EoS.
However, this hybrid EoS model has another interesting feature.
Since the PNJL model possesses a critical point, one may study whether in
the collapse of very massive progenitors with subsequent black hole formation
the trajectory of the collapse in the QCD phase diagram sweeps the critical
point before horizon crossing \cite{Ohnishi:2011jv}.
In that case a characteristic modification of possible observables
from the deconfinement transition is expected,
at variance to trajectories which do not pass the critical point.
The realistic modeling of such processes requires further development of a
phase transition construction which does not remove the critical endpoint.
\section{Summary}
We discussed the possibility to reach the critical conditions for the onset
of deconfinement in core collapse supernovae, comparing two different
quark-hadron hybrid EoS.
Based on the bag model for strange quark matter and choices of low
bag constants, without violating constraints from e.g. heavy-ion collision
experiments, the low critical densities as well as the strong isospin and
temperature dependencies allow for the quark-hadron
phase transition to take place in core collapse supernova
simulations within the first 500~ms post bounce.
For a particular choice of parameters,
we illustrated the dynamical evolution.
It was determined by an adiabatic collapse due to the soft EoS in the
extended mixed phase modeled by applying Gibbs conditions,
and the formation of a strong hydrodynamic shock wave.
It triggered the explosion even in models where otherwise
explosions could not be obtained.
Furthermore, the millisecond neutrino burst released from the
shock propagation across the neutrinospheres may become
observable for a future Galactic event if quark matter occurs
\cite{Dasgupta:2009yj}.
The future observation of such a multi-peaked neutrino signal
may reveal information about the nuclear EoS at high densities and
temperatures that cannot be reached in heavy-ion collision experiments
at present.
For the PNJL model parametrization presented here, the higher critical
density and the narrow mixed phase due to the Maxwell construction,
together with the rising critical density for increasing temperatures,
prevent the quark-hadron phase transition to occur in core collapse
supernova simulations during the post bounce accretion phase.
For such an EoS, a direct correlation between the phase transition to
quark matter and the explosion mechanism cannot be expected.
However, the conditions for the appearance of quark matter
for the PNJL model may allow to obtain a signal from core collapse
supernovae during the PNS deleptonization phase, on timescales
on the order of several seconds after the onset of, e.g., initially
neutrino-driven explosions.
The two classes of hybrid EoS discussed here allow for very different types
of investigations in simulations of core collapse supernovae.
Hybrid EoS based on the class of bag models for quark matter give
access to bulk properties of quark-hadron matter and allow to
model EoS which undergo a quark-hadron phase transition
during the supernova collapse accompanied by a second neutrino burst.
However, the bag model cannot capture the aspect of chiral symmetry
restoration and possible effects of the existence of a critical point.
Moreover, it remains to be shown whether the 2~M$_\odot$ mass constraint
can be explained in agreement with the deconfinement scenario for
this class of models.
The hybrid models based on a PNJL-type approach to quark matter would
allow to study the question of the existence of critical points in the QCD
phase diagram, e.g., during black hole formation where high temperatures
on the order of several 100~MeV are obtained.
A systematic study of possible observables from black hole formation
using PNJL models that include a critical endpoint,
may provide the astrophysical analogue of the energy scan programs
in heavy-ion collisions.
However, the 2~M$_\odot$ mass constraint is obtained at the price
of a too high critical density for a deconfinement transition during
collapse.
This statement is not the final word.
It is possible, e.g., via a density dependence of the gluon thermodynamics
which is absent in the present model, to allow for a simultaneous description
of deconfinement in the supernova interior and high mass (proto)
neutron stars.
\section*{Acknowledgment}
The authors would like to thank Hovik Grigorian for his comments and helpful discussions. The project was funded by the Swiss National Science Foundation (SNF) under project numbers~PP00P2-124879/1, 200020-122287, and the Helmholtz Research School for Quark Matter Studies. T.F. is supported the SNF under project number~PBBSP2-133378 and by HIC for FAIR. G.M.P. is partly supported by the Sonderforschungsbereich~634, the ExtreMe Matter Institute EMMI, the Helmholtz International Center for FAIR, and the Helmholtz Association through the Nuclear Astrophysics Virtual Institute (VH-VI-417). D.B., T.K. and R.{\L}. receive support from the Polish Ministry for Science and Higher Education. D.B. acknowledges support from the Russian Fund for Basic Research under grant No. 11-02-01538-a. R.~{\L}. received support from the Bogoliubov-Infeld programme for visiting JINR at Dubna where part of his work was done. The work of G.P. is supported by the Deutsche Forschungsgemeinschaft (DFG) under Grant No.~PA~1780/2-1 and J.S.-B. is supported by the DFG through the Heidelberg Graduate School of Fundamental Physics. M.H. acknowledges support from the High Performance and High Productivity Computing (HP2C) project. S.T. is supported by the DFG cluster of excellence ''Origin and Structure of the Universe''. F.S. acknowledges support from the Belgian fund for scientific research (FNRS). I.S. is supported by the Alexander von Humboldt foundation via a Feodor Lynen fellowship and wishes to acknowledge the support of the Michigan State University High Performance Computing Center and the Institute for Cyber Enabled Research. The authors are additionally supported by CompStar, a research networking program of the European Science Foundation, and the Scopes project funded by the Swiss National Science Foundation grant.~no.~IB7320-110996/1.
|
\chapter{MHD turbulence}
\author[S.M.~Tobias, F. Cattaneo \& S. Boldyrev]{S.M.~Tobias, F. Cattaneo \& S. Boldyrev}
\input introduction.tex
\input dynamo_new.tex
\input meanfield.tex
\input conclusions.tex
\input all_driver_edited.bbl
\end{document}
\section{Conclusions}
The last two decades have seen many developments in the field of MHD turbulence. There is little doubt that the driving force behind many of these has been the remarkable
increase in computing power. In a field in which it so difficult to carry out experiments or detailed observations, researchers have relied almost entirely on numerical simulations for the
underpinning of their theoretical speculations and modelling. It is now a matter of routine to carry out simulations with in excess of one billion grid-points; and the expectation is that in
the near future simulations with $10^{12}$ grid-points will become feasible. With these resolutions it is possible to explore fully three-dimensional configurations with moderately high
Reynolds numbers. In other words, the turbulent regime has become available to computations. Given these advances and these possibilities, it is natural to conclude this review with some
assessment of what has been learned --- and is now considered fairly certain --- and what still remains unclear or controversial, and will probably occupy the minds of researchers, and the CPU's of
their computers, in the years to come.
Within the framework of dynamo theory it is now widely accepted that given a random velocity, dynamo action is always possible provided the magnetic Reynolds number is large enough. In the
case of reflectionally symmetric turbulence (i.e.\ non-helical) the effort required to drive a dynamo appears to be determined by the slope of the energy spectrum of the velocity at
spatial scales comparable to those on which magnetic reconnection occurs. The flatter the spectrum the harder it is to drive a dynamo. If this property is expressed in terms of the
velocity roughness exponent, a marked difference emerges between the mechanisms leading to dynamo action for a smooth velocity as opposed to those in which the velocity is rough. This
distinction between smooth and rough velocities then maps very naturally in two distinct regimes corresponding to large and small values of the magnetic Prandtl number. What is also clear now, is that this stochastic theory must be modified if coherent structures are present in the turbulence. These may, depending on their properties, take over the control of the dynamo growth rate. We are now only beginning to
understand which properties of coherent structures contribute to the dynamo process, and certainly this will be an active area of future research. In the nonlinear regime, there are now
compelling models for the saturation process leading to the establishment of a stationary turbulent state for dynamos at large Magnetic Prandtl number. These models, in general, rely on some
specific geometrical property of the the magnetic field, like foliation, and seem to be borne out by the existing simulations. It remains to be seen if the predictions of these, mostly
phenomenological models, continue to hold at higher values of $Rm$. By contrast, the low magnetic Prandtl number regime remains largely unexplored. The difficulties are both conceptual,
since the velocity correlator has a non-analytic behaviour at the reconnection scales, and numerical, since reproducing a wide separation of scales between the magnetic and velocity boundary
layers is extremely computationally expensive; in this review we have determined the size of the calculation required to begin to settle the issue. The general expectation is, however, that the turbulence should become independent of of the magnetic Prandtl number provided the latter is small enough.
The existence of this asymptotic regime remains conjectural pending the availability of either much bigger computers, or much better analytical approaches.
When the flows are affected by the large scale component of the magnetic field, irrespective of whether the latter is self generated or imposed externally, the techniques utilised in analysing the turbulence are different.
Two important ideas have emerged that
have strongly influenced our understanding of the turbulent process in this case. The first is that depending on the dominant mechanism responsible for energy transfer MHD turbulence can be either weak
or strong. In the weak case, mean-fluctuation interactions dominate, whereas in the strong case, mean-fluctuation and fluctuation-fluctuation interactions are comparable. The common wisdom is
that the strength of the fluctuation-fluctuation interactions increases at small scales, so that if the inertial range is sufficiently extended there will always be a transition from weak to
strong turbulence. At present there are no cases in which this transition has been convincingly demonstrated numerically. The problem is associated with difficulties in reproducing a deep
inertial range. The situation will no doubt improve with the next generation of supercomputers. The second important idea is that MHD turbulence, unlike its unmagnetised counterpart, has a richly geometrical
structure. There is now universal agreement that the energy cascade is anisotropic with most of the energy being transferred transverse to the (local) mean field. The exact geometry of this
process is controlled by the requirement of a ``critical balance" between the crossing time of counter propagating wavepackets and the time for nonlinear transfer in the
field-perpendicular direction. Furthermore, there are strong indications, both analytical and numerical, that the polarization vectors of magnetic and velocity fluctuations tend to align in
the transverse plane. Although this phenomenon is intimately connected with the forward cascade of cross-helicity it has measurable consequences for the slope of the energy spectrum. This
appears shallower than predicted in the absence of alignment. Both the elongation of the wavepackets and the degree of alignment increase at small scales, and hence the picture that emerges of MHD
turbulence in the deep inertial range is that of a collection of interacting ribbon-like structures. Another problem that has recently emerged, and is at the moment the focus of considerable
interest, is that of unbalanced cascades. This occurs when there is an excess of wavepackets propagating in one direction relative to those propagating in the opposite
direction. The basic question here is whether unbalanced MHD turbulence is fundamentally different from balanced turbulence, or whether both are the same, with even balanced turbulence
being made up of interwoven unbalanced patches that appear balanced only on average. Again, a definite resolution of these issues may have to wait for a further increase in computing resources.
Finally, we should remark about the areas of active research in MHD turbulence that are not discussed in this review. In dynamo theory, the most important problem remains that of the
generation of large scale fields by turbulence lacking reflectional symmetry. At the moment, there appears to be a fracture between the conventional wisdom, mostly grounded in mean-field
theory, and an increasing body of numerical experiments. It is
possible that large-scale dynamos are ``essentially nonlinear" rather
than ``essentially kinematic" and a major revision of the theory might
be in order. It is also possible that the whole categorisation of dynamos into large and small-scale
is misleading and that it is more useful to seek ``system-scale'' dynamo solutions to the MHD equations (\cite{tcb:2010}).
In general, the lack of reflectional symmetry in
astrophysical turbulence owes its origins to the presence of
rotation. Strong rotation and shear flows are known to modify the
nature of MHD turbulence and indeed may lead to the generation of new
instabilties (\cite{chan:61,balhaw:1991}). Such magnetorotational
turbulence is of great interest owing to its importance for the
accretion process, and the nature of the turbulence may be very
different to that discussed in this review (see
e.g. \cite{balhaw:1998} and the references therein).
For relectionally symmetric turbulence, the most natural extension of
what has been presented is to incorporate compressibility. This bring in one more set of linear waves --- the fast magnetosonic waves --- and modifies the nature of the pseudo-Alfv\'en waves. The
resulting turbulence is extremely rich and is only now beginning to be explored in the nonlinear regime (see e.g. \cite{berlazcho:2005,lietal:2008,chandran08b}). Finally, we should note that all of our considerations have been discussed within the
framework of classical MHD. There is an almost bewildering variety of new effects that become important when the particular nature of a plasma is taken into account (see e.g. \cite{kulsrud:2005}). Again, the full richness
of this system will be a major preoccupation in the near future.
\section{Dynamo}
In this section we consider the case where there is no externally imposed magnetic field. It is well known that the
turbulent motion of
an electrically conducting fluid can lead to the amplification of a seed magnetic field (\cite{Parker:1979,Moffatt:1978}).
This generation process is termed dynamo action and can lead
to a substantial level of magnetisation. Two questions naturally arise. Under what conditions does dynamo action take
place and what is the
final state of the turbulence and magnetic field in the magneto-fluid? We discuss these questions in turn.
\begin{figure}
\begin{minipage}[b]{0.5\linewidth}
\centering
\includegraphics[width=0.99\textwidth]{tmp_b_new.pdf}
\end{minipage}
\hspace{0.1cm}
\begin{minipage}[b]{0.5\linewidth}
\centering
\includegraphics[width=0.99\textwidth]{tmp1.pdf}
\end{minipage}
\caption{Possible schematic spectra for kinetic energy (solid line) and magnetic energy (dashed line) for dynamo action
at (a) high and (b) low $Pm$.
When $Pm$ is large the resistive scale for the magnetic field is much smaller than a typical eddy and so the flow
appears large-scale and smooth. When $Pm$ is small, the resistive scale lies in the inertial range of the turbulence
where the velocity is rough.}
\label{low_vs_high_pm}
\end{figure}
\subsection{The onset of turbulent dynamo action}
We consider a situation in which an electrically conducting fluid, confined within some bounded region of space, is in a
state of
turbulent motion, driven by stationary forces. At some time, a very weak magnetic perturbation is introduced into the
fluid. The turbulent motions will stretch the magnetic field and
will therefore typically lead either to its amplification or its decay. In the case of growth the field will eventually
become strong, and the nonlinear Lorentz force term in the momentum
equation will become important. This force will modify the turbulence in such a way that the field no longer grows and
the system will settle down to a stationary state. However, initially, as the initial field is small, the
Lorentz force is negligible and the question of growth versus decay of the field can be addressed by assuming that the
velocity is given and by solving the induction
equation alone. This defines the kinematic dynamo problem.
For a prescribed velocity field, this is a linear problem for the magnetic field $\bB$. There are three cases that are
typically considered within the literature: for a steady velocity
one can seek solutions of the form $\bB = \bB({\bf x}) \exp \sigma t$ and the induction equation becomes a classical
eigenvalue problem for the dynamo growth-rate $\sigma$. The value of $Rm$
for which the real part of the growth-rate ($\Re(\sigma)$) becomes positive is termed the critical magnetic Reynolds number $Rm^{crit}$ for the onset of
dynamo action. Another commonly considered case is one where the velocity is periodic in time, then the induction
equation defines a Floquet problem and solutions consist of a periodic and an exponentially growing part and again one
can define
$Rm^{crit}$ in analogy with the steady case. Finally there is the case in which the velocity is a stationary random
process. Here one utilises a statistical description of the field, and seeks conditions under which the moments of the
probability distribution function for $B = |\bB|$ grow exponentially (\cite{zelruzsok:1990}). In all of these cases
$Rm^{crit}$ corresponds to that value at which the induction processes overcome diffusion. Na\"ively one would expect
that once $Rm > Rm^{crit}$ dynamo action would be guaranteed, but
this, actually, is not necessarily so. In fact establishing
dynamo
action in the limit $Rm \rightarrow \infty$, the so-called fast-dynamo problem, is technically very difficult (see for
instance \cite{chilgil:1995}). For instance it has been shown that flows that do not have chaotic streamlines can not be
fast dynamos (\cite{klapper:1995}). This exemplifies the somewhat paradoxical role of diffusion in dynamo action. On the
one hand too much diffusion suppresses dynamo action, on the other hand not enough diffusion also makes dynamo action
impossible. The reason is that reconnection is required in order to
change the magnetic topology to allow for the growth of the
field
(see e.g. \cite{dormysoward:2007}). In fact, as we shall see, the diffusive scale at which reconnection occurs plays an important
role in determining the properties of dynamo action.
Once a growing solution has been identified, it is of interest to determine properties of the eigenfunction. The
detailed properties depend on the precise form of the velocity, but typically most of the energy of the growing
eigenfunction is concentrated at the reconnection scales. However there is considerable interest, mostly astrophysically
motivated, in cases where a significant fraction of the energy is found on scales larger than a typical scale for the
velocity. This is called the large-scale dynamo problem and is most commonly discussed within the framework of mean
field electrodynamics (\cite{Moffatt:1978,krauraed:1980}). One of the early successes of this kinematic theory was to
establish that a lack of reflectional symmetry of the underlying flow is a necessary condition for the generation of
large-scale fields. There is a substantial body of literature that discusses the large-scale dynamo problem. It is fair
to say that currently there is considerable controversy as to whether mean-field electrodynamics can be applied in cases
where $Rm$ is large (see e.g. \cite{bransub:2005,bolcatt:2004,hugcatt:2009}). In this present review we shall not discuss at
length the large-scale problem, but instead focus on the amplification and saturation of magnetic fields of any scale.
Although anti-dynamo theorems exclude the possibility of dynamo action for certain flows and magnetic fields that
possess too much symmetry, for example two-dimensional flows or axisymmetric fields (\cite{cowling:1933,Zeldovich:1956}),
it is
well established that most sufficiently complicated laminar flows do lead to growing fields at sufficiently high $Rm$.
However here we are interested in the corresponding question for turbulent flows. In order to address this issue we need
to discuss two distinct possibilities corresponding to the cases of small and large magnetic Prandtl number $Pm$.
\subsection{High $Pm$ versus low $Pm$, smooth versus rough}
For the purposes of this discussion, we shall assume that the fluid Reynolds number is high so that the flow is turbulent
with a well-defined inertial range that extends from the integral scale $l_0$ to the dissipative scale $l_\nu$, and a
dissipative sub-range for the scales $l < l_\nu$.
It is useful to define the second order structure function $\Delta_2(r) = \langle |(\bv(\bx,t) - \bv(\bx+{\bf
r},t)).{\bf e}_r|^2 \rangle$, where $r = |{\bf r}|$ and ${\bf e}_r={\bf r}/r$, where we have assumed homogeneity and
isotropy. We can then characterise the inertial and dissipative ranges by the scaling exponents of $\Delta_2(r)$ with
$\Delta_2(r) \sim r^{2 \alpha}$, where $\alpha$ is the roughness exponent. In the dissipative sub-range we expect the
velocity to be a smooth function of position and therefore $\alpha=1$, whilst for the inertial range the velocity
fluctuates rapidly in space, i.e. it is {\it rough} and $\alpha < 1$;
for example, for Kolmogorov turbulence $\alpha = 1/3$.
The inertial range can also be characterised by the slope of the energy spectrum $E_k \sim k^{-p}$ where $p$ is related
to the roughness exponent by $p = 2 \alpha+1$. The dissipative scale is also related to $\alpha$ and given by
$l_\nu = l_0 Re^{-1/1+ \alpha}$.
For this given velocity field the scale at which magnetic dissipation becomes important and reconnection occurs can be
analogously defined by $l_\eta = l_0 Rm^{-1/1+ \alpha}$. The ratio
of the two dissipative scales is given by $l_\nu / l_\eta = Pm^{1/1+ \alpha}$. Clearly, irrespective of the value of
$\alpha$, if $Pm \gg 1$ then the resistive scale $l_\eta$ is much smaller than the viscous scale $l_\nu$, and therefore
reconnection occurs in the dissipative sub-range where the velocity is spatially smooth, as in
Figure~(\ref{low_vs_high_pm}a). By contrast, if $Pm \ll 1$ the dissipative scale is much greater than the viscous scale
and therefore reconnection occurs in the inertial range where the velocity is rough and therefore fluctuates rapidly, as
shown in Figure~(\ref{low_vs_high_pm}b). Therefore $Pm=1$ defines the boundary between dynamo action driven by rough and
smooth velocities.
It will become apparent that in general it is harder to drive a dynamo with a rough velocity than with a smooth
velocity. We now examine some specific examples.
\subsection{Random dynamos - the Kazantsev formulation}
\begin{figure}
\begin{minipage}[b]{0.5\linewidth}
\begin{center}
\includegraphics[width=0.99\textwidth]{Figure_2.pdf}
\end{center}
\end{minipage}
\hspace{0.1cm}
\begin{minipage}[b]{0.5\linewidth}
\begin{center}
\includegraphics[width=0.99\textwidth]{Figure_3.pdf}
\end{center}
\end{minipage}
\caption{(a) Schematic representation of the potential $V(r)$ and the spatial part of the longitudinal magnetic correlator
$H_L(r)$ in the Kazantsev model. The potential has a $1/r^2$ behaviour in the inertial range. Its overall height is
determined by the velocity roughness, and its large $r$ behaviour by the boundary conditions. The magnetic correlator is
peaked at small scales and decays exponentially for large values of $r$.
(b) Critical value of the size parameter $L$ for dynamo action in the Kazantsev model as a function of the
velocity roughness $1+\alpha$. As the velocity becomes rougher the effort required to drive a dynamo increases}
\label{fig_2}
\end{figure}
We now look at the simplest model of (kinematic) dynamo action driven by a random flow, the so-called
Kazantsev model (Kazantsev 1968). This is the only known solvable model for dynamo action and,
despite its simplicity, it captures many of the relevant features of turbulent field generation.
The model is based on a prescribed Gaussian, delta correlated
(in time) velocity field, with zero mean and covariance
\begin{eqnarray}
\langle v_i({\bf x+r}, t)v_j({\bf x}, \tau) \rangle
=\kappa_{ij}({\bf x},{\bf r})\delta(t-\tau).
\label{correlator}
\end{eqnarray}
Isotropy and homogeneity imply that the velocity correlation
function has the form
\begin{eqnarray}
\kappa_{ij}({\bf r})=\kappa_N(r)\left(\delta_{ij}-\frac{r_ir_j}{r^2}
\right)+\kappa_L(r)\frac{r_ir_j}{r^2}.
\label{kappa}
\end{eqnarray}
Further, incompressibility gives
$\kappa_N=\kappa_L+(r\kappa_L^\prime)/2$, where the primes denote differentiation with respect to $r$, and now the
velocity statistics can be characterized by the single scalar function
$\kappa_L(r)$.
A corresponding expression for the magnetic correlator can be defined by
\begin{eqnarray}
\langle B_i({\bf x+r}, t)B_j({\bf x}, t) \rangle=H_{ij}({\bf x},{\bf r}, t),
\label{bcorrelator}
\end{eqnarray}
where, analogously to (\ref{kappa}), the $H_{ij}$ function can be
represented as
\begin{eqnarray}
H_{ij}=H_N(r,t)\left(\delta_{ij}-\frac{r_ir_j}{r^2}
\right)+H_L(r,t)\frac{r_ir_j}{r^2}.
\label{hdef}
\end{eqnarray}
Similarly, the
condition $\nabla\cdot {\bf B}=0$ gives $H_N=H_L+(rH_L^\prime)/2$.
It now remains to determine the equation governing the evolution of $H_L(r,t)$ in terms of the input function
$\kappa_L(r)$. This equation is derived by differentiating equation~(\ref{bcorrelator}) with respect to time, using the
dimensional version of the induction equation and expressions
(\ref{correlator}), (\ref{kappa}) and (\ref{hdef}). Straightforward manipulation yields
\begin{eqnarray}
{\partial_t H}_L=\kappa H_L^{\prime \prime} + \left(\frac{4}{r}\kappa+\kappa' \right)H_L^{\prime}
+\left( \kappa''+\frac{4}{r}\kappa' \right)H_L,
\label{kequation}
\end{eqnarray}
where the
the `renormalized' velocity correlation
function $\kappa(r)=2\eta+\kappa_L(0)-\kappa_L(r)$ has been introduced, and $\eta$ is the magnetic diffusivity.
Equation~(\ref{kequation}) was originally derived
by Kazantsev (1968), and can be rewritten in a related form
that formally coincides with the Schr\"odinger equation in imaginary time.
This is achieved by effecting the change of variable,
$H_L=\psi(r,t) r^{-2}\kappa(r)^{-1/2}$, to obtain
\begin{eqnarray}
\partial_t \psi=\kappa(r)\psi''-V(r)\psi,
\label{sequation}
\end{eqnarray}
which describes the wave function of a quantum particle with variable
mass, $m(r)=1/[2\kappa(r)]$, in a one-dimensional potential~($r>0$):
\begin{eqnarray}
V(r)=\frac{2}{r^2}\kappa(r)-\frac{1}{2}\kappa''(r)-\frac{2}{r}\kappa'(r)-
\frac{(\kappa'(r))^2}{4\kappa(r)}.
\end{eqnarray}
Different authors have investigated the solutions of equation~(\ref{sequation}) for various choices of $\kappa(r)$ (see
\cite{zelruzsok:1990, arphorvai:2007,chertetal:1999}).
Here we restrict attention to the two extreme cases in which $l_\eta$ is in the deep dissipative subrange ($Pm \gg 1$)
and the velocity is smooth, and the case where $l_\eta$ is in the inertial range ($Pm \ll 1$) and the velocity is rough,
so that $\alpha < 1$. For the smooth case, exponentially growing solutions of equation~(\ref{kequation}) can be found
and the magnetic energy spectrum $E_M$ is peaked at the dissipative scale. The spectrum in the range $1/l_\eta < k <
1/l_\nu$ has a power law behaviour with $E_M \sim k^{3/2}$, irrespective of the spectral index for the velocity in the
inertial range (\cite{kuland:1992}). This regime for a smooth velocity
is also referred to as the Batchelor regime, since this is the regime
studied by Batchelor (1959a) for passive-scalar advection
For the rough case $\kappa(r) \sim r^{1+\alpha}$ in the inertial range. This scaling follows from the fact that Eq.~(\ref{sequation}) depends only on the time-integral of the velocity correlation function (\ref{kappa}), that is, on the turbulent diffusivity. When matching the Kazantsev model with a realistic velocity field, one therefore needs to match the turbulent diffusivities. In the Kazantsev model the diffusivity is given by $\kappa(r)$, while for a realistic velocity field it is estimated as $\Delta_2(r)\tau(r)$, where $\tau(r)\sim r/[\Delta_2(r)]^{1/2}$ is the typical velocity decorrelation time. This leads to the inertial-interval scaling $\kappa(r)\sim \Delta_2(r)\tau(r)\sim r^{1+\alpha}$ presented
above. The
Schr\"odinger equation~(\ref{sequation}) therefore has the effective
potential $U_{eff}(r)=V(r)/\kappa(r)=A(\alpha)/r^2$ in the inertial range,
where $A(\alpha)=2-3(1+\alpha)/2-3(1+\alpha)^2/4 $. At small scales ($l < l_\eta$) this effective potential is regularised
by magnetic diffusion as shown in Figure~(\ref{fig_2}a). Growing dynamo solutions
correspond to bound states for the wave-function $\psi$, which are guaranteed to exist if $A(\alpha) < -1/4$. Since this
is always the case for $0 < \alpha < 1$, this demonstrates that dynamo action is always possible even in the case of a rough
velocity (\cite{bolcatt:2004}). The corresponding wave-function will be concentrated around the minimum of the potential
at $r \sim l_\eta$ and it will decay exponentially for $r > l_\eta$ (see Figure~\ref{fig_2}). We stress at this point that
the effective potential always remains $1/r^2$ in the inertial range with its depth decreasing as the roughness
increases ($\alpha \rightarrow 0$). This justifies our previous statement that it is harder to drive dynamo action the
rougher the velocity.
An asymptotic analysis of the solution to equation~(\ref{sequation}) demonstrates that the growth-time, $\tau$, is of
the order of the turnover time of the eddies at the diffusive scale ($l_\eta$). This makes good physical sense since
these are the eddies that have the largest shear rate. Furthermore the spatial part of the wave-function for large $r
\gg l_0$ decays exponentially as $\exp \left[-2 \tau^{-1/2}r^{(\alpha-1)/2}/(1-\alpha)\right]$. The presence of the factor $1-\alpha$
in the denominator implies that if $\alpha$ is close to unity (smooth velocities) the wave-function is localised close
to the resistive scale. On the other hand, if $\alpha$ is close to zero (rough velocity) the wavefunction is more spread
out. This can be used to determine the requirements for the onset of dynamo action, which can be expressed either in
terms of a critical magnetic Reynolds number or in terms of a ``size parameter" $L$. The latter is a more useful measure,
since it relates directly to the effort, computational or experimental, that is needed to achieve dynamo action.
Consider equation~(\ref{sequation}) as a two-point boundary value problem in a finite domain of size $l_0$ (see e.g.
\cite{bolcatt:2004}), with boundary conditions $\psi(0) = \psi(l_0) = 0$. As the velocity becomes rougher the
wave-function spreads out and a larger domain (i.e.\ larger $l_0$) is required to contain the wave-function.
Figure~(\ref{fig_2}(b)) shows the minimum value of $l_0$ measured in units of $l_\eta$ (which is the size parameter $L$) for
which a growing solution can be found as a function of $1+\alpha$. Clearly there is a dramatic increase in $L$ as the
velocity becomes rougher (1+$\alpha$ gets smaller). Hence the effort required to drive a dynamo also increases. \footnote{An earlier attempt to solve the Kazantsev model for small $Pm$ was made in (Rogachevskii \& Kleeorin 1997). However, this paper employed an incorrect asymptotic matching procedure for deriving the dynamo threshold and the dynamo growth rate. The correct analysis was presented in \cite{bolcatt:2004}.}
We now return to the question posed at the start of this section, what is the effect of changing $Pm$ for the onset of
dynamo action in a random flow? At high $Pm$ the dynamo operates in the dissipative sub-range where the velocity is smooth and so the
effort necessary to drive a dynamo is modest. This state of affairs continues until $Pm$ decreases through unity. At
that point, the viscous scale becomes smaller than the resistive scale and the dynamo begins to operate in the inertial
range, where the velocity is rough, with a corresponding increase in the effort required to sustain dynamo action. Once
the dynamo is operating in the inertial range, further decreases in $Pm$ do not make any difference to the effort
required. In terms of critical $Rm$ as a function of increasing $Pm$ the curve takes the form of a low plateau and a high
plateau joined by a sharp increase around $Pm=1$ --- see for example the curves in Figure~3. On the other hand, if a sequence of calculations is carried out at
fixed $Rm$ and increasing $Re$ a sharp drop in the growth-rate $\sigma$ will be observed when $Re \approx Rm$
(\cite{bolcatt:2004}). If the initial $Rm$ is moderate, this could lead to the loss of dynamo action at some $Re$, which
could be misinterpreted as a critical value of $Pm$ at which dynamo action becomes impossible (\cite{christensenetal:1999,
yousefetal:2003,
schekochihietal:2004c,schekochihinetal:2005c}). We shall return to this theme later.
The Kazantsev model is, of course, based on a number of somewhat restrictive assumptions, one of which is that the flow considered is completely random and has no coherent part. We discuss in the next section the role of coherent structure in dynamo problems. Within the statistical framework there has been substantial
effort to extend the basic model to more general cases. One common criticism is that delta-correlated velocities are
artificial, with real turbulence having correlation times of order the turnover time. However, one should remember that
in most turbulent flows the turnover time decreases as one goes to small scales. Near onset the dynamo growth time can
be very large compared with the turnover time --- and hence the correlation time --- of the small eddies that typically
participate in the dynamo process. Noting that the growing magnetic fluctuations are predominantly concentrated at the resistive scales, where the relative motion of magnetic-field lines is affected by magnetic diffusion and, therefore, these lines do not separate with the eddy turnover rate, one expects that the assumption of zero correlation time is not as restrictive as may first
appear.
Once this is realised, it is to be expected that the dynamo behaviour near onset should be similar for cases with short
but finite correlation time to that with zero correlation time, and indeed this is confirmed by models in which the
correlation time of the turbulence is finite
(\cite{vainkit:1986,kleerogsok:2002}). A second possible extension is to flows that lack reflectional symmetry. In this
case the velocity and magnetic correlators are each defined by two functions, one as before related to the energy
density (either kinetic or magnetic), the other related to the helicity (either kinetic or magnetic)
(\cite{vainkit:1986}). In this case similar analysis leads to the derivation of a pair of coupled Schroedinger-like
equations for the two parts of the magnetic correlator (\cite{vainkit:1986,bergerros:1995,kimhug:1997}). Although the
analysis is now more involved, it is possible to show that the system remains self-adjoint (\cite{bolcattros:2005}). It
can also be shown that for sufficient kinetic helicity there is the possibility of extended states --- these are
states that do not decay exponentially at infinity and can be interpreted as large-scale dynamo solutions, in direct analogy with the solutions of the well-known alpha-dynamo model by \cite{steenbeck66}. However, at
large $Rm$ the largest growth-rates remain associated with the localised bound states, so that the overall dynamo
growth-rate remains controlled by small-scale dynamo action
(\cite{bolcattros:2005,malbol:2007,malbol:2008a,malbol:2008b}). Anisotropic versions of the Kazantsev model have also
been constructed by Schekochihin {\it et al.} (2004b).
\begin{figure}
\centering
\includegraphics[width=1.0\textwidth]{nj2.pdf}
\caption{Onset of dynamo action at moderate $Pm$, from Schekochihin {\it et al.} (2007). (a) Growth/decay rate of
magnetic energy as a function of Pm for five values of Rm. (b) Growth/decay rates in the parameter space (Re, Rm). Also
shown are the interpolated stability curves $Rm^{crit}$ as a function of $Re$ based
on the Laplacian and hyperviscous runs are shown separately. For an in-depth discussion of the reason for the apparent
discrepancy between the curves with and without a mean flow see Schekochihin {\it et al.} (2007)}
\label{crit_rm_low_pm}
\end{figure}
Finally there are extensions of the model that take into account departures from Gaussianity. Again on physical grounds
one does not expect this to introduce fundamental changes from the Gaussian case. The reason is similar to the one given
above; near marginality the growth time is long compared with the eddy turnover time and so the dynamo process feels the
sum of many uncorrelated events, which approaches a Gaussian even if the individual events themselves are not. This situation
arises in numerical simulations of dynamo action based on the full solution of the induction equation for velocities
derived from solving the Navier-Stokes equations. Many such efforts are summarised in figure~(\ref{crit_rm_low_pm}), which shows the critical
magnetic Reynolds number as a function of Reynolds number for a collection of such calculations (\cite{scheketal:2007}).
It is clear that at large and moderate $Pm$ these results are consistent with the predictions of the Kazantsev model
above --- the plateau at high $Pm$ is succeeded by a jump in the critical $Rm$ when $Pm$ approaches unity --- such a curve is visible in Figure~\ref{crit_rm_low_pm}. Even the size
of the jump is consistent with the analysis above. However, since the numerical cost of resolving the very thin viscous
boundary layers rapidly becomes prohibitive at small $Pm$, this regime is not really accessible to direct numerical
simulation (DNS).
One scheme to alleviate this problem is to use large eddy simulations (LES) to generate the velocity
(\cite{pontyetal:2007}). We would advise great caution when using this approach. As the discussion above shows, small
changes in the roughness exponent can lead to huge changes in the critical $Rm$ or alternatively the dynamo growth-rate.
In the case of LES, one would need to be able to control very delicately how the velocity roughness is related to the
smoothing scale. We also note that the jump in critical $Rm$ is captured by shell models of dynamo action (see e.g.
\cite{fricksok:1998}) for which it is possible to achieve a large separation of the viscous and resistive dissipation
scales. We are not certain what, if anything, to make of this.
One of the interesting features of the Kazantsev model is that {\it for random flows} the only thing that matters for
the onset of dynamo action is the roughness exponent of the spectrum in the neighbourhood of the dissipative scale.
Therefore, within this framework, features associated with high-order moments, or large-scale boundary conditions do not
matter.
Thus from the point of view of dynamo onset a velocity derived from solving the Navier-Stokes equations, with random
forcing should yield qualitatively similar results to a synthesised random velocity with the same spectrum.
So far this seems to be consistent with the results of numerical simulations. However, for flows with a substantial non-random
component, outside of the range of validity of the Kazantsev model it may be that characteristics other than the
spectral slope of the velocity do play a key role in determining dynamo onset, and it is this possibility that we
address in the next section. The results of the next section should therefore act as a caveat when considering the applicability of
the results of this section.
\subsection{Coherent Structure dynamos}
\begin{figure}
\centering
\includegraphics[width=1.0\textwidth]{rfig4.pdf}
\caption{Dynamos with and without coherent structures. Out-of-plane velocity (left) and $B_x$ (right) for flows with
coherent structures (top) and
purely random structure (bottom) --- after Tobias \& Cattaneo (2008b). The velocity with coherent structures is a better
dynamo and generates more filamentary magnetic fields.}
\label{coh_stru}
\end{figure}
We now turn to the case where the flow consists of two components, a random component as described above and a more
organised component whose coherence time is long compared with the turnover time (see e.g. \cite{Mininni-etal:2005}). There are many examples of naturally
occuring flows that can be characterised in this way; for example, flows in an accretion disks that consists of
small-scale turbulence and coherent long-lived vortices (see e.g. \cite{Bracco:1999,GodLiv:2000,BalmKor:2001}, Taylor
columns in rapidly rotating turbulent planetary interiors (see e.g. \cite{busse:2002}) or the large-scale flows driven
by propellors in laboratory experiments (see e.g. \cite{monchauxetal:2007}). In all these cases the turbulence arises
because the Reynolds number is huge, whilst the coherent part is associated with some constraint, such as rotation,
stratification or large-scale forcing. The natural question then is what determines the dynamo growth-rate?
In the last section we gained some understanding for the case of entirely random flows --- we know that if the dynamo
were entirely random then the dynamo growth-rate would largely be determined by the spectral index of the flow at the
dissipative scale. Conversely, if the flow were laminar again we could appeal to laminar dynamo theory --- but would not
try to characterise the flow in terms of spectra.
Here we need to consider the case where the flow is the sum of the two components, bearing in mind that the dynamo of
the sum is not the sum of the dynamos (see e.g. \cite{catttob:2005}). We address these issues by examining a specific case. We consider a flow that has
both long-lived vortices and a random component with a well-defined spectrum. For those who are interested, such flows
arise naturally, for instance, as solutions to the active scalar equations (\cite{tobcatt:2008b}). In tandem we also
generate a second flow with the same spectrum but no coherent structures obtained by randomising the phases. Figure~(\ref{coh_stru})
compares the out-of plane velocity for the two flows. We consider the dynamo properties for large $Rm$ so we anticipate
that dynamo action would occur for either flow. Therefore, if the considerations of the previous section were to apply,
these two flows would exhibit very similar dynamo properties. However, this is actually not so --- the flow with the
coherent structures is a more efficient dynamo in the sense that at the same $Rm$ the dynamo growth-rate is higher.
Moreover the structure of the resulting magnetic field (shown in Figure~\ref{coh_stru}) indicates that the coherent
structures play the major role in field generation.
Given the above considerations, it is reasonable to ask what controls dynamo action in a flow with a superposition of
coherent structures with different spatial scales and different turnover times? This is equivalent to a turbulent
cascade, but here each of the eddies are long-lived with a coherence time greater than their local turnover time. For
the general case this is a very difficult question to answer. However, there is a case in which some statements can be
made --- namely when each of the dynamo eddies are `quick dynamos'. A `quick dynamo' (as defined by
\cite{tobcatt:2008a}) is one that reaches a neighbourhood of its maximal growth-rate quickly as a function of $Rm$. In
that case it can be shown that the dynamo is driven by the coherent eddy which has the shortest turnover time $\tau$ and
has $Rm \geq {\cal O}(1)$. Since both the turnover time of the eddy and the local $Rm$ are functions of the spatial
scale of the eddy and the slope of the spectrum, the location in the spectrum of the eddy responsible for dynamo action
also depends on the spectral slope as well as the magnetic Reynolds number on the integral scale (\cite{tobcatt:2008a}).
Some of these ideas become particularly germane in explaining the behaviour of dynamos in liquid metal experiments. The
typical laboratory dynamo device consists of a confining vessel and propellors, designed to drive a large-scale flow
with desirable laminar dynamo properties; i.e.\ with low $Rm^{crit}$. Because the magnetic Prandtl number of liquid
metals is tiny ${\cal O} (10^{-6})$ the corresponding Reynolds numbers are huge. Thus the actual flow consists of the
large-scale flow planned by the designers plus turbulent fluctuations that can be comparable in magnitude to the mean
flow. Invariably, it has been found that the actual critical magnetic Reynolds number for the mean flow and turbulence
is significantly higher than that envisaged for the laminar flow alone. It is important to realise that these devices
operate in a regime where at best the magnetic Reynolds number is close to the marginal one for the laminar flow; it is
definitely below the critical $Rm$ for the turbulent part. In all the cases discussed the role of turbulence is to
hinder the dynamo, and this can be understood in terms a renormalised diffusivity that increases with the degree of
randomness. This increase in diffusivity in astrophysical situations is irrelevant, since $Rm$ is so far above critical
that increasing the diffusivity makes little difference. However in the case of the experiments, this increase has a
catastrophic effect on the chances of success for the experiment\footnote{ Therefore we conclude that all funding for
experiments should immediately be channeled to theorists studying dynamos at large $Rm$ (or writing reviews)}.
\subsection{Saturation of turbulent dynamos}
The exponential growth of the magnetic field described in the last section is accompanied by an
exponential growth of the magnetic forces, which will eventually become comparable with
those driving the turbulence.
In this second nonlinear phase the exponential growth of
the magnetic field will become saturated and the magneto-turbulence will settle down to
some stationary level of magnetization. It is of interest to speculate about the nature of the saturation process, and
about the
general properties of the final state, although in general the saturation mechanism may be subtle (see e.g. \cite{catttob:2009})
As before there is a large difference between high and low magnetic Prandtl number regimes. In the high $Pm$ regime the
dynamo is operating at scales in the sub-inertial range of the velocity for which the Reynolds number is small.
Therefore the inertial term in the
momentum equation can be neglected and the velocity can be split into two components; one driven by the external forces,
which is the original velocity of the kinematic dynamo problem and has a characteric scale $\gg l_{\eta}$, the other,
driven by the Lorentz force, encodes the back-reaction of the magnetic field and is mostly concentrated around the
diffusive scale.
In this high $Pm$ regime saturation can be successfully addressed by semi-analytical models, phenomenological models and
numerical experiments. The semi-analytical models
are ultimately based on some closure of the MHD equations.
Within the framework of the Kazantsev equation the magnetically driven velocity produces a change in the velocity
correlator, which leads to the nonlinear saturation (\cite{sub:1999,sub:2003}). A similar approach can be taken by
constructing a Fokker-Planck equation for the probability distribution function for magnetic fluctuations rather than
an equation for the magnetic correlator. It can be shown that the coefficients of this equation again are determined by
the velocity correlation function which can be modified nonlinearly in a similar manner to above (\cite{bol:2001}).
An interesting phenomenological model has been proposed by Schekochihin and collaborators (\cite{scheketal:2002-amodel}).
The model begins with the kinematic growth of fields at the resistive scale, which is much smaller than the viscous
scale. In this kinematic phase the eddies driving the dynamo are the ones at the viscous scale. The authors anticipate
that nonlinear effects begin to be important when $\bv \cdot \nabla \bv \approx \bB \cdot \nabla \bB$ at that scale. The
left hand side is easily estimated to be $v^2/l_\nu$. To estimate the right hand side the authors use the foliated
structure of the magnetic field.
To have a geometrical understanding of what this means, it is useful to distinguish between the orientation and the
direction of a vector field. For example a change from horizontal to vertical is a change in orientation, whereas a
change from up to down is a change in direction. For a typical foliated field the orientation changes slowly on the
scale of the velocity, whereas the direction changes rapidly on the scale $l_\eta$ (\cite{finnott:1988}). Hence the
regions of high curvature occupy practically none of the volume and the right hand side is estimated to be $b^2/l_\nu$,
where $b$ is a typical field strength at the scale
$l_\eta$. Thus the nonlinear saturation begins when the magnetic energy comes into equipartition with the energy at the
viscous scale. The effect is to suppress the dynamo growth associated with the eddies at the viscous scale. This
suppression is not necessarily achieved by a dramatic reduction of the amplitude of the eddies but rather by a subtle
modification of their geometry. In particular if the velocity becomes more aligned with the local magnetic field then it
cannot distort that field and therefore contribute to its amplification. However, slightly larger eddies can still
sustain growth, albeit at a slightly slower rate. Growth will continue until the magnetic energy comes into
equipartition with the energy of these eddies. The process will continue to larger and larger eddies until the magnetic
energy reaches equipartition with the energy of the flow at the integral scale. This nonlinear adjustment is
characterised by a growth of the magnetic energy on an algebraic (rather than exponential) time and there is a
corresponding shift of the characteristic scale of the magnetic field to larger scales. This idea of scale-by-scale
modification of the velocity to reach some form of global equipartition can be formalised in terms of either
Fokker-Planck equations (\cite{scheketal:2002-amodel}) or a Kazantsev model which is appropriately modified to take
account of the growing degree of anisotropy and finite correlation time (\cite{scheketal:2004-satu}). There are even
models constructed where the final state does not reach global equipartition but only a fraction of equipartition.
According to Schekochihin {\it et al.} (2002), this occurs when $Pm$ is large but $Pm \le Re^{1/2}$ with $B^2/U^2 \approx
Pm/Re^{1/2}$.
In this scenario, however, by the time the final state is reached the characteristic scale of the magnetic field is
still smaller than the viscous scale. Schekochihin {\it et al.} (2002) estimate this to be the case when $Pm \gg Re^{1/2}$.
There have been a number of simulations at moderate to high $Pm$, and within the normal restrictions of numerical
simulations the results seem to conform to this general picture (\cite{maronetal:2004}). There is good evidence that in
the kinematic phase the magnetic spectrum is compatible with the $k^{3/2}$ prediction of the Kazantsev model. The
appearance of the magnetic field is indeed that of a foliated structure. In the saturated state, the magnetic and
kinetic energies are comparable, although for moderate values of $Pm$ the magnetic energy increases with $Pm$. As the
saturation progresses the magnetic spectrum grows and flattens, which is compatible with the creation of magnetic
structures larger than the resistive scale.
It is always the case that the magnetic energy exceeds the kinetic energy at small scales (see e.g.
\cite{caratietal:2006}). Moreover the magnetic field is more intermittent than the velocity --- indeed the pdf for the
velocity field remains close to that of a Gaussian whilst that for the magnetic field is better described by an
exponential --- although the degree of intermittency is reduced in the saturated state as compared with the kinematic
state (\cite{cattaneo:1999}).
\begin{figure}
\centering
\includegraphics[width=1.0\textwidth]{Figure_5.pdf}
\caption{Velocity and magnetic spectra in a numerical simulation of magnetised turbulence; courtesy of Mason and Malyshkin (private communication)}
\label{f5}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=1.0\textwidth]{Figure_6.pdf}
\caption{Second order longitudinal velocity structure function $\Delta_2(r)$ measured in a numerical simulation of
magnetised turbulence as a function of separation $r$. At small scales $\Delta_2(r) \sim r^2$; at larger separations its
power law behaviour defines the velocity roughness. Clearly, in order to represent numerically a rough velocity,
at least of the order of a few tens of gridpoints is required. Calculations courtesy of Bodo \& Cattaneo.}
\label{f6}
\end{figure}
Remarkably many of these features of the saturated state persist in the $Pm \approx {\cal O}(1)$ regime, which is much
more accessible to numerical simulations. These address various issues ranging from the degree of intermittency to the
nature of the kinetic and magnetic energy spectrum. Although there is no universal agreement there are a number of
features that are robust. The magnetic fluctuations are always more intermittent than the velocity fluctuations.
Furthermore there is a statistical anisotropy along and transverse to the local magnetic field, with a stronger degree
of intermittency transverse
(\cite{mullbisgrapp:2003}). These authors have argued that the intermittency results are consistent with a She-Leveque
scaling in which the more singular structures are current sheets. Concerning energy spectra, results show that the
kinetic and magnetic energy spectra track each other, with the magnetic energy spectra exceeding the kinetic one at all
but the largest scales, as in Figure~(\ref{f5}). The existence of an inertial range, characterised by a well-defined
power law, is still a matter for debate. There are various reasons for
this. One, of course, is that the resolution of the
numerical simulations is limited. Another one is the presence of bottleneck effects, whose intensity is linked to the
use of hyperdiffusivities. This, coupled with the limited resolution, can have effects that propagate into the putative
inertial range. Finally there is a tendency to present results sometimes in terms of the shell-averaged spectrum,
sometimes in terms of the one-dimensional spectrum relevant to a direction chosen for computational convenience and
sometimes in terms of a one-dimensional spectrum perpendicular to the local mean field. All these issues
notwithstanding, there are two things that can be said with certainty. Most researchers agree that the spectrum, if it exists, has a slope somewhere
between $-1.5$ and $-1.7$ --- we shall return to this issue in the next section. The second is that people have a
tendency to find in their results confirmation of their expectations, and to attribute deviations to other effects, such
as intermittency or bottlenecks. For example it is not unheard of for people to interpret a spectrum compatible with a
spectral slope of $-3/2$ as a spectrum of $-5/3$ with corrections induced by a bottleneck (\cite{haugenetal:2004}) or
intermittency (\cite{beresnyaklazarian:2006}).
No doubt as computers get bigger and numerics get better some kind of consensus will emerge.
At present the low $Pm$ regime remains largely unchartered. In this regime the dynamo is operating in the inertial range
of the turbulence. Inertial terms are certainly important and the velocity can no longer be thought of as consisting of
two components.
In this case analytical progress can only be made by resorting to standard closure models such as EDQNM
(\cite{pouquetetal:1976}). One of the basic ingredients of these theories is the decorrelation time of triple moments.
Whereas in hydrodynamic turbulence there is overwhelming evidence that this quantity is of the order of the turnover
time, there is no consensus for of what this quantity should be in MHD theory. Indeed it may depend sensitively on the
field strength and magnetic Reynolds number of the magnetised turbulence. For this reason it is not clear that utilising EDQNM is
a reliable way to make analytical progress. One of the problems is that, whereas in
hydrodynamic turbulence there is a vast body of experimental evidence, there is no such thing in MHD, and therefore one
is forced to rely almost entirely on numerical experiments, with all of their limitations. These limitations become
extremely severe in the low $Pm$ regime. There is a generic expectation that for sufficiently low $Pm$ an asymptotic
regime is reached where the level of saturation becomes independent of $Pm$ (\cite{petfauve:2008}). Some steps have been
taken to move in the direction of simulating dynamos at low $Pm$, but these are still in the $Pm \sim {\cal O}(1)$ regime,
with all the associated dynamics (see e.g. \cite{mininni:2006,iskakovetal:2007}). It is important to appreciate the
difficulties inherent in reaching the low $Pm$ regime numerically, and why caution should be exercised when interpreting
the results of numerical simulations that are described as being in this regime. The essence of the problem is that the
velocity field driving the dynamo is rough. It is possible to ask what sort of resolution is required to represent such
a velocity. Figure~(\ref{f6}) shows the second order structure function for a turbulent velocity; the distance here is
expressed in units of the grid spacing. At small distances the slope approaches two, implying that on a few
grid-points the function becomes {\it smooth}. It is important to realise that this feature persists even if the
dissipation is entirely numerical. For larger distances (i.e.\ more gridpoints) the slope becomes less than two and
eventually approaches a power law behaviour, characteristic of a rough velocity. So for example, in this case the range
of scales over which the function becomes rough is of the order of 20-30 gridpoints. In the low $Pm$ regime we should
think of this size as the thickness of a magnetic boundary layer that is able to recognise the velocity as rough.
Inspection of Figure~(\ref{fig_2}b) gives an estimate of the size of the domain required to contain a growing eigenfunction
in units of the magnetic boundary layer thickness. For a Kolmogorov-like velocity $1+\alpha = 1.333$ and so this is of the
order of 30. This would give of the order 600-900 gridpoints simply to capture the magnetically growing eigenfunction.
If one then wants a reasonable description of the velocity inertial range one probably requires at least a decade below
this, giving a requirement in terms of gridpoints of several thousands (or tens of thousands). Possibly one could save a
factor of ten by not matching the velocity inertial range to a resolved viscous sub-range by using LES. This is indeed
the approach used by Ponty {\it et al.} (2008), but again we caution the reader that playing this game requires exceptional control of
the LES, and even with the LES, simulations of the order of $1000^3$ are unavoidable.
Again, as machines get larger, the task of exploring this regime may not appear so daunting.
\section{Introduction}
Magnetic fields are ubiquitous in the universe (\cite{Parker:1979, zelrussok:1983}). Their interaction with an electrically conducting fluid gives rise to a complex system--a
magnetofluid--whose dynamics is quite distinct from that of either a non conducting fluid, or that of a magnetic field in a vacuum (\cite{cowling:1978}). The scales of these interactions
vary in nature from metres to megaparsecs and in most situations, the dissipative processes occur on small enough scales that the resulting flows are turbulent.
The purpose of this review is to discuss a small fraction of what is currently known about the properties of these turbulent flows. We refer the reader to several recent reviews for a
broader view of the field (\cite{biskamp:2003, galtier:2008, galtier:2009, lazarian:2006, lazariancho:2005, mullerbusse:2007, kulsrudzweibel:2008}, Bigot {\em et~al.} 2008, Sridhar 2010, Brandenburg \& Nordlund 2010).
The electrically conducting fluid most commonly found in nature is
ionized gas, i.e. a plasma, and its description in terms of all its
fundamental constituents is extremely complex (see e.g.
\cite{kulsrud:2005}). In many circumstances, however, these complexities can be neglected in favour of a simplified description in term of a single fluid interacting with a magnetic field.
Formally, this approach is justifiable when the processes of interest occur on timescales long compared with the light-crossing time, and on spatial scales much larger that any
characteristic plasma length. Despite its simplified nature, the resulting Magneto-Hydrodynamic (MHD) approximation is of great general applicability and can adequately describe many
astrophysical systems ranging from galaxy clusters, to the interstellar medium, to stellar and planetary interiors, as well as laboratory experiments in liquid metals. The dynamics of turbulent flows under this approximation will form the
basis for our discussion.
There are many similarities between turbulence in a magnetofluid and in a non-electrically conducting fluid --- hereinafter a regular fluid. The most fundamental is the idea of a turbulent
cascade whereby energy is transferred from a large injection scale by nonlinear interactions through an inertial range to small scales where it is converted into heat by dissipative
processes (see e.g. \cite{davidson:2004}). In analogy with regular turbulence, one of the objectives of theories of MHD turbulence is to predict the general form of the inertial range and
dissipative subrange, and to identify the important processes therein. There are, however, many significant differences. First of all, MHD describes the evolution of two vector fields, the
velocity and the magnetic field, and hence the specification of the state of the system is more involved.
A further crucial difference concerns the influence of global constraints. In regular turbulence global constraints become progressively weaker at small scales, and in general it is always
possible to find a scale below which the motions become unconstrained. For example, in a rotating fluid scales smaller than the Rossby radius hardly feel the rotation at all; in a
stratified fluid, motions on a scale much smaller than a representative scale height hardly feel the stratification.
In contrast, there is no scale in MHD turbulence below which the fluid becomes unmagnetized. This has an important consequence: the cascade in regular turbulence approaches an isotropic state at small
scales, whereas the cascade for MHD turbulence becomes progressively more anisotropic at small scales, this being true even if there is no large-scale field.
Moreover, an important effect in MHD turbulence that has no direct correspondence in regular turbulence is the so-called dynamo process whereby a turbulent electrically conducting fluid
becomes self magnetized. Finally, there are several differences between regular and MHD turbulence that are social in origin. We are constantly immersed in regular turbulence. We have a
direct experience of it in our everyday life. Thus our development of models and theories of regular turbulence is both strongly guided and strongly constrained by experimental data and
intuitions. Not so for MHD turbulence. Even though MHD turbulence is very widespread in the universe, we have practically no direct experience of it in our daily pursuits. Laboratory
experiments can be conducted but they are, in general difficult to perform, and difficult to diagnose (\cite{verhille-etal:2009}). As a result, MHD theorists have enjoyed a greater creative
freedom and have expressed it by generating an impressive array of competing theories; computer simulations being the only obstacle standing in the way of the expression of unbounded
imagination. As a testament to this, we note that whereas the form of the energy spectrum in isotropic homogeneous (regular) turbulence was very much a settled issue by the late forties
(for a historical perspective see \cite{Frisch:1995}), the corresponding problem in MHD turbulence is still very much open to debate today.
\subsection{Formulation and equations}
In this review we shall restrict our discussion to incompressible
MHD. In terms of the fluid velocity this is appropriate for subsonic
flows. In MHD an additional constraint on the magnitude of the
magnetic fluctuations is also required for the applicability of
incompressibility. It is
customary to define the plasma $\beta$ as the ratio of the gas
pressure to the magnetic pressure --- then incompressibility is appropriate for situations in which $\beta \gg 1$. This
condition is satisfied, for instance, in planetary and stellar interiors, but not in dilute plasmas such as those found in the solar corona and in many fusion devices (\cite{kulsrud:2005}). We note here also that in some specific circumstances the incompressible equations provide a good approximation even if $\beta \ll 1$.
The evolution of the magnetofluid is described by the momentum equation together with the induction equation and the requirements that both the velocity and magnetic fields be solenoidal. In
dimensionless form, these can be written as
\begin{eqnarray}
&
\partial_t \bfv + \bfv \cdot \nabla \bfv = - \nabla p + \bfJ \times \bfB + Re^{-1} \nabla^2 \bfv + \bff,
&\label{eq_mom}
\\
&
\partial_t \bfB = \nabla \times \left( \bfv \times \bfB \right) + Rm^{-1} \nabla^2 \bfB,
&\label{eq_ind}
\\
&
\nabla\cdot\bfB=\nabla\cdot\bfv=0,
&\label{eq_div}
\end{eqnarray}
where $\bfv$ is the fluid velocity, $\bfB$ is the magnetic field intensity, $p$ the pressure, $\bfJ$ is the cuurent density, and $\bff$ is the total body force, that could include rotation,
buoyancy, etc. Here, $Re=v_o \ell /\nu$ in the momentum equation is the familiar Reynolds number and $Rm$ in the induction equation is the corresponding magnetic Reynolds number defined
analogously by $Rm=v_o \ell /\eta$, where $\eta$ is the magnetic diffusivity; it measures the stength of inductive processes relative to diffusion. As is customary, we have assumed that the
density $\rho_o$ is constant and uniform, and we measure $\bfB$ in units of the equivalent Alfv\'en speed ($B/\sqrt{4 \pi \rho_o}$).
The ratio of the magnetic to the kinetic Reynolds number is a property of the fluid and is defined as the magnetic Prandtl number $Pm$. In most naturally occurring systems, $Pm$ is either
extremely large, like in the interstellar medium, the intergalactic medium and the solar wind, or extremely small like in the dense plasmas found in stellar interiors or liquid metals.
Interestingly, it is never close to unity except in numerical simulations. This has some important consequences in terms of what we can compute as opposed to what we would like to
understand; we shall return to this issue later. In this review, we shall concentrate on the cases in which $Re$ is large, so that the flows are turbulent, and $Rm$ can be small, as in
some liquid metal experiments, moderate, as in planetary interiors, or large as in most astrophysical situations.
For an ideal fluid, i.e.\ one in which $Rm$ is infinite, magnetic field lines move with the fluid as if they were frozen in. This is analogous to the behaviour of vortex lines in an inviscid
fluid. Indeed, Alfv\'en's Theorem states that the magnetic flux through a material surface is conserved in the same way as Kelvin's theorem asserts that the circulation of a material
contour is conserved (\cite{cowling:1978, Moffatt:1978}). In fact, there is a formal analogy between the induction and vorticity equations. It should be noted however that, whereas the
vorticity is the curl of the velocity, no such relationship exists between $\bfv$ and $\bfB$. As a result, arguments based on the formal analogy between the two equations can sometimes be
useful, and sometimes be misleading. In general, as a rule of thumb, if the magnetic field is weak compared with the velocity it behaves analogously to the vorticity; if it is comparable it
behaves like the velocity. This point will be discussed more fully later.
As in hydrodynamic turbulence, much insight can be gained from examining conservation laws. In the ideal limit there are three quadratic conserved quantities: the total energy
\begin{equation}
E = \half \int_V \left( v^2+B^2 \right) \, d^3 x,
\label{def_energy}
\end{equation}
the cross-helicity
\begin{equation}
H^c = \int_V \bfv \cdot \bfB \, d^3 x,
\label{def_cross}
\end{equation}
and the magnetic helicity
\begin{equation}
H = \int_V \bfA \cdot \bfB \, d^3 x,
\label{def_maghel}
\end{equation}
where $\bfA$ is the vector potential satisfying $\bfB = \nabla \times
\bfA$. Here the volume $V$ is either bounded by a material flux
surface, or is all space provided that
the fields decay sufficiently fast at infinity
(\cite{Moffatt:1978})\footnote{Many numerical simulations utilise
periodic boundary conditions, for which care must be taken in
defining the magnetic helicity.}. It should be noted that the limit $\bfB \rightarrow 0$ is a delicate one; only the energy survives as a conserved
quantity and a new conserved quantity the kinetic helicity appears. This emphasises the fundamental difference between hydrodynamics and MHD. It can be argued analytically and verified
numerically that, in the presence of small dissipation, energy decays faster than magnetic helicity and cross-helicity (\cite{biskamp:2003}). Therefore, in a turbulent state, energy cascades
toward small scales (analogous to the energy cascade in 3D hydrodynamics), with the magnetic fluctuations approximately in equipartition with the velocity fluctuations. The magnetic
helicity cascades toward large scales (analogous to the energy cascade in 2D hydrodynamics); the inverse cascade of magnetic helicity may lead to the formation of large-scale magnetic
fields, which are not in equipartition with the velocity fluctuations (\cite{pouquetetal:1976}).
The role of cross-helicity is more subtle. Having the dimension of energy, it also cascades towards small scales (\cite{biskamp:2003}). However, cross-helicity dissipation is not
sign-definite: cross-helicity may be either amplified or damped locally. As we shall see, this leads to local self-organization in the turbulent inertial interval.
We can now discuss the most significant difference between hydrodynamic and MHD turbulence. In hydrodynamic turbulence, mean flows or large-scale eddies advect small-scale fluctuations
without affecting their dynamics (\cite{batchelor:1953}). This is a consequence of the Galilean invariance of the Navier-Stokes equation. The situation is quite the opposite in MHD
turbulence. While the large-scale velocity field can be removed by Galilean transformation, the large-scale magnetic field cannot. Such a large-scale magnetic field could arise owing to a
number of different processes. It could be either generated by large-scale eddies, as in the interstellar medium, or imposed by external sources, as in the solar wind or plasma fusion
devises. The large-scale magnetic field mediates the energy cascade at all scales in the inertial interval. As a consequence, as we pointed out earlier, weak small-scale turbulent
fluctuations become anisotropic, since it is much easier to shuffle strong magnetic field lines than to bend them. The smaller the scale, the stronger the anisotropy caused by the
large-scale field. This is in a stark contrast with hydrodynamic turbulence where large-scale anisotropic conditions get ``forgotten'' by smaller eddies, so that the turbulent fluctuations
become isotropic as their scale decreases.
In this review, we shall organise our discussion into two main sections describing the dynamics of turbulence with and without a significant mean field. If the mean field is unimportant then
we neglect it altogether and consider the ``dynamo'' case, whereas if it is important we shall assume that it is strong and shall not concern ourselves about its origin. One should note
that this distinction may depend on scale and the same system may be well described by the ``dynamo'' case at some scales and the ``guide-field'' case at others. Finally, we shall mostly be
concerned with driven turbulence, that eventually evolves to a stationary state. There is a substantial body of literature considering turbulent decay (see e.g. \cite{biskamp:2003}) that
will not be discussed here
\section{Mean field}
We now turn our attention to the case in which there is an externally imposed magnetic field. As mentioned earlier, here the motivations are that in many astrophysically significant
situations the regions of interest are threaded by a large scale magnetic field, or that we are interested in small scales whose dynamics is influenced by a magnetic field at
larger scales. It is reasonable, and customary, to concentrate on the idealized case in which the large scale field is uniform.
We begin by discussing small perturbations to a uniform field; the magnetohydrodynamic equations support linear waves that propagate on the stationary and uniformly magnetized background. We
denote the uniform background magnetic field by ${\bf B}_0$ and choose a coordinate frame with the $z$-axis along~${\bf B}_0$. The magnetic field is then given by ${\bf B}({\bf x},t)={\bf
B}_0+{\bf b}({\bf x}, t)$, where ${\bf b}({\bf x}, t)$ is the fluctuating part. Instead of the guide field ${\bf B}_0$ we will often use the Alfv\'en velocity ${\bf v}_A={\bf B}_0/\sqrt{4\pi
\rho_0}$, where $\rho_0$ is a uniform fluid density. We seek wave-like solutions of the MHD equations for the velocity and magnetic fluctuations in the form ${\bf v}({\bf x}, t)={\bf
v}_k\exp(-i\omega t+i{\bf k}\cdot{\bf x})$ and ${\bf b}({\bf x}, t)={\bf b}_k\exp(-i\omega t+i{\bf k}\cdot{\bf x})$. Recall that we are considering the incompressible case, ${\bf k}\cdot
{\bf v}_k$=0, which rules out acoustic modes.
The remaining waves can be classified according to their polarizations.
There are two so-called ``shear-Alfv\'en'' waves. Their dispersion relation is~$\omega=|k_z|v_A$, and their polarization vector~${\bf b}_k$ is normal to both the wavevector~${\bf k}$ and
the guide field~${\bf B}_0$ (i.e.\ ${\bf b}_k \cdot {\bf k} = {\bf b}_k \cdot {\bf B}_0 = 0$). If the group velocity of a shear-Alfv\'en wave is in the direction of the guide field, its
velocity polarisation is anti-parallel to the magnetic polarisation, i.e. ${\bf v}_k =- {\bf b}_k$. If the group velocity of such a wave is in the direction opposite to the guide field, its
velocity polarisation is parallel to the magnetic polarisation, i.e. ${\bf v}_k= {\bf b}_k$.
The other two modes are the so-called ``pseudo-Alfv\'en'' waves. Their dispersion relation is also~$\omega=|k_z|v_A$, however, the polarisations are different. The magnetic polarisation is
normal to the wavevector (i.e.\ ${\bf b}_k \cdot {\bf k} = 0$) and lies in the plane of~${\bf k}$ and~${\bf B}_0$. If the group velocity of a pseudo-Alfv\'en wave is in the direction of the
guide field, its velocity polarisation is anti-parallel to the magnetic polarisation, i.e., ${\bf v}_k=- {\bf b}_k$. If the group velocity of such a wave is in the direction opposite to the
guide field, its velocity polarisation is parallel to the magnetic polarisation, i.e., ${\bf v}_k= {\bf b}_k$.
For those familiar with magnetoacoustic waves, it is helpful to note that the shear-Alfv\'en and pseudo-Alfv\'en linear waves are the limiting cases of the so-called ``Alfv\'en'' and
``slow'' waves, respectively, in the general picture of compressible MHD in the limit of low compressibility, that is $v_A\ll v_s$, where $v_s$ is the speed of sound. An arbitrarily small
perturbation of a uniformly magnetised fluid is a superposition of non-interacting MHD waves. In the purely incompressible case, however, both the shear-Alvf\'en and pseudo-Alfv\'en modes
become rather significant. As we show in the next section, shear-Alfv\'en and pseudo-Alfv\'en wave packets even at finite amplitudes become {\em exact} solutions of the incompressible nonlinear MHD
equations. Any perturbation of incompressible magnetized fluid can be expanded in these modes.
\subsection{Formulation}
The MHD equations take an especially simple form when written in the so-called Elsasser variables, ${\bf z}^\pm={\bf v}\pm {\bf b}$,
\begin{equation}
\left( \frac{\partial}{\partial t}\mp{\bf v}_A\cdot\nabla\right) {\bf
z}^\pm+\left({\bf z}^\mp\cdot\nabla\right){\bf z}^\pm = -\nabla P+\frac{1}{2}(\nu +\eta)\nabla^2{\bf z}^{\pm}+\frac{1}{2}(\nu -\eta)\nabla^2{\bf z}^{\mp} +{\bf f}^{\pm},
\label{mhd-elsasser}
\end{equation}
where ${\bf v}$ is the fluctuating plasma velocity, ${\bf b}$ is
the fluctuating magnetic field normalized by $\sqrt{4 \pi \rho_0}$,
${\bf v}_A={\bf B}_0/\sqrt{4\pi \rho_0}$ is the contribution from the
uniform magnetic field ${\bf B}_0$, $P=(p/\rho_0+b^2/2)$ includes the
plasma pressure $p$ and the magnetic pressure $b^2/2$, and the forces ${\bf f}^{\pm}$ mimic possible driving mechanisms. In the case of incompressible fluid, the pressure term is not an independent function, but it ensures incompressibility of the ${\bf z}^+$ and ${\bf z}^-$ fields. Turbulence can be excited in a number of ways, for example by driving velocity fluctuation (the ``dynamo-type'' driving), or by launching
colliding Alfv\'en waves, etc. The steady-state inertial interval should be independent of the details of the
driving.
In the incompressible case, it follows from
these equations that for ${\bf z}^\mp({\bf x},t) \equiv 0$, neglecting dissipation and driving, there is an exact
nonlinear solution that represents a non-dispersive wave packet propagating
along the direction $\mp {\bf v}_A$, i.e.\ ${\bf z}^\pm({\bf x},t)=F^\pm({\bf x}\pm {\bf v}_A t)$ where $F^\pm$ is an arbitrary function. A wave packet ${\bf z}^\pm$ therefore
propagates without distortion until it reaches a region where ${\bf z}^\mp$ does not vanish. Nonlinear interactions are thus solely the result
of collisions between counter-propagating Alfv\'en wave packets. These exact solutions may therefore be thought of as typical nonlinear structures in incompressible MHD turbulence, somewhat
analogous to eddies in incompressible hydrodynamic turbulence.\footnote{For other types of turbulence, e.g. rotating or stratified, there are no corresponding exact nonlinear solutions, and the interactions do not occur via collisions in this manner. Care must therefore be taken in applying the ideas and techniques of MHD turbulence more generally.} Any perturbations of the velocity and magnetic fields can be rewritten as perturbations of the Elsasser fields. The
conservation of energy and cross-helicity by the ideal incompressible MHD equations discussed earlier is equivalent to the conservation of the Elsasser energies $E^+=\int ({\bf
z}^+)^2\,d^3x$ and $E^-=\int ({\bf z}^-)^2\,d^3x$.
The energies $E^+$ and $E^-$ are independently conserved and they cascade in a turbulent state towards small scales owing to nonlinear interactions of oppositely moving $z^+$ and $z^-$
Alfv\'en packets.
Incompressible MHD turbulence therefore consists of Alfv\'en wave packets, which are distorted and split into smaller ones owing to nonlinear interaction, until their energy is converted
into heat by dissipation. However, when the amplitudes of the wave packets are small they can survive many independent interactions before getting destroyed, in which case MHD turbulence
can be considered as an ensemble of weakly interacting linear waves.
To estimate the strength of the nonlinear interaction, we need to compare the linear terms, $({\bf v}_A\cdot\nabla) {\bf z}^\pm$, which describe advection of Alfv\'en wave packets along the
guide field, with the nonlinear terms, $\left({\bf z}^\mp\cdot\nabla\right){\bf z}^\pm$, which are responsible for the distortion of wave packets and for the energy redistribution over
scales. Denote $b_{\lambda}$ as the rms magnetic fluctuation at the field-perpendicular scale $\lambda \propto 1/k_{\perp}$, and assume that the typical field-parallel wavenumber of such
fluctuations is~$k_\|$. For Alfv\'en waves, magnetic and velocity fluctuations are of the same order, $v_\lambda \sim b_\lambda$, so one estimates $({\bf v}_A\cdot\nabla) {\bf z}^\pm \sim
v_A k_\| b_\lambda$ and $\left({\bf z}^\mp\cdot\nabla\right){\bf z}^\pm \sim k_\perp b_\lambda^2$. Then the turbulence is called ``weak'' when the linear terms dominate, i.e.\
\begin{eqnarray}
k_\| v_A\gg k_{\perp} b_{\lambda},
\label{weak-turb}
\end{eqnarray}
and it is called ``strong'' otherwise. Depending on the driving force, turbulence exhibits either a weak or strong regime in a certain range of scales. In the next section we describe early
approaches to MHD turbulence that essentially treated the turbulence as weak.
\subsection{Early Approaches}
The first model of MHD turbulence was proposed by Iroshnikov (1963)
and Kraichnan (1965) who invoked the picture of interacting Alfv\'en packets propagating along a strong
large-scale magnetic field.
Iroshnikov and Kraichnan considered two wave packets of size~$\lambda$
propagating with the Alfv\'en velocity in opposite directions along a magnetic-field line.
They further assumed that the wave packets are isotropic; that is, their field-parallel scale is also~$\lambda$. Assuming that significant interaction occurs between eddies of comparable sizes\footnote{This is a standard assumption of ``locality'' of turbulence, see, e.g., (\cite{aluie_eyink10}).},
one then estimates from equations~(\ref{mhd-elsasser}) that during one collision, that is one crossing time $\lambda/v_A$, the
distortion of each wave packet is $\delta v_{\lambda}\sim ( v_{\lambda}^2/\lambda)(\lambda/v_A)$.
The distortions add up randomly, therefore, each wave packet will be distorted relatively strongly after
$N\sim (v_{\lambda}/\delta v_{\lambda})^2\sim (v_A/v_{\lambda})^2$ collisions with uncorrelated wave packets moving in the opposite direction. The energy transfer time is, therefore,
\begin{eqnarray}
\tau_{IK}(\lambda)\sim N\lambda/v_A\sim \lambda v_A/v_{\lambda}^2= \lambda/v_{\lambda}(v_A/v_{\lambda}).
\label{tauik}
\end{eqnarray}
It is important that in the Iroshnikov-Kraichnan interpretation,
a wave packet has to experience {\em many} uncorrelated
interactions with oppositely moving wave packets (since $\tau_{IK}\gg \lambda/v_A$) before its
energy is transferred to a smaller scale. The requirement of constant energy flux over
scales $\epsilon = v^2_{\lambda}/\tau_{IK}(\lambda)={\sf const}$ immediately
leads to the scaling of the fluctuating fields $v_{\lambda}\propto b_{\lambda}\propto
\lambda^{1/4}$, which results in the energy spectrum (the
Iroshnikov-Kraichnan spectrum),
\begin{eqnarray}
E_{IK}(k)\sim |v_k|^2k^2\propto k^{-3/2}.
\label{eik}
\end{eqnarray}
Substituting the derived scaling of the fluctuating fields into formula~(\ref{weak-turb}) one concludes that, since $k_\|\sim k_\perp \sim 1/\lambda$, turbulence
becomes progressively {\em weaker} as the scale of the fluctuations decreases. Therefore, the Iroshnikov-Kraichnan picture is the picture of weak MHD turbulence.
Iroshnikov and Kraichnan did not consider anisotropy of the
spectrum, so~$E_{IK}(k)$ was assumed to be
three-dimensional and isotropic. As discussed earlier, over the years the assumption of isotropy was shown to be incorrect. The phenomenological picture of anisotropic
weak MHD turbulence was proposed by Ng \& Bhattacharjee (1996) and
Goldreich \& Sridhar (1997), and it was put on more formal
mathematical ground by Galtier {\it et al.} (2000). This theory is
discussed in the next section. \\
\subsection{Weak turbulence}
\label{weak_turbulence}
When the
condition~(\ref{weak-turb}) is satisfied, one may assume that
turbulence consists of weakly interacting shear-Alfv\'en and pseudo-Alfv\'en
waves. The main results of the weak turbulence theory can be explained by using the following dimensional arguments, which were proposed by Ng \& Bhattacharjee (1996) and
Goldreich \& Sridhar (1997) before the more rigorous treatment by Galtier {\it et al.} (2000) was developed. First, let us note that two Alfv\'en waves with wavevectors ${\bf k}_1$ and ${\bf k}_2$ interact with
a third one if the following resonance conditions are satisfied: $\omega(k)=\omega^+(k_1)+\omega^-(k_2)$ and ${\bf k}={\bf k}_1+{\bf k}_2$. Only the waves propagating in opposite directions
along the guide field interact with each other, therefore, $k_{1\|}$ and $k_{2\|}$ should have opposite signs. Since the Alfv\'en waves have a linear dispersion relation, $\omega=|k_\| |v_A$,
the solution of the resonance equations is nontrivial only if either $k_{1\|}=0$ or $k_{2\|}=0$ (\cite{shebalin83}). The turbulence dynamics therefore depend on the fluctuation spectrum at $k_\|=0$. As the
Alfv\'en waves interact with the fluctuations at $k_\|=0$, the~$k_\|$ components of their wavevectors do not change, and the energy of waves with given $k_\|$ cascade in the direction of
large~$k_\perp$.
In the region $k_\perp\gg k_\|$, the polarization of the pseudo-Alfv\'en waves is almost parallel to the guide field ${\bf B}_0$. Therefore, the pseudo-Alfv\'en modes (${\bf z}_p$) influence
the shear-Alfv\'en ones (${\bf z}_s$) through the term $({\bf z}^\pm_p\cdot \nabla){\bf z}^{\mp}_s \sim z^\pm_p k_\| z^{\mp}_s$, whilst the shear-Alfv\'en modes, whose polarization is
normal to the guide field, interact with each other as $({\bf z}^\pm_{s}\cdot\nabla){\bf z}^{\mp}_s\sim z^{\pm}_sk_\perp z^{\mp}_s$. Since $k_\|$ does not change in the energy cascade, in
the inertial interval we have $k_\perp \gg k_\|$, and we obtain the important property of MHD turbulence that the dynamics of the shear-Alfv\'en modes decouple from the dynamics of the
pseudo-Alfv\'en ones (\cite{goldreich_s95}). Pseudo-Alfv\'en modes are passively advected by the shear-Alfv\'en ones. The spectrum of a passive scalar is the same as the
spectrum of the advecting velocity field (\cite{batchelor:1959}), therefore the spectra of pseudo- and shear-Alfv\'en waves should be the same.
These spectra can be derived from dimensional arguments similar to those of Iroshnikov and Kraichnan, if one takes into account the anisotropy of wave packets with respect to the guide
field. As before, denote the field-perpendicular scale of the interacting wave packets as~$\lambda$, and their field-parallel scale as~$l$; but now the field-parallel scale~$l$ does not
change during interactions. The collision time (crossing time) is now given by~$l/v_A$. During one collision, the counter-propagating packets get deformed by $\delta v_\lambda \sim
(v_\lambda^2/\lambda)(l/v_A)$. The number of uncorrelated interactions before the wave packets get destroyed is $N\sim (v_\lambda/\delta v_\lambda)^2\sim \lambda^2 v_A^2/(l^2 v_\lambda^2)$,
and the energy cascade time is $\tau_w\sim Nl/v_A$. From the condition of constant energy flux $\epsilon=v_\lambda^2/\tau_w={\rm const}$, one derives $v_{\lambda}\propto b_\lambda\propto
\lambda^{1/2}$, and the field-perpendicular energy spectrum
\begin{eqnarray}
E(k_\perp)\propto k_\perp^{-2},
\label{weak-spectrum}
\end{eqnarray}
where the two-dimensional Fourier transform is used in the $(x,y)$ plane in order to calculate the spectrum.
These results can be put on a more rigorous mathematical ground by using the approach of weak turbulence theory. The basic assumption of this theory is that
in the absence of nonlinear interaction, the waves have random phases, and that
the Gaussian rule can be applied to express their higher order
correlation functions in terms of the second-order ones.\footnote{Many
papers contributed over the years to the development of fundamental
ideas on MHD turbulence (see e.g., the reviews
in~\cite{biskamp:2003,ng03}). The general methods of weak turbulence theory
have been extensively reviewed (\cite{zakharov,newell}).}
A perturbative theory of weak MHD turbulence has also been developed (\cite{galtier_nnp00}, see also \cite{galtier_nnp02}).
By expanding the Elsasser form of the MHD eq.~(\ref{mhd-elsasser}) up
to the second order in the nonlinear interaction and using the
Gaussian rule to split the fourth-order correlators, the authors
derived a closed system of kinetic equations governing
the wave energy spectra\footnote{It should be noted that the Gaussian assumption is, in fact, not necessary to derive the kinetic equatons. The same closure for the long time behavior of
the spectral cumulants can be derived without a priori assumptions on the statistic of the process, e.g., Newel {\em et~al.} (2001), Elskens \& Escande
(2003).}.
These equations confirm the important fact
that wave energy cascades in the Fourier
space in the direction of large~$k_{\perp}$, and the universal spectrum
of wave turbulence is established in the region~$k_{\perp}\gg k_{\|}$. The equations also demonstrate that
in this region the dynamics of the shear-Alfv\'en waves decouple
from the dynamics of the pseudo-Alfv\'en
waves, and the pseudo-Alfv\'en waves are passively advected by the
shear-Alfv\'en ones, in agreement with the previous qualitative
consideration. The main objects of interest in this theory are the correlation functions of the shear-Alfv\'en waves, $\langle {\bf z}^{\pm}({\bf k}, t){\bf z}^\pm({\bf k}', t) \rangle
=e^{\pm}({\bf k}, t)\delta({\bf k}+{\bf k}')$. In addition, it was assumed that the waves propagating in the opposite directions are not correlated, that is,
$\langle {\bf z}^{\pm}({\bf k}, t){\bf z}^\mp({\bf k}', t) \rangle =0$. The kinetic equations for the energy spectra $e^{\pm}({\bf k}, t)$ then have the form:
\begin{eqnarray}
\partial_t e^{\pm}({\bf k})= \int M_{{\bf k}, {\bf p} {\bf q}}
e^{\mp}({\bf q}) [e^{\pm}({\bf p})-e^{\pm}({\bf k})]\delta(q_\|)
d_{\bf k,pq},
\label{galtier-eq}
\end{eqnarray}
where the interaction
kernel is given by
\begin{eqnarray}
M_{{\bf k}, {\bf p}{\bf q}}=({\pi}/{v_A}){({\bf
k}_{\perp}\times{\bf q}_{\perp})^2({\bf k}_{\perp}\cdot {\bf
p}_{\perp})^2}/({k_{\perp}^2 q_{\perp}^2 p_{\perp}^2}),
\end{eqnarray}
and the shorthand notation $d_{\bf k,pq}
\equiv \delta({\bf k}-{\bf p}-{\bf q})\,d^3p\,d^3q $ is adopted. It is also customary to
use the phase-volume compensated energy spectrum calculated as $E^{\pm}({\bf k}, t)dk_{\|}\,dk_{\perp}= e^{\pm}({\bf k},
t)2\pi k_{\perp}dk_{\|}\,dk_{\perp}$. In this section we consider only statistically balanced turbulence, that is, we assume $E^+=E^-$. The balanced stationary
solution of equation~(\ref{galtier-eq}) was found
(\cite{galtier_nnp00}) to have
the general form~$E^{\pm}({\bf k})=g(k_{\|})k_{\perp}^{-2}$, where
$g(k_{\|})$ is an arbitrary function that is smooth at~$k_{\|}=0$.
It should be noted, however, that the derivation of~(\ref{galtier-eq}) based
on the weak interaction approximation is not rigorous.
It follows from the wave resonance condition, and as is evident from equation~(\ref{galtier-eq}),
that only the $q_\|=0$~components of the energy spectrum $e({\bf q})$ are
responsible for the energy transfer. However, if we apply equation~(\ref{galtier-eq}) to
these dynamically important
components themselves, that is, if we set $k_{\|}=0$ in (\ref{galtier-eq}),
we observe an inconsistency. Indeed, the perturbative approach implies
that the linear
frequencies of the waves are much larger than the
frequency of
their nonlinear
interaction. The nonlinear interaction
in~(\ref{galtier-eq}) remains nonzero as $k_{\|}\to 0$ while the
linear frequency of
the corresponding Alfv\'en modes, $\omega_k=k_{\|}v_A$, vanishes. Therefore, as shown
by Galtier {\it et al.} (2000), an additional assumption that goes beyond the theory of weak turbulence, should be made. Namely, one has to assume the smoothness of the
function $g(k_\|)$ at $k_\| = 0$; this assumption is crucial for deriving
the spectrum $E({\bf k}) \propto k_{\perp}^{-2}$ since, according to (\ref{galtier-eq}) the wave dynamics essentially depend on the energy spectrum at~$k_\|\to 0$.
A definitive numerical verification of such a spectrum
therefore seems desirable. It is however quite difficult to perform direct numerical simulations of weak MHD turbulence based on equation~(\ref{mhd-elsasser}).
The major problem
faced by such simulations
is to satisfy simultaneously the two conditions,
$k_\perp\gg k_{\|}$, which is necessary to reach the universal regime where the dynamically unimportant pseudo-Alfv\'en mode decouples, and $k_{\|}B_0\gg k_\perp b_{\lambda}$, which is the
condition of weak turbulence. These two conditions are hard to achieve with present-day computing power.
\begin{figure}
\centering
\includegraphics[width=1.0\textwidth]{balanced_weak2.pdf}
\caption{Compensated spectra of $E^+$ (solid line) and $E^-$ (dash
line) in numerical simulations of weak MHD turbulence based on
reduced MHD equations, with numerical resolution $1024^2\times 256$,
and Reynolds number ${Re=Rm}=6000$ (from \cite{boldyrev_p09}).}
\label{balanced_weak}
\end{figure}
The first numerical tests (\cite{bhattacharjee_n01} and \cite{ng03}), used a scattering
model based on MHD equations expanded up to the second order in
nonlinear interaction rather than the full MHD equations. Integration of such a model did reproduce
the~`$-2$ exponent'. Recently, direct numerical simulations of the so-called reduced MHD (RMHD) equations (see later), which explicitly neglect pseudo-Alfv\'en fluctuations and present a good
approximation of the full MHD equations in the case of strong guide
field, were conducted (\cite{perez_b08,boldyrev_p09}); more details on numerical simulations are given in Section~(\ref{numerical_frameworks}). They also confirmed the $-2$ exponent, see Figure~(\ref{balanced_weak})
Weak turbulence theory predicts that as $k_\perp$ increases, the turbulence should eventually become strong. Indeed, owing to the obtained scaling of turbulent fluctuations,
$b_{\lambda}\propto \lambda^{1/2}$, the linear terms in equation~(\ref{mhd-elsasser}) scale as $({\bf v}_A\cdot\nabla){\bf z}^{\pm}\sim v_Ab_\lambda/l\propto \lambda^{1/2}$, whilst the nonlinear
ones are independent of $\lambda$ as $({\bf z}^\mp\cdot \nabla){\bf z}^\pm \sim b_\lambda^2/\lambda \propto {\rm const}$. One therefore observes that the condition of weak turbulence
(\ref{weak-turb}) should break at small enough~$\lambda$, which means that at large $k_\perp$ turbulence becomes strong. In the next section we describe strong MHD turbulence.
\subsection{Strong turbulence}
\label{strong_turbulence}
As noted at the end of the previous section, weak MHD turbulence eventually becomes strong as the $k_{\perp}$ increases so that condition (\ref{weak-turb}) is no longer satisfied. If this is the case then one can argue that the linear and nonlinear terms should be
approximately balanced at all scales; in equation~(\ref{weak-turb}) this
would mean that ``$k_\| v_A\sim k_{\perp} b_{\lambda}$''. This is the
so-called critical balance condition (\cite{goldreich_s95}). Before discussing the consequences of critical balance, let us give its more precise definition, and discuss its physical meaning.
In contrast with weak turbulence, in strong turbulence the magnetic field lines are relatively strongly bent by velocity fluctuations. The wave packets following these lines in opposite
directions are strongly distorted in only one interaction, that is, one crossing time. During such an interaction small wave packets are guided not by the mean field obtained through averaging
over the whole turbulent domain, but rather by a local guide field
whose direction is deviated from the direction of the mean field by
larger wave packets (\cite{cho_v00}). Therefore, it would be incorrect to verify the critical balance condition by using the wavenumber $k_\|$ defined through the Fourier transform in the global $z$-direction, as sometimes proposed in
the literature. Rather, the critical balance condition means that the nonlinear interaction time $\lambda/b_\lambda$ should be on the order of the linear crossing time, which can be represented as
$l/v_A$ with some length~$l$ along the {\em local} guide field, at all scales.
The physical meaning of the critical balance can then be understood
from the following causality principle (\cite{boldyrev05}). Suppose that owing to nonlinear interaction the wave packets
are deformed on the time scale $\tau_N\sim \lambda/v_{\lambda}$. During this time, the information about the deformation cannot propagate along the guide field further than a distance
$l\sim v_A\tau_N$, and the fluctuations cannot be correlated at a larger distance along the guide field. This coincides with the statement of the critical balance. We also note that the
condition of critical balance in the GS picture has a useful geometric property. Noting that the individual magnetic field lines in an eddy of size $\lambda$ are locally deviated by the small
angle $b_\lambda/v_A$, one derives that the Alfv\'en wave packet of length $l$ displaces the magnetic field lines
in the field perpendicular direction by a distance~$\xi\sim b_{\lambda}l/v_A $. In the GS picture this displacement happens to be equal to the perpendicular wave-packet size~$\lambda$.
As a consequence of critical balance, oppositely moving Alfv\'en wave-packets are significantly deformed during only one interaction (one crossing time). This is in contrast with weak
turbulence discussed earlier, where a large number of crossing times was required to deform a wave packet. The nonlinear interaction time therefore assumes the Kolmogorov form $\tau_{GS}\sim
\lambda/v_\lambda$, and the energy spectrum attains the Kolmogorov scaling in the field-perpendicular direction (the Goldreich-Sridhar spectrum):
\begin{eqnarray}
E_{GS}(k_\perp)\propto k_\perp^{-5/3},
\label{gs}
\end{eqnarray}
where the spectrum is calculated by using the two-dimensional Fourier transform in the $(x,y)$ plane, in analogy with the anisotropic weak turbulence spectrum~(\ref{weak-spectrum}). As an
important consequence of critical balance, strong MHD turbulence becomes progressively more anisotropic at smaller scales. Indeed, with the GS scaling $v_\lambda\propto \lambda^{1/3}$ and
the critical balance condition it follows that~$l\propto \lambda^{2/3}$, and the ``eddies'' or wave packets get progressively elongated along the local guide field as their scale decreases.
Such scale-dependent anisotropy and the scaling (\ref{gs}) seemed to find some support in early numerical simulations, which did not have strong enough guide field, i.e. they had $v_A\sim v_{rms}$, and
did not have large enough inertial interval, (see e.g.,
\cite{cho_v00,cho_lv02}). Recent high resolution direct numerical simulations with a strong guide field, $v_A\geq 5v_{rms}$, verify the strong anisotropy of the
turbulent fluctuations, supporting the argument that the original IK picture is incorrect. However, they consistently reproduce the field-perpendicular energy spectrum as close to
$E(k_\perp)\propto k_\perp^{-3/2}$, (see, e.g.,
\cite{maron_g01,mullbisgrapp:2003,muller_g05}, Mason {\it et al.} 2006, \cite{mason_cb08,perez_b08}), thus contradicting the GS model, see Figure (\ref{muller_grappin2005}).
\begin{figure}
\centering
\includegraphics[width=1.0\textwidth]{muller_grappin20052.pdf}
\caption{Compensated field-perpendicular total (solid), kinetic
(dashed), and magnetic (dotted) energy spectra. The guide field is
$B_0=5v_{rms}$, resolution $1024^2\times 256$, Reynolds number
$Re=Rm=2300$ (based on field-perpendicular fluctuations).
Dash-dotted curve: high-k part of field-parallel total energy
spectrum. Insert: comparison of the field-perpendicular energy
spectra for field-perpendicular resolutions of $512^2$ (dash-dotted)
and $1024^2$ (solid) (from \cite{muller_g05}).}
\label{muller_grappin2005}
\end{figure}
The flattening of the spectrum compared with the GS theory means that the energy transfer time becomes progressively longer than the Goldreich-Sridhar time~$\tau_{GS}(\lambda)$ at small
$\lambda$. This can happen if the nonlinear interaction in the MHD
equations is depleted by some mechanism. A theory explaining such
depletion of nonlinear interaction has recently been proposed
(\cite{boldyrev05,boldyrev06}). In addition to the
elongation of the eddies in the direction of the guiding field, it is proposed that
at each
field-perpendicular scale $\lambda$ ($\sim 1/k_{\perp}$) in the inertial range, typical
shear-Alfv\'en velocity fluctuations, ${\bf v}_{\lambda}$, and magnetic fluctuations, $\pm {\bf b}_{\lambda}$, tend
to align the directions of their polarizations in the
field-perpendicular plane, and the turbulent eddies become anisotropic in that plane. In these eddies the magnetic and velocity fluctuations change significantly in the direction almost
perpendicular to the directions of the fluctuations themselves,~${\bf v}_\lambda$ and~$\pm {\bf b}_\lambda$. This reduces the strength of the nonlinear interaction in the MHD equations by
$\theta_\lambda \ll 1$, which is the angle between the direction of the fluctuations and the direction of the gradient: $({\bf z}^{\mp}\cdot \nabla){\bf z}^{\pm}\sim v_{\lambda}^2
\theta_\lambda/\lambda$.
One can argue that the alignment and anisotropy
are stronger for smaller scales, with the alignment angle decreasing with scale
as $\theta_{\lambda}\propto \lambda^{1/4}$. This `dynamic alignment'
process progressively reduces the strength of the nonlinear interactions as the scale of the fluctuations decreases, which leads to the velocity and magnetic fluctuations $ v_{\lambda}\sim b_\lambda \propto \lambda^{1/4}$, and to the field-perpendicular energy spectrum
\begin{eqnarray}
E(k_{\perp}) \propto k_{\perp}^{-3/2}.
\end{eqnarray}
Dynamic alignment is a well-known phenomenon of MHD turbulence, (e.g.,
Dobrowolny {\it et al.} 1980, \cite{biskamp:2003}), but the term has taken on a number of meanings.
In early studies it essentially meant that {\em decaying} MHD turbulence asymptotically reaches the so-called Alfv\'enic state where either ${\bf v}({\bf x})\equiv {\bf b}({\bf x})$
or ${\bf v}({\bf x})\equiv -{\bf b}({\bf x})$, depending on initial conditions. Such behaviour is a consequence of cross-helicity conservation, (see e.g., \cite{biskamp:2003}): energy decays faster than cross-helicity, and
such selective decay leads asymptotically to alignment of magnetic and velocity fluctuations.
Regions of polarized fluctuations have also been previously detected in numerical simulations of {\em driven} turbulence (\cite{maron_g01}). The essense of the phenomenon that we discuss here is that in randomly driven MHD turbulence the fluctuations ${\bf v}_\lambda$ and $\pm {\bf b}_\lambda$ tend
to align their directions in such a way that the alignment angle
becomes {\em scale-dependent} (\cite{boldyrev05,boldyrev06,mason_cb06,boldyrev_mc09}).
\begin{figure}
\centering
\includegraphics[width=1.0\textwidth]{alignment_angle2.pdf}
\caption{The alignment angle between shear-Alfv\'en velocity and
magnetic fluctuations as a function of fluctuation scale~$r$. The
results are plotted for five independent simulations that differ by
the large-scale driving mechanisms. The solid line has the slope of
1/4, predicted by the theory (from \cite{mason_cb08}).}
\label{alignment_angle}
\end{figure}
The first numerical observations of the scale-dependent dynamic
alignment were presented in Mason {\it et al.} (2006) and Mason {\it et
al} (2008), (see also Beresnyak \& Lazarian 2006). The results are shown in Figure
(\ref{alignment_angle}). The effect was also observed in magnetized
turbulence in the solar wind (\cite{podesta09}). It has been also
argued (Boldyrev {\it et al.} 2009) that the phenomenon of scale-dependent dynamic alignment, and the corresponding energy spectrum
$k^{-3/2}_{\perp}$ are consistent with the exact relations known for
MHD turbulence (\cite{politano_p98}).
As previously discussed, there are two possibilities for dynamic alignment:
the velocity fluctuation ${\bf v}_\lambda$ can be aligned either
with ${\bf b}_\lambda$ (positive alignment) or with $-{\bf b}_\lambda$ (negative alignment). This implies that the turbulent domain can be fragmented into regions (eddies) of positive and negative
alignment. If no overall alignment
is present, that is, the total cross-helicity of turbulence is zero,
the numbers of positively and negatively aligned eddies are equal on
average. In the case of nonzero total cross-helicity those numbers may
be not equal. However, there is no
essential difference between overall balanced (zero total
cross-helicity) and imbalanced (non-zero total cross-helicity) strong
turbulence (\cite{perez_b09}). Therein it is argued that strong MHD turbulence,
whether balanced overall or not, has the characteristic property that at each scale it is locally imbalanced.
Overall, it can be viewed as a superposition of positively and negatively aligned eddies. The scaling of the turbulent energy spectrum depends on the change in degree of alignment with
scale, not on the amount of overall alignment.
\subsection{Numerical frameworks}
\label{numerical_frameworks}
In the absence of a rigorous theory, the understanding of MHD turbulence largely relies on phenomenological arguments and numerical simulations. Numerical experiments are sometimes able to
resolve controversies among different phenomenological assumptions. In this section we discuss the specific requirements for the numerical simulations imposed by the special features of MHD
turbulence. In particular, the anisotropic nature of MHD turbulence requires the optimisation of the computation domain and the driving force in order to capture the dynamics of strongly
anisotropic eddies. First, one needs to choose the forcing as to ensure that large-scale eddies, driven by the forcing, are critically balanced. Second, the simulation box should be
expanded in the direction of the guide field ($z$-direction) so as to fit the anisotropic eddies, otherwise, residual box size effects can spoil the inertial interval if the Reynolds number
if not large enough.
Numerical simulations, which have an inertial interval of a limited extent, demonstrate that when the guide field is not strong enough, $v_A\leq v_{rms}$, the observed turbulence spectrum
is not very different from the Kolmogorov spectrum (e.g.,
\cite{cho_v00,cho_lv02,mason_cb06}). As~$v_A$ exceeds~$\sim 3
v_{rms}$, the guide field becomes important and the spectral exponent
changes to~$-3/2$, (e.g., \cite{mullbisgrapp:2003}, Mason {\it et al.} 2006).
In the latter case, the simulation box should have the aspect ratio
$L_\|/L_\perp \sim v_A/v_{rms}$. Current state-of-the-art
high-resolution direct numerical simulations of MHD turbulence
(\cite{muller_g05,perez_b09,perez_b10}) discuss in detail the aforementioned effects of anisotropy related to the strong guide field.
We now demonstrate that anisotropic MHD turbulence can, in fact, be effectively simulated using simplified MHD equations. An obvious simplification stems from the fact that in the case of
strong guide field the gradients of fluctuating fields are much smaller along the guide field than across the field. Such a system of equations, the so-called Reduced MHD (RMHD) equations,
was derived in context of fusion devices (\cite{strauss76,kadomtsev_p74}, see also \cite{biskamp:2003}). In this system the following self-consistent ordering is used $L_\perp/L_\| \sim
b_\perp/B_0\sim b_\|/b_\perp \ll 1$, where $L_\|$ and $L_\perp$ are typical scales of magnetic perturbations, i.e., large scales of turbulent fluctuations in our case. As noticed by Goldreich \& Sridhar (1995) this
is precisely the ordering following from the critical balance condition for strong MHD turbulence. Therefore, the Reduced MHD system is suitable for studying strong turbulence. The Reduced
MHD equations have the form:
\begin{eqnarray}
& \left( \frac{\partial}{\partial t} \mp {\bf v}_A\cdot \nabla \right) {\tilde
{\bf z}}^{\pm}+({\tilde {\bf z}}^{\mp}\cdot \nabla){\tilde {\bf
z}}^{\pm}=-\nabla_{\perp} P +\nonumber \\
& +\frac{1}{2}(\nu +\eta)\nabla^2{\bf z}^{\pm}+\frac{1}{2}(\nu -\eta)\nabla^2{\bf z}^{\mp} +{\tilde {\bf f}}_\perp ,
\label{rmhd}
\end{eqnarray}
where the fluctuating fields and the force have only two vector components,
e.g., ${\tilde {\bf z}}^{\pm}=\{{\tilde z}^{\pm}_1, {\tilde z}^{\pm}_2, 0 \}
$, but depend on all three spatial coordinates. Since both the velocity and the magnetic fields are divergence-free, they can be represented through the scalar potentials, ${\bf v}={\hat
{\bf z}}\times \nabla \phi$ and ${\bf b}={\hat {\bf z}}\times \nabla \psi$, where ${\hat {\bf z}}$ is the unit vector in the z-direction. Then, the system~(\ref{rmhd}) yields two scalar
equations,
\begin{eqnarray}
{}& \partial_t \omega + ({\bf v}\cdot \nabla) \omega -({\bf b}\cdot \nabla)j =B_0\partial_z j +\nu \nabla^2\omega +f_\omega , \nonumber \\
{}& \partial_t \psi +({\bf v}\cdot \nabla) \psi=B_0 \partial_z \phi +\eta \nabla^2 \psi +f_\psi,
\label{rmhd2}
\end{eqnarray}
where $j=\nabla^2_\perp \psi$ is the current density and $\omega=\nabla^2_\perp \phi$ is the vorticity. Note that it would be incorrect to suggest that system (\ref{rmhd2}) describes
quasi-two-dimensional turbulence, since the linear terms describing the field-parallel dynamics, that is, the terms containing $\partial_z$, are of the same order as the nonlinear terms.
The system (\ref{rmhd2}) is the best known form of the RMHD equations. The less common symmetric Elsasser form (\ref{rmhd}), on the other hand, has advantages for analytic study, for example, it allows
one to derive relations analogous to those by Politano \& Pouquet (1998) for the case of anisotropic MHD turbulence (see \cite{perez_b08}). Both full and RMHD systems can be used for
numerical simulations of strong MHD turbulence, however, the reduced MHD system has half as many independent variables as the full MHD equations, allowing one to speed up the numerical
computations by a factor of two.
Recently, it has been demonstrated that the system (\ref{rmhd}) can also be effectively
used for numerical simulations of the universal regime of {\em weak} MHD turbulence (Perez \& Boldyrev 2008). This is somewhat surprising and requires an explanation. As we discussed in
Section~\ref{weak_turbulence}, the weak turbulence spectrum becomes universal at $k_\perp \gg k_\|$, when the pseudo-Alfv\'en modes decouple from the cascade dynamics. Simultaneously, one needs to satisfy the
condition of weak turbulence (\ref{weak-turb}). In order to satisfy
both conditions, one needs quite a low ratio of $b/B_0$, which implies quite a short Alfv\'en time and an increased
computational cost. It turn out that the first condition can be relaxed if one uses the system where the pseudo-Alfv\'en modes are explicitly removed. This system is obtained from the full MHD system (\ref{mhd-elsasser}) if one sets $z^\pm_\|\equiv 0$. The resulting restricted system then formally coincides with (\ref{rmhd}), with the exception that it now can be used out of the region of validity of the RMHD equations, that is, $k_\perp \gg k_\|$. The restricted system explicitly neglects the pseudo-Alfv\'en modes, and,
therefore, it describes the universal regime of weak MHD turbulence of shear-Alfv\'en waves as long as the driving force ensures the weak turbulence condition~(\ref{weak-turb}). The condition~$k_\perp \gg
k_\|$ is therefore not required.
To support this argument, one can demonstrate that the weak turbulence derivation based on the reduced system (\ref{rmhd}) leads to exactly the same kinetic equations~(\ref{galtier-eq}), as
the full MHD system~(\cite{galtier_nnp02,galtier_c06}, Perez \& Boldyrev 2008, Galtier 2009). Moreover, direct numerical simulations of system~(\ref{rmhd}) with the broad $k_\|$-spectrum of the driving force,
necessary to ensure the weak turbulence condition~(\ref{weak-turb}), reproduce the energy spectrum of weak turbulence~$k_\perp^{-2}$ (Perez \& Boldyrev 2008), see Figure~(\ref{balanced_weak}). To date, this has been the only
available direct numerical study of the universal regime of weak MHD turbulence. We conclude that the reduced MHD system provides an effective framework for simulating the universal
regimes of MHD turbulence, providing the forcing is chosen accordingly. Depending on the $k_\|$ spectral width of the forcing, the driven turbulence is weak if the fluctuations satisfy the
inequality~(\ref{weak-turb}), and it is strong if approximate equality
holds in~(\ref{weak-turb}) instead. Further examples of numerical
simulations of various regimes of MHD turbulence in the RMHD framework
are available (e.g., \cite{oughton_etal04,rappazzo_etal2007,rappazzo_etal2008}, Dmitruk {\em et~al.} 2003).
It should be reiterated that even if the turbulence is driven in a ``weak'' fashion, it remains weak only for a certain range of scales, and it becomes strong below the field-perpendicular
scales at which the condition (\ref{weak-turb}) breaks down. With rapidly increasing capabilities of numerical simulations it should become possible to reproduce a large enough inertial
interval, and to observe the predicted transition from weak to strong MHD turbulence.
\subsection{Unbalanced turbulence}
In the preceding sections we assumed that the MHD turbulence was statistically balanced, that is, the energies of counter-propagating Elsasser modes $E^+$ and $E^-$ were the same. This,
however, is by no means guaranteed, since the Elsasser energies are independently conserved by the ideal MHD equations. In nature and in the laboratory (e.g., the solar wind, interstellar
medium, or fusion devices) MHD turbulence is typically unbalanced as it is generated by spatially localized sources or instabilities, so that the energy has a preferred direction of
propagation. Moreover, there is a reason to believe that imbalance is an inherent fundamental property of MHD turbulence. As argued earlier, numerical simulations and phenomenological
arguments indicate that even when the turbulence is balanced overall, it is unbalanced locally, creating patches of positive and negative cross-helicity (e.g., \cite{matthaeus08,perez_b09,boldyrev_mc09}). We have already encountered
this phenomenon when discussing dynamic alignment in Section~\ref{strong_turbulence}. Unbalanced MHD turbulence is also called ``cross-helical,'' as the normalized cross helicity,
$H^C/E=\frac{1}{2}(E^+-E^-)/(E^++E^-)$, is a natural measure of the imbalance.
In unbalanced MHD turbulence the energies of counter-propagating Elsasser modes are not equal, and, {\em a priori} might not have same scalings. This raises an interesting question as to whether
the MHD turbulence is universal and scale-invariant. Indeed, if the energy spectra in the unbalanced domains are different, the overall energy spectrum does not have to be scale-invariant
and universal, but rather may depend on the driving and dissipation mechanisms. Below we consider the cases of strong and weak unbalanced turbulence separately. \\
{\em Strong unbalanced turbulence}. Phenomenological treatments of strong unbalanced MHD turbulence are complicated by the fact one can formally construct two time scales for nonlinear energy
transfer. The MHD system (\ref{mhd-elsasser}) suggests that the times of nonlinear deformation of the $z^\pm$ packets are $\tau^\pm\sim \lambda/z_\lambda^\mp$. These time scales are
essentially different in the unbalanced case, which is hard to reconcile with the assumption that most effectively interacting counter-propagating eddies have comparable field-parallel and
field-perpendicular dimensions. Several phenomenological models attempting to accommodate this time difference have been proposed recently. These, however, have led to conflicting
predictions for the turbulent spectra of~$E^+$ and~$E^-$ (\cite{lithwick_gs07,beresnyak_l08,chandran08,perez_b09}).
A possible resolution has been proposed (\cite{perez_b09}). Here the phenomenon of scale-dependent dynamic alignment was invoked to estimate the interaction times. In an unbalanced eddy,
the field-perpendicular fluctuations of ${\bf v}_\lambda$ and ${\bf b}_\lambda$ are aligned in the field-perpendicular plane within a small angle $\theta_\lambda$. In the phenomenology of
scale-dependent dynamic alignment, these fluctuations are almost normal to the direction of their gradient, which is also in the field-perpendicular plane. In the case of strong imbalance,
$z^+\gg z^-$, ${\bf z}^+$ and ${\bf z}^-$ then form different angles with the direction of the gradient, and a geometric constraint requires that $z^+_\lambda\theta^+_\lambda \sim
z^-_\lambda\theta^-_\lambda$. In the aligned case the nonlinear interaction time is increased by the alignment angle, $\tau^\pm\sim \lambda/z^{\mp}_\lambda\theta_\lambda^\mp$, leading to the
conclusion that the nonlinear interaction times are the {\em same} for both $z^+$ and $z^-$ packets. Therefore, $z_\lambda^+$ has the same scaling as $z_\lambda^-$, and $\theta_\lambda^+$ has the same scaling as $\theta_\lambda^-$.
As a result, in strong unbalanced MHD turbulence the spectra of the two sets of modes should have the same scaling, $E^+(k_\perp)\sim E^-(k_\perp)\propto k_\perp^{-3/2}$, but different
amplitudes. This result is consistent with the picture that overall balanced MHD turbulence consists of regions of local imbalance of various strengths. In each of the unbalanced regions the
fluctuations are dynamically aligned and the discussed phenomenology applies. The scaling of the spectrum of strong MHD turbulence is therefore universal, and it does not depend on the
degree of overall imbalance of the turbulence. These results seem to be supported by numerical simulations (\cite{perez_b09,perez_b10}).
\begin{figure}
\centering
\includegraphics[width=1.0\textwidth]{strong-balanced-imbalanced2.pdf}
\caption{Compensated field-perpendicular spectra of strong balanced
MHD turbulence (A1-A2), and strong unbalanced MHD turbulence (B1-B3)
in numerical simulations of reduced MHD system. Runs A1, A2 have
Reynolds numbers $Re=Rm=2400$ and $6000$, correspondingly. Runs B1,
B2, and B3 have Reynolds numbers $900$, $2200$, and $5600$. In runs
B1-B3, the slopes of the Elsasser spectra become progressively
more parallel and close to -3/2 as the Reynolds number increases (from \cite{perez_b10}).}
\label{strong-balanced-imbalanced}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=1.0\textwidth]{evb_kpar2.pdf}
\caption{Field-parallel magnetic and kinetic energy spectra at given $k_\perp=5$ in the inertial interval. The condensate of the residual energy appears at small $k_\|$. The simulations correspond to moderately imbalanced weak MHD turbulence $z^+/z^-\sim 2$, resolution $512^2\times 256$, Reynolds number $Re=Rm=2500$ (from \cite{boldyrev_p09}).}
\label{evb_kpar}
\end{figure}
\begin{figure}
\centering
\hskip-5mm
\includegraphics[width=1.0\textwidth]{pinning2.pdf}
\caption{Compensated field-perpendicular spectra of the Elsasser
fields in unbalanced weak MHD turbulence simulations based on
reduced MHD equations. Resolution $1024^2\times 256$, Reynolds
numbers $Re=Rm=2000$ and $4500$. The spectra are anchored at large
scales and pinned at the dissipation scales. As the Reynolds number
increases, the spectral slopes should become progressively more
parallel thus converging to~-2 (from \cite{boldyrev_p09}).}
\label{pinning}
\end{figure}
To conclude this section we note that the above consideration allows one to predict the scalings of the Elsasser fields, however, it does not allow one to specify their amplitudes. Since the $E^+$ energy is mostly concentrated in positively aligned eddies, while $E^-$ energy in negatively aligned ones, the amplitudes of the $E^{\pm}$ spectra averaged over the whole turbulent domain depend on the numbers of eddies of each kind. In particular, for each eddy one can estimate $(z^+_\lambda /z^-_\lambda)^2 \sim \epsilon^+ / \epsilon^-$, however, this relation should not generally hold for the quantities $\langle (z^{\pm}_\lambda)^2 \rangle $ and $\langle \epsilon^{\pm} \rangle $ averaged over the whole turbulent domain (e.g., \cite{perez_b10b}). In principle, such an average may be not universal, but rather may depend on the structure of the large-scale driving.
A possible refinement of the theory, which takes into account different populations of positively and negatively aligned eddies has been proposed (\cite{podesta_b09}).\\
{\em Weak unbalanced turbulence}. Weak unbalanced MHD turbulence allows for a more detailed consideration. A good starting point is provided by the equations~(\ref{galtier-eq}) describing
the evolution of the Elsasser energies (\cite{galtier_nnp00}). We recall that in deriving these equations one assumes that in the zeroth-order approximation, only the
correlation functions $e^{\pm}({\bf k}) \propto \langle {\bf z}^\pm ({\bf k}) \cdot {\bf z}^\pm (-{\bf k}) \rangle $ are non-zero. So far we have discussed the ``balanced'' solution of this system,
$e^+({\bf k})= e^-({\bf k}) \propto k_\perp^{-3}$. The authors realized that the system can also describe unbalanced MHD turbulence. They noticed that the system
has a broader range of steady solutions: the right hand side integrals of equation~(\ref{galtier-eq}) vanish for a one-parameter family of power-like solutions
\begin{eqnarray}
e^{\pm}({\bf k})=g^{\pm}(k_\|)k_{\perp}^{-3\pm \alpha},
\label{spectra}
\end{eqnarray}
with arbitrary functions $g^{\pm}(k_\|)$, which are smooth at $k_\|= 0$, and $-1<\alpha<1$. For $\alpha \neq 0$ these solutions correspond to unbalanced MHD turbulence. The different energy spectra correspond to different
energy fluxes over scales, which in a steady state are equal to the rates of energy
provided by the large-scale forcing to the Elsasser fields; we denote these rates as~$\epsilon^{\pm}$. One can demonstrate directly from eq.~(\ref{galtier-eq}) that $\alpha$ is uniquely
determined if the ratio of the fluxes $\epsilon^+/\epsilon^-$ is specified. One can further show that the solution with the steeper spectrum corresponds to the larger energy flux, and vice
versa~(\cite{galtier_nnp00,lithwick_g03}).
Once the energy fluxes provided by the large-scale forcing are
specified, the slopes of the energy spectra are specified but their
amplitudes are not. Fully to remove the degeneracy,
one further argues that at the dissipation scale the balance should be
restored, that is, $e^+(k)$ should converge to $e^-(k)$. This
so-called ``pinning'' effect was first pointed out in Grappin {\it et
al} (1983), and has been elaborated upon in greater detail (see
e.g. \cite{galtier_nnp00,lithwick_g03,chandran08}). The pinning can be physically understood, if one notes that the $e^+(k)$ and $e^-(k)$ energy
spectra cannot intersect in the inertial interval, as this would contradict the universality of the turbulence. The alignment of velocity and magnetic fluctuations, preserved by the nonlinear
terms, can be broken only by the dissipation. This pinning effect is indeed observed in the simulations presented below.
According to the above picture, if the rates of energy supply are fixed, then the {\em slopes} of the energy spectra $e^{\pm}(k)$ are fixed as well. If the dissipation scale is now changed,
the {\em amplitudes} of the spectra should change so as to maintain the specified slopes and to make them converge at the dissipation scale. This conclusion, although consistent with
equations (\ref{galtier-eq}), seems however to be at odds with common intuition about turbulent systems, which suggests that small-scale dissipation should not significantly affect
large-scale fields subject to the same large-scale driving.
This seeming contradiction has recently been addressed (\cite{boldyrev_p09}). It was proposed that driven weak MHD turbulence generates a residual energy condensate
\begin{eqnarray}
\langle {\bf z}^+({\bf
k})\cdot {\bf z}^{-}(-{\bf k})\rangle = v^2-b^2 \neq 0 \quad \mbox{at} \,\, k_\|=0.
\label{condensate}
\end{eqnarray}
This condensate has been assumed to be zero in the standard
derivations (Ng \& Bhattacharjee 1996,\cite{goldreich_s97,galtier_nnp00}),
however, this assumption is not necessary and apparently incorrect in the unbalanced case. The presence of the condensate can be explained as follows. Alfv\'en wave fluctuations at $k_\|
\neq 0$ obey ${\bf v}= \pm {\bf b}$, in which case the residual energy vanishes. However, at $k_\|=0$ fluctuations are not waves, and the Alfv\'enic condition should not be necessarily
satisfied. The presence of the condensate (\ref{condensate}) implies that magnetic and kinetic energies are not in equipartition at $k_\|=0$.
It can be argued that the generation of the
condensate is a consequence of the breakdown of the mirror symmetry in unbalanced turbulence. Indeed, non-balanced MHD turbulence is not mirror-invariant, as it possesses non-zero
cross-helicity, i.e. $H^c=\int ({\bf v}\cdot {\bf b})d^3x=(E^+-E^-)/4 \neq 0$. Non-mirror invariant turbulence can also possess nonzero magnetic helicity, $H^m=\int ({\bf A}\cdot {\bf B})\, d^3x \neq 0$. Helical magnetic fields should not necessarily be in equipartition with the velocity field, as is obvious from the fact that the corresponding Lorentz force may be equal to zero (consider, for example, a Beltrami field ${\bf B}=\alpha \nabla \times {\bf B}$, which is maximally helical, but exerts no force on the velocity field). In the case of weak MHD turbulence, the only region of phase space where the equipartition between the magnetic and velocity fluctuations may be broken is the region $k_\|=0$, where such fluctuations are not Alfv\'en waves. In the case of a uniform guide field ${\bf B}_0$, which is relevant for the numerical set-up, these arguments can have further support if one notes that the magnetic helicity of {\em fluctuations} is not conserved, rather, it is generated by the term $\int [{\bf z}^+\times {\bf z}^-]_\|\,d^3 x$, (see e.g., \cite{galtier_nnp00}). Interestingly, the magnetic field of the condensate is generated by the $k_\|=0$ component of the same term, $[{\bf z}^+\times {\bf z}^-]_\|$. Such a condensate is indeed observed in the numerics, see Figure~(\ref{evb_kpar}).
The conclusion of this analysis are the following. When the turbulence is balanced, the energy spectra are $E^{\pm}(k_{\perp})\propto k_{\perp}^{-2}$, in agreement with the analytic
prediction (Ng \& Bhattacharjee 1996, \cite{goldreich_s97,galtier_nnp00}). In the balanced case the evolution of $E^{\pm}$ fields is not affected by the condensate. In the non-balanced case the
interaction with the condensate becomes essential, and the universal power-law spectra may not exist in an inertial interval of limited extent. Both spectra $E^{\pm}(k_{\perp})$
have the large-scale amplitudes fully specified by the external forcing, and they converge at the dissipation scale. As the dissipation scale decreases, however, the spectral scalings (but
not amplitudes) approach each other at large~$k_{\perp}$ such that the universal scaling $k_{\perp}^{-2}$ is recovered for both the $E^{\pm}(k_{\perp})$ spectra as $k_{\perp}\to \infty$, see Figure~(\ref{pinning}).
Finally, we would like to point out that unbalanced MHD turbulence is numerically more challenging than balanced turbulence. Indeed, in the case $z^+ \gg z^-$
the formal Reynolds number associated with the $z^+$ mode is much smaller than the Reynolds number associated with $z^-$. The stronger the imbalance the larger
the Reynolds number and the resolution required to reproduce the inertial intervals for both fields. For example, numerical simulations with the resolution of $1024^2$ points in the field-perpendicular direction
do not allow one to address turbulence with imbalance stronger than $z^+/z^-\sim 2$--$3$, (see e.g., \cite{perez_b10}).
|
\section{Introduction}\label{intro}
Driven transport and coherent control at the nanoscale are well established areas of research.
Quantum ratchets,\cite{HanggiNM11,Kemerink2011} molecular charge,\cite{RothNM10}
spin\cite{KowenhovenRMP07,WernsdorferNM08} and heat pumps,\cite{Nitzan2007,Dlott2007}
and nano-plasmonics\cite{HalasNL10} are just several examples of areas of recent developments.
Advances in optical techniques, in particular near-field optical microscopy, allow single
molecule manipulation\cite{Hildner2011} and induction of bond specific
chemistry.\cite{SeidemanACR99}
Combined with molecular junction fabrication techniques,\cite{NatelsonNL07}
optical spectroscopy methods are becoming an important observation and diagnostic tool
in molecular electronics.\cite{NatelsonNL08,Ioffe2008,NatelsonNN11}
Experimental developments led to surge of theoretical activity in the field of optically assisted
transport\cite{HanggiCP04,GalperinNitzanRatnerPRL06,CuevasPRB07,CuevasPRB08}
and optical response of molecular junctions.\cite{GalperinNitzanPRL05,GalperinNitzanJCP06,Harbola2006b,GalperinTretiakJCP08,Sukharev2010,GalperinRatnerNitzanNL09,GalperinRatnerNitzanJCP09}
In particular, Ref.~\onlinecite{GalperinNitzanRatnerPRL06} considered molecular junctions
composed of molecules with strong charge-transfer transition into their excited
state\cite{PonderMathies1983,ColvinAlivisatos1992,SmirnovBraun1998}
as a possible constituent for light-induced molecular charge pump, when change of molecular
dipole occurs along the junction axis. Consideration was done within a two-level (HOMO-LUMO)
model with ground and excited (HOMO and LUMO) states of the molecule strongly coupled to
different contacts. In junction setup optical excitation brings electron from occupied ground
to empty excited state, and asymmetry in coupling to contacts assures appearance of current.
The model was treated within non-equilibrium Green function approach, and perturbation theory
in coupling to laser field was employed.
Later Ref.~\onlinecite{FainbergNitzanPRB07} generalized the consideration of
Ref.~\onlinecite{GalperinNitzanRatnerPRL06} to strong laser fields.
Pumping optical field was treated as a classical driving force, and closed set of
EOMs for observables (electronic populations and coherences of the levels and single time exciton
correlation function) was formulated. One of the most important advances in
Ref.~\onlinecite{FainbergNitzanPRB07} was consideration of chirped laser pulses,
which allowed formulation of charge transfer between ground and excited states in terms
of Landau-Zener problem. Chirped laser pulses enable to produce complete population
inversion in molecular systems (a molecular bridge) where the well-known
$\pi$-pulse excitation\cite{All75} fails.
In realistic molecular junctions optical field driving the molecule is a {\em local field} formed by
both incident radiation and scattered response of the system (mostly plasmonic response of
metallic contacts). Another feature of molecular junctions is hybridization of states of a molecule
with those of contacts. The latter leads to non-Markov effects in response of the junction.
In this paper we generalize studies reported in Ref.~\onlinecite{FainbergNitzanPRB07}
incorporating the aforementioned effects into consideration.
Dynamics of local electromagnetic fields is simulated within the FDTD technique
for realistic geometry of a molecular junction similar to our previous
publication.\cite{Sukharev2010}
Non-Markov effects of junction response are introduced within non-equilibrium
Green functions equation-of-motion (NEGF-EOM) approach.
Structure of the paper is the following. After introducing the model in section~\ref{model},
we describe a junction geometry and numerical approach used in calculations of local
electromagnetic fields in section~\ref{optics}. Section~\ref{cur} discusses calculation of local
field-induced electron flux through the junction, and section~\ref{eom} introduces
set of NEGF-EOMs. Numerical results and discussion are given in section~\ref{res}.
Section~\ref{conclude} summarizes our findings.
\section{Model}\label{model}
A model junction consists of a molecule coupled to two metallic contacts driven by external
radiation field. The radiation is a time-dependent local electromagnetic field $E(t)$ calculated
within FDTD technique for bowtie geometry of the contacts (see section~\ref{optics} for details).
Molecule is represented by a two-level system $|1>$ and $|2>$ (HOMO and LUMO
or ground and excited states), and is placed in a `hot spot' of the local field.
Contacts $L$ and $R$ are assumed to be free charge carrier reservoirs,
each at its own equilibrium. Difference in their electrochemical
potentials defines bias applied to the junction $eV=\mu_L-\mu_R$.
Following Refs.~\onlinecite{GalperinNitzanRatnerPRL06,FainbergNitzanPRB07}
we consider two types of coupling between molecule and contacts: charge and energy transfer.
Hamiltonian of the junction is
\begin{align}
\label{H}
\hat H(t) =& \hat H_0(t) + \hat V
\\
\label{H0}
\hat H_0(t) =& \sum_{m=1,2}\varepsilon_m \hat n_m
+ \sum_{k\in \{L,R\}}\varepsilon_k\hat n_k
\\
-& \mu E(t)\left(\hat D_{12}+\hat D_{12}^\dagger\right)
\nonumber \\
\label{V}
\hat V =& \sum_{m=1,2}\sum_{k\in\{L,R\}}\left(V_{km}\hat c_k^\dagger\hat d_m +\mbox{H.c.}\right)
\nonumber \\
+& \sum_{k_1\neq k_2\in\{L,R\}}\left(V^{en}_{k_1k_2}\hat c_{k_1}^\dagger\hat c_{k_2}\hat D_{12}
+\mbox{H.c.}\right)
\end{align}
Here $\hat d_m^\dagger$ ($\hat d_m$) and $\hat c_k^\dagger$ ($\hat c_k$)
are creation (annihilation) operators of electron in level $m$ of the molecule and
in state $k$ in the contact(s), respectively, $\hat n_m=\hat d_m^\dagger\hat d_m$
is the operator of electronic population in level $m$,
$\hat D_{12}\equiv\hat d_1^\dagger\hat d_2$ is operator of molecular de-excitation
($\hat D_{21}\equiv\hat D_{12}^\dagger$), and
$\mu$ is molecular transition dipole moment. Terms on the right-hand side of (\ref{H0})
represent molecular structure (two-level system), contacts, and coupling to the driving field.
Right-hand side of Eq.(\ref{V}) introduces electron and energy transfer
between molecule and contact(s).
Eqs.~(\ref{H})-(\ref{V}) introduce the model of Ref.~\onlinecite{FainbergNitzanPRB07}
with rotating wave approximation relaxed, and with driving force treated as
a local electromagnetic field.
\begin{figure}[t]
\centering\includegraphics[width=\linewidth]{fig1}
\caption{\label{fig1}
(Color online) A sketch of local field driven molecular charge pump.
}
\end{figure}
To simulate molecules with strong charge-transfer transition with dipole moment
oriented along the junction axis below we assume that ground state (or HOMO), $|1>$,
is coupled strongly to the left contact $L$, while excited state (or LUMO) - to the right
contact $R$. Such setup works as a local field driven charge pump (see Fig.~\ref{fig1}).
Note that similar selective coupling can also be obtained for the bridge made of
a quantum dots as discussed in Refs.~\onlinecite{Bryllert02,Brandes08PRB,Fai10PRB}.
\section{Local field simulations}\label{optics}
Calculations of the local electromagnetic field dynamics are carried out utilizing FDTD
technique.\cite{TafloveBook}
Following Ref.~\onlinecite{FainbergNitzanPRB07} we assume that the incident field,
$E_{\mbox{inc}}(t)$, has the form of a linear chirped pulse
\begin{equation}
\label{chirp}
E_{\mbox{inc}}(t)=\mbox{Re}\left( \mathcal{E}_0
\exp \left( -\frac{\left(\delta^2-i\bar\mu\right)t^2 }{2}-i\omega_0t \right) \right),
\end{equation}
where $\mathcal{E}_0$ is the incident peak amplitude, $\omega_0$ is the incident frequency, and
parameters $\delta$ and $\bar\mu$ describing incident chirped pulse are given by
\begin{align}
\delta^2 =& \frac{2\tau_0^2}{\tau_0^4+4\Phi''{}^2(\omega_0)}, \\
\bar\mu =& -\frac{4\Phi''(\omega_0)}{\tau_0^4+4\Phi''{}^2(\omega_0)},
\end{align}
with $\tau_0\equiv t_{p0}/\sqrt{2\log{2}}$ (the value of the pulse duration $t_{p0}$
of the corresponding transform-limited pulse used in
simulations is $9.34$~fs) and $\Phi''(\omega_0)$ is
the chirp rate in the frequency domain.
Throughout the simulations the incident field is taken in the form of (\ref{chirp})
and is normalized to preserve the total energy of a laser pulse at different chirp rates
according to
\begin{equation}
\label{norm}
\int_{-\infty}^{+\infty}dt E_{inc}^2(t)=\mbox{const}.
\end{equation}
\begin{figure}[t]
\centering\includegraphics[width=\linewidth]{fig2}
\caption{\label{fig2}
(Color online) Results of FDTD simulations for chirped pulses exciting the bowtie antenna
schematically depicted in the inset of panel (a). Panel (a) shows the enhancement of the local
intensity $\left| \vec E\right|^2$ detected in the gap of the bowtie antenna as a function of the
incident frequency. Panel (b) presents local field component of electric field along the axis of
symmetry of the structure, $E_x$, as a function of time for two chirp rates at the plasmon resonance,
$\omega_0=2.057$ eV: black solid line - $\Phi''(\nu_0)=-3000$ fs$^2$, red dashed line
- $\Phi''(\nu_0)=3000$ fs$^2$. Panel (c) shows the amplitude of the local field in the frequency
domain (note that it is independent from the phase rate). Panel (d) represents the phase of the local
field in the frequency domain at two chirped rates: black solid line
- $\Phi''(\nu_0)=-3000$ fs$^2$, red dashed line - $\Phi''(\nu_0)=3000$ fs${}^2$.}
\end{figure}
The geometry considered is depicted in the inset of Fig. {\ref{fig2}}a showing the top view of the
bowtie antenna. To investigate the influence of chirped incident pulses on plasmon dynamics we
choose incident field in the form (\ref{chirp}) and vary $\Phi''(\nu_0)=4\pi^2\Phi''(\omega_0)$.
Below we shall write $\Phi''$, having in mind $\Phi''(\nu_0)$.
We further presume that
the incident pulse is $x$-polarized and propagates along the $z$ axis with the incident frequency
at the plasmon resonance (see the inset of Fig. {\ref{fig2}}a). Material dispersion of silver is taken in
the Drude form with other numerical parameters as in Ref.~\onlinecite{Sukharev2010}. For a given
set of material and geometric parameters the local electric field enhancement exhibits well
pronounced plasmon resonance as seen in Fig. {\ref{fig2}}a reaching the value of 2800 near
$2$~eV.
Our goal is to take plasmonic effects (local field enhancement and phase accumulation) directly
into account and investigate how such crafted local fields affect transport properties of molecular
junctions placed in the gap of bowtie antennas. However it is informative first to examine general
features of chirped pulses interacting with plasmonic materials. It has been noted in several
papers \cite{StockmanBergman2004,Gray2005,Brixner2007,Gray2010} that local field
enhancement depends sensitively on the sign of chirped excitation pulses. Moreover careful
examination of spatiotemporal dependence of local fields on chirp rates\cite{Gray2010} revealed a
complex dynamics of plasmon wavepackets that are noticeably influenced by chirped laser pulses
- one may find different local points for a given plasmonic system, where positive chirps lead to
higher local fields and the other way around.
Generally speaking, plasmonic materials can be considered as pulse shapers \cite{Weiner2009}
due to high material dispersion near plasmon resonance, which induces a phase in the frequency
domain resulting in shaping of the total electromagnetic field in time domain. This is illustrated in
Fig. {\ref{fig2}}b-d, where one can clearly see that the positive chirp leads to the compression of the
local field (Fig. {\ref{fig2}}b) and hence stronger field enhancement. While the field amplitude in
the frequency domain is not affected by the chirp sign (Fig. {\ref{fig2}}c),
obviously the phase of the field
is significantly different for positive and negative chirp as shown in Fig. {\ref{fig2}}d. We note that
one can not recover data obtained for negative chirp, for instance, by simply flipping the sign of the
phase for the positive chirp. Additional phase induced by the plasmonic system, which depends on
the sign of the chirp rate, makes this problem time irreversible.\cite{StockmanBergman2004}
\section{Current through the junction}\label{cur}
Time-dependent current through the junction under external driving
is\cite{Jauho1994,Sukharev2010}
\begin{align}
\label{IKt}
I_K(t)=& -\frac{e}{\hbar}\left(
\mbox{Tr}\left[\mathbf{\Gamma}^K\,\mathbf{\rho}(t)\right]
\right. \\ +& \left.
\frac{1}{\pi}\mbox{Im}\int_{-\infty}^{+\infty}dE\,f_K(E)\,
\mbox{Tr}\left[\mathbf{\Gamma}^K\,\mathbf{G}^r(t,E)\right]\right)
\nonumber
\end{align}
Here the trace is taken over molecular subspace,
$\mathbf{\Gamma}^K$ is matrix of electronic decoherence due to coupling to
contact K
\begin{equation}
\label{Gamma}
\Gamma^K_{mm'}\equiv 2\pi\sum_{k\in K}V_{mk}\,V_{km'}\,\delta(E-\varepsilon_k)
\end{equation}
which is energy independent in the wide-band approximation,
$f_K(E)$ is Fermi-Dirac thermal distribution in the contacts,
$\mathbf{\rho}(t)\equiv -i\mathbf{G}^{<}(t,t)$
is non-equilibrium reduced density matrix of molecular subsystem,
$\mathbf{G}^{r,<}(t,t')$ are matrices in molecular subspace of retarded and lesser projections
of single-electron Green function
\begin{equation}
\label{G}
G_{mm'}(\tau,\tau') \equiv -i\langle T_c\,\hat d_m(\tau)\,\hat d_{m'}^\dagger(\tau')\rangle
\end{equation}
($T_c$ is contour ordering operator), and $\mathbf{G}^r(t,E)$ is the right side Fourier transform
of the retarded Green function
\begin{equation}
\label{GrtE}
\mathbf{G}^r(t,E) = \int_{-\infty}^{+\infty}dt'\, e^{iE(t-t')}\, \mathbf{G}^r(t,t')
\end{equation}
We are interested mostly in effectiveness of the device as a charge pump, i.e.
we will calculate {\em excess charge} transferred through the system during the laser pulse
\begin{equation}
\label{QKt}
Q_K(t) = \int_{-\infty}^t dt'\, \left(I_K(t')-I_K^{dc}\right)
\end{equation}
where $I_K(t)$ is defined in Eq.(\ref{IKt}) and $I_K^{dc}$ is current at bias induced
steady-state condition, i.e. in the absence of radiation -- $E(t)=0$.
\section{Equations of motion}\label{eom}
Markov approximation employed in Ref.~\onlinecite{FainbergNitzanPRB07}
comes from consideration of time-local quantities only. This approach is sufficient
when one can neglect broadening of molecular states induced by hybridization
with states in the contacts. In realistic molecular junctions such hybridization is
non-negligible, since molecules are usually chemisorbed on at least one of the contacts.
Here (in addition to local field formation) we are going to explore how
non-Markovian effects influence characteristics of laser pulse induced charge pumping.
To keep non-Markov effects a time-nonlocal quantity --
single-particle Green function, Eq.(\ref{G}) --
is at the focus of our consideration. We employ Keldysh contour based EOM approach,
similar to the one employed in our earlier publication\cite{GalperinNitzanRatnerPRB07}
(see Appendix~\ref{appA} for derivation)
\begin{align}
\label{EOMG1}
&i\frac{\partial}{\partial\tau}G_{mm'}(\tau,\tau') = \delta_{m,m'}\delta(\tau,\tau')
+\varepsilon_m G_{mm'}(\tau,\tau')
\\ &
-\mu E(t) G_{\bar m m'}(\tau,\tau')+\sum_{m_1}\int_c d\tau_1\Sigma_{mm_1}(\tau,\tau_1)
G_{m_1m'}(\tau_1,\tau')
\nonumber \\ &
-i\sum_{k_1\neq k_2}\sum_{m_1}\left|V^{en}_{k_1k_2}\right|^2\int_c d\tau_1
g_{k_2}(\tau,\tau_1)g_{k_1}(\tau_1,\tau)
\nonumber \\ &\qquad\qquad\qquad\qquad\qquad\qquad\times
\mathcal{G}_{\bar m\bar m_1,m'm_1}(\tau,\tau_1;\tau',\tau_1+)
\nonumber
\end{align}
Here $\bar m$ is molecular level other than $m$ (e.g. for $m=1$ $\bar m=2$),
$g_k(\tau,\tau')$ is a single-particle Green function of free electron in the contacts
\begin{equation}
g_k(\tau,\tau') \equiv -i\left<T_c\,\hat c_k(\tau)\,\hat c_k^\dagger(\tau')\right>
\end{equation}
$\Sigma_{mm'}(\tau,\tau')\equiv\sum_{K=L,R}\Sigma^K_{mm'}(\tau,\tau')$ is
the self-energy due to coupling to contacts with
\begin{equation}
\label{SK}
\Sigma^K_{mm'}(\tau,\tau') \equiv \sum_{k\in K}V_{mk} g_k(\tau,\tau') V_{km'}
\end{equation}
and $\mathcal{G}$ is molecular subspace two-particle Green function
\begin{align}
\label{G2}
&\mathcal{G}_{m_1m_2,m_3m_4}(\tau_1,\tau_2;\tau_3,\tau_4)\equiv
\\ &
-\left<T_c\,\hat d_{m_1}(\tau_1)\,\hat d_{m_2}(\tau_2)\,\hat d^\dagger_{m_4}(\tau_4)\,
\hat d_{m_3}^\dagger(\tau_3)\right>
\nonumber
\end{align}
Note, in derivation of (\ref{EOMG1}) we treated the energy transfer term, Eq.(\ref{V}),
at the second order of the perturbation theory.
Presence of many-body interaction does not allow to close hierarchy of equations exactly.
To make the problem tractable we employ Markov approximation
in treating energy transfer, last term on the right in (\ref{EOMG1}), and
in writing EOM for two-particle GF (see below). These approximations are similar to those
introduced previously in Refs.~\onlinecite{GalperinNitzanJCP06,FainbergNitzanPRB07}.
Molecule-contact coupling in (\ref{EOMG1}) is treated exactly,
thus introducing non-Markov effects into description. This leads to system of equations
(see Appendix~\ref{appA} for derivation)
\begin{widetext}
\begin{align}
\label{EOMGrE}
&
i\frac{\partial}{\partial t}G^r_{mm'}(t,E) = \delta_{m,m'}+\left(\varepsilon_m-E\right) G^r_{mm'}(t,E)
-\mu E(t) G^r_{\bar m m'}(t,E)-\frac{i}{2}\sum_{m_1=1,2} \Gamma_{mm_1}G^r_{m_1m'}(t,E)
\\
\label{EOMnm}
&
\frac{d}{dt}n_m(t) = 2(-1)^m \mu E(t)\mbox{Im}\left[p(t)\right]-\Gamma_{mm}n_m(t)
-\Gamma_{m\bar m}\mbox{Re}\left[p(t)\right]
+2\mbox{Re}\sum_{m_1}\int_{-\infty}^{+\infty}\frac{dE}{2\pi}G^r_{mm_1}(t,E)\Sigma^{<}_{m_1m}(E)
\\ & \qquad\qquad
-(-1)^m\bigg(B(\varepsilon_{21})N_M(t)-B(\varepsilon_{12})\left[n_1(t)-n_2(t)+N_M(t)\right]\bigg)
\nonumber \\
\label{EOMp}
&
\frac{d}{dt}p(t) = -i\mu E(t)\bigg(n_2(t)-n_1(t)\bigg)-i\bigg(\varepsilon_2+\varepsilon_1\bigg)p(t)
-\frac{\Gamma_{21}}{2}\bigg(n_1(t)+n_2(t)\bigg)-\frac{\Gamma_{11}+\Gamma_{22}}{2}p(t)
\\ &
+\sum_{m_1=1,2}\int_{-\infty}^{+\infty}\frac{dE}{2\pi}\bigg(
G^r_{2m_1}(t,E)\Sigma_{m_11}^{<}(E)-\Sigma_{2m_1}^{<}(E)\overset{*}{G}{}^r_{1m_1}(t,E)
\bigg)
-i\sum_{m_1}B(\varepsilon_{m_1\bar m_1})\mbox{Im}\left[p(t)\right]
\nonumber \\
\label{EOMNM}
&
\frac{d}{dt}N_M(t) = 2\mu E(t)\mbox{Im}\left[p(t)\right]
-i\Sigma_{22}^{<}(\varepsilon_2)n_1(t)
+i\Sigma_{11}^{>}(\varepsilon_1)n_2(t)
-2i\bigg(\Sigma_{12}^{>}(\varepsilon_1)+\Sigma_{12}^{<}(\varepsilon_2)\bigg)\mbox{Re}
\left[p(t)\right]
\\ & \qquad\qquad
-\bigg(\Gamma_{11}+\Gamma_{22}+B(\varepsilon_{21})\bigg)N_M(t)
+B(\varepsilon_{12})\bigg(n_1(t)-n_2(t)+N_M(t)\bigg)
\nonumber
\end{align}
\end{widetext}
Here $\varepsilon_{m\bar m}\equiv\varepsilon_{m}-\varepsilon_{\bar m}$,
$n_m(t)\equiv\rho_{mm}(t)$ ($m=1,2$) are populations of molecular levels,
$p(t)\equiv\rho_{21}(t)$ is molecular coherence,
$N_M(t)\equiv\left\langle\hat D^\dagger(t)\hat D(t)\right\rangle\equiv\mathcal{G}_{12,12}(t+,t;t,t+)$
is the molecular excitation correlation function,
$\Gamma_{mm'}\equiv\sum_{K=L,R}\Gamma^K_{mm'}$
is the matrix of electronic decoherence due to electron transfer between the molecule and
contacts,
with $\Gamma^K_{mm'}$ defined in Eq.(\ref{Gamma}),
$\Sigma_{mm'}^{>,<}(E)=\sum_{K=L,R}\Sigma^{K>,<}_{mm'}(E)$ greater (lesser)
projections of self-energy due to coupling to contacts with
\begin{align}
\Sigma^{K<}_{mm'}(E) \equiv & i\Gamma^K_{mm'}f_K(E) \\
\Sigma^{K>}_{mm'}(E) \equiv & -i\Gamma^K_{mm'}\left[1-f_K(E)\right]
\end{align}
and $B(E)$ is the dissipation rate due to energy transfer
\begin{align}
\label{B}
B(E) \equiv& 2\pi\sum_{K=L,R}\sum_{k_1\neq k_2\in K} \left|V_{k_1k_2}^{en}\right|^2
\delta(\varepsilon_{k_1}-\varepsilon_{k_2}+E)
\nonumber\\ &\qquad\qquad\qquad\qquad\times
f_K(\varepsilon_{k_1})[1-f_K(\varepsilon_{k_2})]
\end{align}
Note that in (\ref{EOMGrE}) we omitted term coming from energy transfer,
since contribution to the total retarded self-energy $\Sigma^r$
from molecule-contacts electron transfer $\sim\Gamma$ is much bigger than
corresponding contribution from energy transfer
$\sim B(\varepsilon_{21})n_2$ ($\sim B(\varepsilon_{21})(1-n_1)$)
for $m=1$ ($m=2$) in a reasonable parameter
range.\cite{GalperinNitzanJCP06,GalperinNitzanRatnerPRL06}
EOMs~(\ref{EOMGrE})-(\ref{EOMNM})
form a closed set of time-dependent equations to be solved simultaneously
on energy grid starting from a steady-state initial condition corresponding
to biased junction before the laser is switched on.
Density matrix $\mathbf{\rho}(t)$ and retarded GF $\mathbf{G}^r(t,E)$
obtained as the solution are used in (\ref{IKt}) and (\ref{QKt})
to calculate time-dependent current and excess charge pumped through the junction, respectively.
In the limit of weak molecule-contact coupling $\Gamma\to 0$,
neglecting local field and non-Markov effects, disregarding
off-diagonal terms in spectral function, and assuming rotating-wave approximation
Eqs.~(\ref{EOMGrE})-(\ref{EOMNM}) are reduced to results of
Ref.~\onlinecite{FainbergNitzanPRB07} (see Appendix~\ref{appB} for details).
\section{Results and discussion}\label{res}
Here we present results of numerical simulations for the model (\ref{H})-(\ref{V}) with local field
formation and non-Markov effects taken into account as described above.
Time dependent local electromagnetic field is calculated solving Maxwell's equations on a grid (see section~\ref{optics})
for metallic contacts of a bowtie geometry.
Molecule is placed in a `hot spot' situated between the contacts, and local field plays a role
of external driving force in electronic calculations (as described in Section~\ref{eom}).
Unless stated otherwise parameters of the electronic simulations:
temperature is $300$K, molecular electronic level positions
$\varepsilon_1=-1$eV and $\varepsilon_2=1$eV, elements of electronic decoherence matrix
are $\Gamma^L_{11}=\Gamma^R_{22}=0.1$eV, $\Gamma^L_{22}=\Gamma^R_{11}=0.01$eV,
$\Gamma^{L,R}_{12}=\Gamma^{L,R}_{21}=0$, coupling to external field
$\mu\mathcal{E}_0=0.008$eV (after normalization (\ref{norm}) for $\Phi''=2000$~fs${}^2$;
also below coupling to external field below is given renormalized according to (\ref{norm})
for particular $\Phi''$).
Fermi energy is taken as origin $E_F=0$, bias is applied symmetrically
$\mu_{L,R}=E_F\pm |e|V_{sd}/2$. All calculations except Fig.~\ref{fig6} below are
done at equilibrium, $V_{sd}=0$.
Only processes of energy relaxation on the molecule
are taken into account with $B(\varepsilon_{12})=0$ and $B(\varepsilon_{21})=0.1$eV.
Time grid is taken from the external driving field simulations. Energy grid spans region
from $-20$ to $20$eV with step $0.001$eV. Other parameters are introduced separately
for each calculation.
\begin{figure}[t]
\centering\includegraphics[width=\linewidth]{fig3}
\caption{\label{fig3}
(Color online) Time dependence of charge pumped through the junction Q(t), Eq.(\ref{QKt})
for several chirp rates.
Shown are results for (a) $\Phi''=2000$ and $-2000$ fs${}^2$
without energy transfer $B(\varepsilon_{21})=0$ (solid line, red and dashed line, blue, respectively), and taking into account energy transfer term $B(\varepsilon_{21})=0.1$~eV
(dash-dotted line, red and dotted line, blue, respectively), and
(b) $\Phi''=10000$ and $-10000$ fs${}^2$ for $B(\varepsilon_{21})=0.1$~eV and
$\mu\mathcal{E}_0=0.03$ (solid line, red and dashed line, blue, respectively)
and $\mu\mathcal{E}_0=0.003$~eV (dash-dotted line, red and dotted line, blue, respectively).
}
\end{figure}
Figure~\ref{fig3} demonstrates pumped charge build-up during the laser pulse excitation.
One sees that the local field formation leads to asymmetry in pumped charge for opposite
chirp rates. Negatively chirped incoming field creates longer local pulse (see Fig.~\ref{fig2}b),
which results in increase in total charge pumped through the junction.
Role of electron-hole excitations in the contacts on charge buildup is shown in Fig.~\ref{fig3}a.
Since processes of escape from LUMO into the right contact and energy relaxation on
the molecule compete for the excited state population, current (and consequently pumped charge)
diminish with increase of coupling to electron-hole excitations in the contacts.
Fig.~\ref{fig3}b shows effect of intensity of incoming pulse on the transfered charge buildup.
For higher intensity the build-up demonstrates saturation in the middle of the pulse.
The reason for this behavior is the competition between timescales related to
Rabi oscillation induced by local field between molecular levels
and electronic escape rate from molecule into contacts ($\sim 1/\Gamma$).
On the one hand, both negatively and positively chirped pulses in the middle have
frequency approximately at resonance with HOMO-LUMO transition,
$\omega\approx\varepsilon_2-\varepsilon_1$, which is a prerequisite to
effective electron transfer and thus increase in pumped charge.
On the other hand, at resonance Rabi oscillations\cite{Landau_v3_1991}
at high enough intensities compete with electron escape rate, thus
effectively blocking current through the junction. Depending on parameters
this may lead either to most effective charge transfer in the middle of the pulse
(dash-dotted and dotted lines in panel b), or to suppression of charge transfer
at this point (solid and dashed line in panel b).
Note that the effect is not related to non-Markov relaxation,
i.e. this behavior is observed also in the absence
of hybridization between molecule and contact(s) states, and its relation
to Landau-Zener problem\cite{Nitzan_2006} in terms of total charge pumped across the junction
was discussed in Ref.~\onlinecite{FainbergNitzanPRB07}.
Note also, that with positively chirped pulse changing frequency from lower to higher
transfered charge buildup is more effective at the start of the pulse (at lower frequencies),
while for negatively chirped pulse more effective buildup takes place at the end of the pulse
(compare solid and dashed lines in Figs.~\ref{fig3}b and \ref{fig5}b).
Contrary to buildup suppression in the middle of the pulse, this effect is due
to molecule-contact hybridization. The latter leads to broadening of molecular levels, and
effectiveness of HOMO-LUMO charge transfer depends (among other conditions)
on integral of occupied states at HOMO and empty states at LUMO
separated by frequency of incident light $\int dE G^{<}_{11}(E) G^{>}_{22}(E+\omega)$.
Clearly, at frequencies below resonance the latter is greater than at frequencies above it.
\begin{figure}[htbp]
\centering\includegraphics[width=\linewidth]{fig4}
\caption{\label{fig4}
(Color online) Charge pumped through the junction during pulse vs. chirp rate:
(a) total charge, i.e. integral of Q(t), Eq.(\ref{QKt}), over local field
pulse duration, and (b) asymmetry in charge transfer between positively and negatively chirped
incoming laser pulses, $\Delta Q\equiv Q(|\Phi''|)-Q(-|\Phi''|)$.
Shown are results with ($B(\varepsilon_{21})=0.1$eV, dashed line with triangles, blue)
and without ($B(\varepsilon_{21})=0$, solid line with circles, red) energy transfer.
}
\end{figure}
Local field asymmetry relative to the sign of the chirp rate in the frequency domain
leads to asymmetry in charge pumping contrary to symmetric situation presented
in Ref.~\onlinecite{FainbergNitzanPRB07}, as is demonstrated in Fig.~\ref{fig4}a.
One can see that the pumped charge is almost symmetric at high rates with
asymmetry confined to the low rate region. Difference between
charge pumped through the junction at positive and negative chirp rates is shown in
Fig.~\ref{fig4}b. As discussed above duration of local field due to positively chirped
incoming pulse is shorter than the one due to negatively chirped analog.
This compression is the cause of less charge pumped through the system in the former case,
which results in decrease in $\Delta Q\equiv Q(\Phi'')-Q(-\Phi'')$ in the region of $\Phi''(\nu_0)$
from $0$ to $3000$ fs${}^2$.
Indeed, at the very low rates frequency of the pulse does not change much, so the asymmetry
is solely due to difference in pulse length. At a higher rates an additional factor appears:
the most effective contribution to charge transfer takes place at a particular region of frequencies
(at and just below resonance, as is discussed above).
This region is passed quicker in the positively chirped local pulse, and
in the region up to $3000$~fs${}^2$ this results in increase of asymmetry,
since negatively chirped pulse spends more time in its effective frequencies zone.
Further increase of chirp rate leads to decrease and almost disappearance of the asymmetry.
The reason is decrease of ratio of the pulses difference to overall local pulse duration.
Coupling to electron-hole excitations not only diminishes pumped charge
(compare solid and dashed lines in panel a), but also decreases asymmetry (panel b).
The latter results from the fact that rate for molecular energy relaxation (LUMO ${}\to{}$ HOMO
transition due to coupling to excitations in the contacts) is proportional to population in
the LUMO (see discussion in Ref.~\onlinecite{GalperinNitzanJCP06}). So for higher currents also
energy relaxation will be more efficient, thus effectively compensating for the difference.
\begin{figure}[t]
\centering\includegraphics[width=\linewidth]{fig5}
\caption{\label{fig5}
(Color online) Dependence of charge pumping on molecule-contact states hybridization
at $\mu\mathcal{E}_0=0.03$eV.
Shown are (a) Charge pumped through the junction during pulse vs. electronic
escape rate for chirp rate $\Phi''=10000$ fs${}^2$ (solid line, red) and $-10000$ fs${}^2$
(dashed line, blue), and
(b) Normalized transferred charge build-up (charge normalized to total charge transferred
during the pulse) vs. time for chirp rate $\Phi''=10000$ fs${}^2$ and
$\Gamma^L_{11}=\Gamma^R_{22}=0.1$eV (solid line, red)
and $1$eV (dashed line, blue).
Red line in panel (b) is the same as solid red line in Fig.~\ref{fig3}b.}
\end{figure}
Importance of non-Markov behavior for charge pump is demonstrated in Figure~\ref{fig5}.
Fig.~\ref{fig5}a shows pumped charge as function of level width (for two opposite
choices of chirp rate).
Increase in total charge pumped through the junction with increase in hybridization
saturates at high strength of coupling between molecule and contacts.
Such behavior is expected: at low hybridization there is only
one frequency corresponding to resonance, where pumping is most
effective, so only an 'instant' of chirped pulse contributes to charge transfer.
As molecule-contact coupling grows the condition of resonance transition becomes less and
less strict. Eventually any frequency within the chirped pulse has roughly
same effectiveness -- this is the reason for saturation. Also, stronger coupling
means more effective molecule-contact electron transfer, which competes more effectively with
intra-molecular Rabi oscillation at resonance. This competition is demonstrated in Fig.~\ref{fig5}b,
where middle-of-the-pulse saturation (see discussion of Fig.~\ref{fig3}) disappears
for stronger molecule-contact coupling.
\begin{figure}[t]
\centering\includegraphics[width=\linewidth]{fig6}
\caption{\label{fig6}
(Color online) Dependence of charge pumping on bias.
(a) Total excess charge pumped through the junction during pulse vs. bias
for chirp rate $\Phi''=10000$ fs${}^2$ (solid line, red) and $-10000$ fs${}^2$
(dashed line, blue).
(b) Excess charge build-up (normalized to total excess charge transferred
during the pulse) vs. time at chirp rate $\Phi''=10000$ fs${}^2$
for $V_{sd}=-1$V (dashed line, blue), $0$V (solid line, red), and
$1$ V (dotted line, black).
Here $\Gamma^L_{11}=\Gamma^R_{22}=0.5$~eV.
Red line in panel (b) is the same as solid red line in Fig.~\ref{fig3}b.
}
\end{figure}
Finally, in Figure~\ref{fig6} we discuss influence of bias on charge pumping.
Here we define optically pumped charge (excess charge) as a difference between
charge pumped through the junction with and without laser field.
Fig.~\ref{fig6}a demonstrates total excess pumped charge during laser pulse
for opposite choices of chirp rate as function of bias.
Application of bias has 2 effects on the pumping process:
1. it depletes (populates) the HOMO (the LUMO) and 2. it may block or release
channels for electron transfer from LUMO to contact $R$.
This leads to a situations when most effective optical pumping does not
correspond to zero bias, rather we see shallow peak at $V\sim 1$V.
Explanation is related to the fact that broadened molecular levels are
essentially a set of scattering channels with different transmission
probabilities: high conducting channels are in the center of Lorentzian,
while channels in the sides of distribution are poor conductors.
Optical process takes electron from an occupied ground state and puts it into
one of empty excited state. Effectiveness of the charge pump is
defined by increase or decrease of current through the junction
under optical pulse (see discussion in Ref.~\onlinecite{Sukharev2010}).
In particular, negative bias decreases effectiveness of the pump mostly due to
blocking part of LUMO-$R$ escape routes. Positive bias opens additional
escape routes at the tail of LUMO Lorentzian, facilitating increase in pump efficiency.
However, additional effect of depleting HOMO and populating LUMO
partially blocks optically induced HOMO-LUMO electron transfer, thus
reducing overall effectiveness of optical pumping. Competition between the two proceses
reveals itself as a shallow peak at $\sim 1$V.
Time resolved charge buildup is presented in Fig~\ref{fig6}b. Middle-of-the-pulse
saturation observed previously at equilibrium (solid line, same as in Fig.~\ref{fig3}b)
is enhanced at negative (dashed line) and disappears at positive bias (dotted line).
The reason is similar to competition between Rabi frequency and escape rate discussed
above. Indeed, with negative bias partially blocking fast escape route for the electron from
excited state into right contact, Rabi oscillation plays an important role at quasi-resonant situation
in the middle of the pulse. Positive opening additional routes makes Rabi oscillation less effective.
\section{Conclusion}\label{conclude}
We consider a two-level (HOMO-LUMO) model for optically-driven molecular charge pump.
Such pump may be realized as a junction formed by a molecule with strong charge-transfer
transition between its ground and excited states. The junction is driven by both applied bias
and laser pulse. The latter is treated as an external classical driving force.
Our consideration is the generalization of previous study\cite{FainbergNitzanPRB07}
which takes into account effects of local field (`hot spot' formation) and
hybridization between the states of molecule and contact(s) (non-Markov effects).
We formulate approximate closed set of EOMs for single- and two-particle GFs.
Electron transfer in the former is treated exactly. To close set of equations the latter
are considered within Markov approximation. Our EOMs are reduced to
set of equation derived in Ref.~\onlinecite{FainbergNitzanPRB07}
under several simplifying assumptions: weak molecule-contact coupling (neglect of hybridization),
neglect of non-diagonal terms in molecular spectral function, and within
rotating wave approximation.
Incoming laser pulse is assumed to be linearly chirped. Local field is calculated
within FDTD technique on a grid with bowtie antenna geometry used to represent
junction metallic contacts.
We find that contrary to symmetric behavior of the pump relative to sign
of the chirp rate, duration of the corresponding local field pulse depends on the sign of incoming chirp, which results in asymmetric operation of the pump. The asymmetry depends on
the incoming pulse chirp rate in a non-monotonic manner. Junction response to optical driving
is symmetric to both low and high chirp rates, going through a maximum between the two extremes.
We find that this behavior is caused by the correspondence between pulse duration of
the local field and detuning of its frequency at the end of the pulse from energy difference
between molecular states ($\varepsilon_2-\varepsilon_1$).
We note that at quasi-resonance
charge pump becomes ineffective, due to the competition between intra-molecular
Rabi oscillation induced by the pulse with electron transfer from molecule to contact.
Increase of the molecule-contact coupling strength increases electron escape rate,
thus reducing ineffectiveness of the pump due to Rabi oscillations.
Also we study the effect
of bias on optically-facilitated charge transfer through the junction. We find that in the non-Markov
situation (i.e. when hybridization between molecule and contacts is non-negligible)
most effective charge pump regime is at finite positive bias, rather than at equilibrium
as one may expect from Markov consideration of Ref.~\onlinecite{FainbergNitzanPRB07}.
The effect comes from optically-assisted charge redistribution between low and high
conducting scattering channels in broadened molecular states, as was discussed in
our previous publication.\cite{Sukharev2010} Within the model negative bias reduces
(positive increases) excess charge pumping due to blocking (facilitating) outgoing
scattering channels in the excited molecular state and thus increasing (decreasing)
the role of intra-molecular Rabi oscillations.
Finally, direct electron-hole excitation in contacts, heating, and inelastic effects
are examples of effects beyond current consideration which may also have
a significant impact on the properties of a molecular charge pump.
\begin{acknowledgments}
We acknowledge support by the National Science Foundation (MG, CHE-1057930),
the US-Israel Binational Science Foundation (BF and MG, \#2008282),
and by the UCSD (MG, startup funds).
\end{acknowledgments}
|
\section{Introduction}
\label{sec:intro}
The mixing of Frenkel excitons (FEs) and charge-transfer excitons (CTEs) has been studied both experimentally and theoretically in many one-compo\-nent molecular crystals, \emph{e.g.}, polyacenes \cite{sebastian81,sebastian83,siebrand83,petelenz96}, perylene derivatives \cite{henessy99,hoffmann00,hoffmann02}, fullerenes \cite{jeglinski92,pac98}, and other. The excitonic and vibronic spectra of quasi-one-dimensional crystals like 3,4,9,10-perylenetetracarboxylic dianhydride (PTCDA) and $N,N^{\prime}$-dimethylperylene-3,4,9,10-perylenetetracarbo\-ximide (MePTCDI) have been treated in \cite{henessy99,hoffmann00,hoffmann02} and \cite{schmidt02,lalov06a,lalov07a,lalov06b,lalov06c}. The short molecular distance in quasi-one-dimensional stacks causes a strong FE--CTEs mixing and mixing of their vibronic satellites as well, especially in the absorption spectra in the spectral region of $2$--$3$ eV ($15\,000$--$23\,000$ cm$^{-1}$).
In the present paper, we calculate the linear absorption spectra of one-component MePTCDI and PTCDA crystals applying the vibronic approach developed in our previous papers \cite{lalov06b,lalov07b}. It allows complex calculations of the absorption in pure excitonic, one-phonon vibronic, two-phonon vibronic, etc.\ spectra. In paper \cite{lalov08} the vibronic approach is the tool of studying the FE--CTEs mixing in a two-component stack of alternatively arranged donor--acceptor (DA)-molecules. While the FE--CTEs mixing in \cite{lalov08} is a probable but hypothetical model, in the present study we turn to the real absorption spectra and use the parameters of the FE--CTEs mixing from Refs.~\cite{henessy99,hoffmann00,schmidt02}.
Two general differences exhibit the models of the FE--CTEs mixing in one- and two-component molecular stacks \cite{lalov08}, notably: (i) The intermolecular transfer both of electrons and holes must be a feature of the model for the case of one-component stack whereas only one transfer mechanism is sufficient in a DA-molecular stack. In the last case the Frenkel exciton represents a collectivized electronic excitation of a donor or an acceptor and the strongest transfer on the closest neighbor would be either of a hole or correspondingly of an electron. Obviously in an one-component stack both types of transfer on the neighbors could be of the comparable probability. The two-step processes of successive transfer of the electron and the hole on the same molecule ensure the transfer of a FE even in the case of its relatively weak direct intermolecular transfer \cite{schmidt02}. In this way, our present study extends the calculations in \cite{lalov06b} where only one transfer mechanism has been considered. (ii) In the most widely studied DA-crystal, anthracene-PMDA the excitonic absorption lines are narrow \cite{haarer75,brillante80,weiser04} and many details in the vibronic spectra can be seen. In PTCDA and MePTCDI crystals the absorption lines are two order of magnitude wider. Thus we pay attention to the general structure of the excitonic and one-phonon vibronic spectra supposing a width of the excitonic lines of $300$--$500$ cm$^{-1}$ (not $2$--$10$ cm$^{-1}$ which is the absorption width of the anthracene-PMDA).
In the next section of the paper we involve the initial Hamiltonian in the case of a FE--CTEs mixing. The Hamiltonian contains one mode of intramolecular vibration linearly coupled to the FE and CTEs. In Section~3 the linear optical susceptibility has been calculated in the excitonic and one-phonon vibronic spectra. In Section~4 the linear absorption has been modelled using the excitonic and vibrational parameters of the PTCDA and MePTCDI crystals fitted in Refs.~\cite{hoffmann00,hoffmann02,schmidt02}. Section~5 summarizes the main findings derived in this paper. In the Appendix we treat some problems of the FE--CTEs mixing in a molecular stack of the $C_i$ point group of symmetry.
\section{Hamiltonian for the Case of a FE--CTEs--Phonon Coupling}
\label{sec:Hamiltonian}
We consider the excitonic and vibronic excitations in a linear molecular stack of $N$ identical molecules which are regularly arranged at a distance $d$ each other. The point group of symmetry of the stack is $C_i$ as is for the molecular stacks of PTCDA and MePTCDI \cite{hoffmann00}. The origin of the Frenkel exciton is a non-degenerate molecular electronic excitation with excitation energy $E_{\rm F}$ and $L$ being the transfer integral between neighboring molecules. We denote by $B_n$ ($B_n^{+}$) the annihilation (creation) operator of the electronic excitation on molecule $n$ and get the following FE-part of the Hamiltonian:
\begin{equation}
\label{eq:fehamilt}
H_{\rm FE} = \sum_n E_{\rm F} B_n^{+}B_n + \sum_{n,n^{\prime}} L \left( \delta_{n^{\prime},n+1} + \delta_{n^{\prime},n-1} \right) B_{n^{\prime}}^{+}B_n.
\end{equation}
As usually, we consider two CTEs of equal excitation energy $E_{\rm c}$ and $C_{n,1}$ is the annihilation operator CTE, $\sigma =1$, with hole located on the site $n$ and electron on the site $n + 1$, whereas the electron of the second CTE, $\sigma =2$, ($C_{n,2}$) is located at molecule $n - 1$. We neglect the transfer of CTEs as a whole and the mutual coupling of the two CTEs since those processes can be realized through the transfer of the electron or hole at distance $2d$ which is less probable than the FE--CTEs mixing caused by the electron (hole) transfer at the neighbor molecule (see Refs.~\cite{lalov08,brillante80}). We have the following CTEs-part of the Hamiltonian:
\begin{equation}
\label{eq:cteshamilt}
\hat{H}_{\rm CTE} = \sum_{n,\sigma = 1,2} E_{\rm c}C_{n \sigma}^{+}C_{n \sigma}
\end{equation}
and suppose the following operator for the FE--CTEs mixing:
\begin{eqnarray}
\label{eq:fcoperator}
\hat{H}_{\rm FCT} = \sum_n \left[ \varepsilon_{\rm e1}B_n^{+}C_{n,1} + \varepsilon_{\rm e2}B_n^{+}C_{n,2} \right. \nonumber \\
\left.
{}+ \varepsilon_{\rm h1}B_n^{+}C_{n-1,1} + \varepsilon_{\rm h2}B_n^{+}C_{n+1,2} + \mbox{h.c.} \right],
\end{eqnarray}
where $\varepsilon_{\rm e1}$ and $\varepsilon_{\rm e2}$ are the transfer integrals of the electron from molecule $n$ to molecules $n+1$ and $n-1$ correspondingly, and $\varepsilon_{\rm h1}$, $\varepsilon_{\rm h2}$ denote the transfer integrals of the hole from molecule $n$ to molecules $n+1$ and $n-1$. Certainly the model with four transfer integrals is more complicated than the model in Refs.~\cite{hoffmann00,hoffmann02} with two mixing parameters only. But our model is more realistic because we take into account the inclination of the flat molecules of PTCDA and MePTCDI relatively to the stack axis (see the Appendix).
One intramolecular mode is only supposed to be coupled with the FE and CTEs and the phonon part of the Hamiltonian is
\begin{equation}
\label{eq:phonhamilt}
\hat{H}_{\text{ph}} = \sum_n \hbar \omega_0 a_n^{+} a_n,
\end{equation}
where $\omega_0$ is the vibrational frequency and $a_n$ is the annihilation operator of one vibrational quantum on molecule $n$. The linear exciton--phonon coupling only is manifested in the treated crystals \cite{hoffmann02} and we get the following exciton--phonon part \cite{henessy99,hoffmann00,hoffmann02,lalov05}
\begin{eqnarray}
\label{eq:exc-phon}
\hat{H}_{\text{ex--phon}} = \sum_{n,\sigma=1,2}\hbar \omega_0\left[ \xi_{\rm F}B_n^{+} B_n \left( a_n^{+} + a_n \right)\right. \nonumber \\
\left. {}+ \xi C_{n\sigma}^{+} C_{n\sigma} \left( a_n^{+} + a_n + a_{n+\sigma_1}^{+} + a_{n+\sigma_1} \right) \right],
\end{eqnarray}
where $\xi_{\rm F}$ and $\xi$ are dimensionless parameters of the linear FE--phonon and CTEs--phonon coupling, correspondingly; $\sigma_1 = +1$ if $\sigma = 1$, and $\sigma_1 = -1$ if $\sigma = 2$. For the sake of simplicity, we suppose the same linear exciton--phonon coupling in the positive and negative ions which create CTEs.
The full Hamiltonian $\hat{H}$ contains all the parts (\ref{eq:fehamilt})--(\ref{eq:exc-phon}) and can be transformed using the canonical transformation which eliminates the linear exciton--phonon coupling (\ref{eq:exc-phon}), see \cite{lalov05,davydov71},
\begin{equation}
\label{eq:h1}
\hat{H}_1 = \exp(Q) \hat{H} \exp(-Q),
\end{equation}
where
\begin{eqnarray}
\label{eq:q}
Q = \sum_{n,\sigma=1,2} \left[ \xi_{\rm F}B_n^{+} B_n \left( a_n^{+} - a_n \right) \right. \nonumber \\
\left. {}+ \xi C_{n\sigma}^{+} C_{n\sigma} \left( a_n^{+} - a_n + a_{n+\sigma_1}^{+} - a_{n+\sigma_1} \right) \right].
\end{eqnarray}
We introduce the vibronic operators
\begin{equation}
\label{eq:vn}
V_n = \exp(Q) B_n \exp(-Q),
\end{equation}
\begin{equation}
\label{eq:unsigma}
U_{n,\sigma} = \exp(Q) C_{n\sigma} \exp(-Q)
\end{equation}
and get the following transformed Hamiltonian
\begin{eqnarray}
\label{eq:h1new}
\hat{H}_1 = \sum_n \left( E_{\rm F} - \hbar \omega_0 \xi_{\rm F}^2 \right)V_n^{+}V_n + \sum_{n n^{\prime}} L \left( \delta_{n^{\prime},n+1} + \delta_{n^{\prime},n-1} \right) V_{n^{\prime}}^{+} V_n \nonumber \\
{}+ \sum_{n,\sigma} \left( E_{\rm c} - 2\hbar \omega_0 \xi^2 \right)U_{n \sigma}^{+}U_{n \sigma} + \sum_n \hbar \omega_0 a_n^{+} a_n + \sum_n \left[ \varepsilon_{\rm e1}V_n^{+}U_{n1} \right. \nonumber \\
\left.
{}+ \varepsilon_{\rm e2}V_n^{+}U_{n2} + \varepsilon_{\rm h1}V_n^{+}U_{n-1,1} + \varepsilon_{\rm h2}V_n^{+}U_{n+1,2} + \mbox{h.c.} \right]
\end{eqnarray}
In a stack with inversion center, point group $C_i$, the excitons are \emph{gerade\/} or \emph{ungerade\/} (in the center of the Brillouin zone, at $k = 0$). The ungerade excitons only are dipole-active and influence the linear optical susceptibility and the absorption spectra. In the case under consideration, the operator of the transition dipole moment has the following form \cite{lalov08,haarer75}:
\begin{equation}
\label{eq:p}
P = \sum_n \left[ \mathbf{p}_{\rm F}\left( V_n^{+} + V_n \right) + \mathbf{p}_{\rm CT}\left( U_{n2} - U_{n1} + U_{n2}^{+} - U_{n1}^{+} \right) \right].
\end{equation}
The gerade FEs, as well as the symmetrical combination of CTEs, $\left( U_{n2}^{+}\right.$ $\left.+\; U_{n1}^{+} \right)|0\rangle$, can be also mixed, but due to their vanishing transition dipole moment they will not be considered here.
Introducing the Fourier transform in the momentum space of the vibronic operators,
\begin{equation}
\label{eq:vk}
V_k = \frac{1}{\sqrt{N}}\sum_n V_n \exp(\mathrm{i}knd),
\end{equation}
\begin{equation}
\label{eq:uksigma}
U_{k,\sigma} = \frac{1}{\sqrt{N}}\sum_n U_{n,\sigma} \exp(\mathrm{i}knd),
\end{equation}
we obtain the following form of the Hamiltonian (\ref{eq:h1new})
\begin{eqnarray}
\label{eq:h1new1}
\hat{H}_1 \!\!\!\!\!&=&\!\!\!\!\! \sum_k \left( E_{\rm F} - \hbar \omega_0 \xi_{\rm F}^2 + 2L\cos k \right)V_k^{+}V_k \nonumber \\
&&\!\!\!\!\!
{}+ \sum_{k,\sigma} \left( E_{\rm c} - 2\hbar \omega_0 \xi^2 \right) U_{k\sigma}^{+}U_{k\sigma} \nonumber \\
&&\!\!\!\!\!
{}+ S + \sum_n \hbar \omega_0 a_n^{+} a_n,
\end{eqnarray}
where
\[
S = \sum_k \left\{ A V_k^{+}\left( U_{k,1} + U_{k,2} \right) + B V_k^{+}\left( U_{k,2} - U_{k,1} \right) + \mbox{h.c.} \right\}
\]
with
\[
A = \varepsilon_{\rm e}^{\prime} + \varepsilon_{\rm h}^{\prime}\cos k + \mathrm{i}\varepsilon_{\rm h}^{\prime \prime} \sin k \quad \mbox{and} \quad
B = \varepsilon_{\rm e}^{\prime \prime} + \varepsilon_{\rm h}^{\prime \prime}\cos k
+ \mathrm{i}\varepsilon_{\rm h}^{\prime} \sin k .
\]
Here
\begin{equation}
\label{eq:epsprime}
\varepsilon_{\rm e}^{\prime} = \left( \varepsilon_{\rm e1} + \varepsilon_{\rm e2} \right)/2, \qquad \varepsilon_{\rm h}^{\prime} = \left( \varepsilon_{\rm h1} + \varepsilon_{\rm h2} \right)/2,
\end{equation}
\begin{equation}
\label{eq:epssec}
\varepsilon_{\rm e}^{\prime \prime} = \left( \varepsilon_{\rm e2} - \varepsilon_{\rm e1} \right)/2, \qquad \varepsilon_{\rm h}^{\prime \prime} = \left( \varepsilon_{\rm h2} - \varepsilon_{\rm h1} \right)/2.
\end{equation}
For the case of ungerade FEs, their mixing with the symmetrical (even) combination $\left( U_{k,1} + U_{k,2} \right)$ at $k = 0$ is impossible, and thus the mixing parameters $\varepsilon_{\rm e}^{\prime}$ and $\varepsilon_{\rm h}^{\prime}$ vanish, $\varepsilon_{\rm e}^{\prime} = \varepsilon_{\rm h}^{\prime} = 0$. The final expressions for operators (\ref{eq:fcoperator}) and (\ref{eq:p}) are
\begin{eqnarray}
\label{eq:fc}
\hat{H}_{\rm FCT} = \sum_k \left[ \left( \varepsilon_{\rm e}^{\prime \prime} + \varepsilon_{\rm h}^{\prime \prime}\cos k \right) V_k^{+} \left( U_{k,2} - U_{k,1} \right) \right. \nonumber \\
\left.
{}+ \mathrm{i}\varepsilon_{\rm h}^{\prime \prime}\sin k V_k^{+} \left( U_{k,2} + U_{k,1} \right) + \mbox{h.c.} \right]
\end{eqnarray}
and
\begin{eqnarray}
\label{eq:pfinal}
\hat{P} = \sqrt{N}\left[ \mathbf{p}_{\rm F}\left( V_{k=0} + V_{k=0}^{+} \right) \right. \nonumber \\
\left.
{}+ \mathbf{p}_{\rm CT}\left( U_{k=0,2} - U_{k=0,1} + U_{k=0,2}^{+} - U_{k=0,1}^{+} \right) \right],
\end{eqnarray}
respectively.
\section{Calculation of the Linear Optical Susceptibility}
\label{sec:optical}
The linear optical susceptibility can be calculated by using the formula \cite{agranovich83}
\begin{equation}
\label{eq:chiij}
\chi_{ij} = \lim_{\epsilon \to 0}\left\{ \frac{1}{2\hbar V}\left[ \Phi_{ij}(\omega + \mathrm{i}\epsilon) + \Phi_{ij}(-\omega + \mathrm{i}\epsilon) \right] \right\}
\end{equation}
with
\begin{equation}
\label{eq:phiij}
\Phi_{ij}(t) = -\mathrm{i} \theta(t)\langle 0| \hat{P}_i(t)\hat{P}_j(0) + \hat{P}_j(t)\hat{P}_i(0)|0\rangle,
\end{equation}
where $V$ is the crystal's volume [in our case being proportional to $Nv$ ($v$ is the volume occupied by one molecule)] and $\hat{P}$ is the operator (\ref{eq:pfinal}). The Green functions (\ref{eq:phiij}) have been calculated as average over the ground state $|0\rangle$ only by taking into account the large values of $E_{\rm F}$, $E_{\rm c}$, $\hbar \omega_0 \gg k_{\rm B}T$.
We calculate the Green functions (\ref{eq:phiij}) following the vibronic approach \cite{lalov07b,lalov08}. In the next expression, the $x$ axis is supposed to be oriented along the vector $\mathbf{p}_{\rm F}$ which includes angle $\gamma$ with the vector $\mathbf{p}_{\rm CT}$, and $a = \left| p_{\rm CT}/p_{\rm F} \right|$. Then we can represent the linear optical susceptibility of one stack as
\begin{equation}
\label{eq:chixx}
\chi_{xx} = -\frac{p_{\rm F}^2}{v}\frac{1}{\alpha_1 \alpha_2 - 2\alpha_{12}^2} \left[ \alpha_2 + 4a\alpha_{12}\cos \gamma + 2a^2 \alpha_1 \cos^2 \gamma \right].
\end{equation}
The functions $\alpha_1$, $\alpha_2$, $\alpha_{12}$ have been calculated for the excitonic and one-phonon vibronic regions (see below). The PTCDA and MePTCDI crystals contain two types $A$ and $B$ of parallel molecular stacks, however, the excitonic and vibronic excitations in each stack interact very weakly with the excitations of the other stacks. In the same way as in Ref.~\cite{lalov06b}, we calculate the crystal's susceptibility in an oriented gas model. We denote by $2\varphi$ the angle between the vectors $\mathbf{p}_{\rm F}^A$ and $\mathbf{p}_{\rm F}^B$ of two different stacks and suppose that these vectors are positioned in the $(XY)$ plane, the crystal $X$ axis been oriented along the sum $\mathbf{p}_{\rm F}^A + \mathbf{p}_{\rm F}^B$ \cite{note}.
The components of the linear optical susceptibility of the crystal correspondingly are:
\begin{eqnarray}
\label{eq:chixxnew}
\chi_{XX} = -\frac{p_{\rm F}^2}{v}\frac{2}{\alpha_1 \alpha_2 - 2\alpha_{12}^2} \left[ \alpha_2 \cos^2 \varphi \right. \nonumber \\
\left.
{}+ 4a\alpha_{12}\cos^2 \varphi \cos \gamma + a^2 \alpha_1 (1 + \cos 2\gamma \cos \varphi) \right]
\end{eqnarray}
and
\begin{eqnarray}
\label{eq:chiyy}
\chi_{YY} = -\frac{p_{\rm F}^2}{v}\frac{2}{\alpha_1 \alpha_2 - 2\alpha_{12}^2} \left[ \alpha_2 \sin^2 \varphi \right. \nonumber \\
\left.
{}+ 4a\alpha_{12}\sin^2 \varphi \cos \gamma + a^2 \alpha_1 (1 - \cos 2\gamma \cos \varphi) \right].
\end{eqnarray}
We find the following expressions for functions $\alpha_1$, $\alpha_2$, $\alpha_{12}$:
(1) In the excitonic region expressions practically coincide with formulas in Ref.~\cite{lalov08}, namely
\begin{equation}
\label{eq:alpha1}
\alpha_1 = \hbar \omega - \left( E_{\rm F} + 2L \right) - \hbar \Omega_{\rm 0F}(1),
\end{equation}
\begin{equation}
\label{eq:alpha12}
\alpha_{12} = \varepsilon_{\rm e}^{\prime \prime} + \varepsilon_{\rm h}^{\prime \prime},
\end{equation}
\begin{equation}
\label{eq:alpha2}
\alpha_2 = \hbar \left[ \omega - \Omega_{\rm 0c}(1) \right] - E_{\rm c},
\end{equation}
where $\Omega_{\rm 0F}(1)$ and $\Omega_{\rm 0c}(1)$ are expressed through the continuous fractions following from recursions:
\begin{equation}
\label{eq:omega0f}
\Omega_{\rm 0F}(n) = \frac{n\omega_{\rm a}^2}{\omega - \left( E_{\rm F} + 2L \right)/\hbar - n\omega_0 - \Omega_{\rm 0F}(n+1)},
\end{equation}
\begin{equation}
\label{eq:omega0c}
\Omega_{\rm 0c}(n) = \frac{2n\omega_{\rm a1}^2}{\omega - E_{\rm c}/\hbar - n\omega_0 - \Omega_{\rm 0c}(n+1)},
\end{equation}
\begin{equation}
\label{eq:omegaa}
\omega_{\rm a} = \xi_{\rm F}\omega_0, \qquad \omega_{\rm a1} = \xi \omega_0.
\end{equation}
(2) In the one-phonon vibronic region
\begin{equation}
\label{eq:alfa1}
\alpha_1 = \hbar \omega - \left( E_{\rm F} + 2L \right) - \frac{\omega_{\rm a}^2}{mD} \left[ (\hbar \omega_{\rm 1c} - \beta)\sigma - \frac{\beta \left( \varepsilon_{\rm e}^{\prime \prime} + \varepsilon_{\rm h}^{\prime \prime} \right)^2}{m}\sigma_1 \right],
\end{equation}
\begin{equation}
\label{eq:alfa12}
\alpha_{12} = \left( \varepsilon_{\rm e}^{\prime \prime} + \varepsilon_{\rm h}^{\prime \prime} \right)\left[ 1 + \frac{\omega_{\rm a} \omega_{\rm a1}}{mD}\left( \sigma + \sigma_1 \right) \right],
\end{equation}
\begin{eqnarray}
\label{eq:alfa2}
\alpha_2 = \hbar \omega - E_{\rm c} - \frac{2\omega_{\rm a1}^2}{mD} \left\{ \frac{m}{\omega_{\rm 1c}} - \alpha_{\rm F}\sigma + \left( \varepsilon_{\rm e}^{\prime \prime} + \varepsilon_{\rm h}^{\prime \prime} \right)^2 \right. \nonumber \\
\left. \times \! \left[ \frac{\sigma + \sigma_1 (2 + t)}{\omega_{\rm 1c}} - \frac{\alpha_{\rm F}\sigma_1}{m} \right] \right\},
\end{eqnarray}
where
\begin{equation}
\label{eq:omega1c}
\omega_{\rm 1c} = \omega - \omega_0 - \Omega_{\rm 1c}(1) - E_{\rm c}/\hbar,
\end{equation}
\vspace{1mm}
\begin{equation}
\label{eq:m}
m = 2\left[ L\hbar \omega_{\rm 1c} + 2 \varepsilon_{\rm e}^{\prime \prime} \varepsilon_{\rm h}^{\prime \prime} \right],
\end{equation}
\begin{eqnarray}
\label{eq:t}
t = (1/m)\left\{
\left\{ \hbar \left[ \omega - \omega_0 - \Omega_{\rm 1F}(1) \right]- E_{\rm F} \right\} \hbar \omega_{\rm 1c}
- 2\left[ \left( \varepsilon_{\rm e}^{\prime \prime} \right)^2 + \left( \varepsilon_{\rm h}^{\prime \prime} \right)^2 \right] \right\},
\end{eqnarray}
\begin{eqnarray}
\label{eq:D}
D = \left( 1 - \frac{\alpha_{\rm F} \omega_{\rm 1c} \sigma}{m} \right) \left( 1 - \frac{\beta}{\omega_{\rm 1c}} \right) \nonumber \\
{}+ \frac{\beta \left( \varepsilon_{\rm e}^{\prime \prime} + \varepsilon_{\rm h}^{\prime \prime} \right)^2}{m} \left[ \frac{\alpha_{\rm F} \sigma_1}{m} - \frac{\sigma + \sigma_1(2 + t)}{\omega_{\rm 1c}} \right],
\end{eqnarray}
\vspace{2mm}
\begin{equation}
\label{eq:sigma}
\sigma = \left\{ \begin{array}{cc}
-1/\sqrt{t^2 - 1} & \mbox{if\,\,\, $\left| t + \sqrt{t^2 - 1} \right| < 1$,}
\\ \\
1/\sqrt{t^2 - 1} & \mbox{if\,\,\, $\left| t + \sqrt{t^2 - 1} \right| < 1$,}
\end{array}
\right.
\end{equation}
and
\begin{equation}
\label{eq:sigma1}
\sigma_1 = t\sigma -1.
\end{equation}
The functions $\Omega_{\rm 1F}(1)$ and $\Omega_{\rm 1c}(1)$ also represent continuous fractions from recursions:
\begin{equation}
\label{eq:omega1f}
\Omega_{\rm 1F}(n) = \frac{n\omega_{\rm a}^2}{\omega - E_{\rm F}/\hbar - (n+1)\omega_0 - \Omega_{\rm 1F}(n+1)},
\end{equation}
\begin{equation}
\label{eq:Omega1c}
\Omega_{\rm 1c}(n) = \frac{2n\omega_{\rm a1}^2}{\omega - E_{\rm c}/\hbar - (n+1)\omega_0 - \Omega_{\rm 1c}(n+1)}.
\end{equation}
Finally, the functions $\beta$ and $\alpha_{\rm F}$ can be expressed as follows:
\begin{equation}
\label{eq:beta/h}
\beta/\hbar = \Omega_{\rm 0c}(2) - \Omega_{\rm 1c}(1),
\end{equation}
\begin{equation}
\label{eq:alphaf/h}
\alpha_{\rm F}/\hbar = \Omega_{\rm 0F}(2) - \Omega_{\rm 1F}(1).
\end{equation}
\section{Simulations of the Excitonic and Vibronic Spectra of MePTCDI and PTCDA Crystals}
\label{sec:simul}
In this section, we calculate the absorption spectra of the two crystals finding the imaginary parts of the components of the linear optical susceptibility (\ref{eq:chixxnew}) at $\left( 2p_{\rm F}^2/v \right) = 1$ and supposing an imaginary part equal to $\mathrm{i}\delta/\hbar$ of the frequency $\omega$. We put the excitonic and vibrational parameters for the studied crystals as they have been fitted in Refs.~\cite{henessy99} and \cite{schmidt02} and used in our previous papers \cite{lalov06a,lalov07a,lalov06b,lalov06c}, see table 1:
\begin{table}
\caption{\label{data}Excitonic and vibrational parameters (in cm$^{-1}$) of the MePTCDI and PTCDA crystals.}{\smallskip}
\begin{small}\centering
\begin{tabular}{lccccrccc}
\hline \noalign {\smallskip}
& $E_{\rm F}$ & $E_{\rm c}$ & $\hbar \omega_0$ & $L$ & $\varepsilon_{\rm e}$ & $\varepsilon_{\rm h}$ & $2\varphi$ & $\gamma$ \\
\hline \noalign
{\smallskip}
MePTCDI & $17\,992$ & $17\,346$ & $1\,400$ & $345$ & $-380$ & $-137$ & $36.8^\circ$ & $68.1^\circ$ \\
PTCDA & $18\,860$ & $18\,300$ & $1\,400$ & $330$ & $-48$ & $-436$ & $82^\circ$ & $143^\circ$ \\
\hline
\end{tabular}
\end{small}
\end{table}
The data for angle $\gamma$ and ratio $a = \left| p_{\rm CT}/p_{\rm F} \right|$ have been calculated in \cite{hoffmann00} using quantum chemical evaluations. We use the values $a = 0.1$ for the PTCDA crystal and $a = 0.135$ for the MePTCDI. The data for angle $2\varphi$ are derived from the crystal structure (for MePTCDI see \cite{haedicke86}).
The aforementioned parameters are permanent in our calculations. We vary the values of the following parameters (intending to observe their impact on the absorption spectra and find a better similarity with the experimental absorption spectra \cite{henessy99,hoffmann00,hoffmann02}):
(i) The excitonic damping quantity $\delta$ which varies from $1$ to $500$ cm$^{-1}$.
(ii) The linear exciton--phonon coupling parameters $\xi_{\rm F}$ and $\xi$ which vary near the values
\begin{eqnarray*}
\xi_{\rm F} \!\!\!\!\!&=&\!\!\!\!\! 0.82 \mbox{ and } 1.1 \quad \mbox{for PTCDA} \quad \mbox{(see \cite{henessy99})},
\\
\xi_{\rm F} \!\!\!\!\!&=&\!\!\!\!\! 0.88 \mbox{ and } 1.1 \quad \mbox{for MePTCDI} \quad \mbox{(see \cite{hoffmann02})},
\end{eqnarray*}
and
\[
\xi = \xi_{\rm F}, \mbox{ }\xi_{\rm F}/\sqrt{2}, \mbox{ and }\; \xi_{\rm F}/2 \;\, \mbox{for both crystals}.
\]
In calculating continuous fractions (\ref{eq:omega0f}), (\ref{eq:omega0c}) and (\ref{eq:omega1f}), (\ref{eq:Omega1c}) we take twenty steps.
\subsection{MePTCDI}
\label{subsec:meptcdi}
The fine structure of the excitonic spectra of this crystal is presented in Figure~1. The absorption curves are calculated at small value $\delta = 20$ cm$^{-1}$ of the excitonic damping for $\xi_{\rm F} = 0.88$ with $\xi = \xi_{\rm F}$ (blue curve 1), $\xi = \xi_{\rm F}/\sqrt{2}$ (red curve 2) and $\xi = \xi_{\rm F}/2$ (green curve 3). We have calculated the imaginary part of expression~(\ref{eq:chixxnew}) using the `excitonic' formulas (\ref{eq:alpha1})--(\ref{eq:alpha2}) (in the following denoted as ``exc'' program). For the smaller values of $\xi$ equal to $\xi_{\rm F}/\sqrt{2}$ and $\xi_{\rm F}/2$ we obtain spectral doublets which correspond to the CTEs--FE splitting. This splitting generates a spectral triplet at $15\,100$ cm$^{-1}$ (very weak), $16\,500$ cm$^{-1}$, and $17\,200$ cm$^{-1}$ (for $\xi = \xi_{\rm F}$, blue curve 1).
\begin{figure}[!ht]
\centering\includegraphics[height=.25\textheight]{lalov-fg1}
\caption{Linear absorption in the excitonic region of the MePTCDI crystal at $\delta = 20$ cm$^{-1}$, $\xi_{\rm F} = 0.88$ with $\xi = \xi_{\rm F}$ (blue curve 1), $\xi = \xi_{\rm F}/\sqrt{2}$ (red curve 2) and $\xi = \xi_{\rm F}/2$ (green curve 3) calculated using formulas (24)--(26) (``exc'' program).}
\label{fig:fg1}
\end{figure}
\begin{figure}[!ht]
\centering\includegraphics[height=.25\textheight]{lalov-fg2}
\caption{Linear absorption in the excitonic region of the MePTCDI crystal at $\delta = 300$ cm$^{-1}$ (``exc'' program). For the values of $\xi_{\rm F}$ and $\xi$ see Figure~\ref{fig:fg1}.}
\label{fig:fg2}
\end{figure}
\begin{figure}[!ht]
\centering\includegraphics[height=.25\textheight]{lalov-fg3}
\caption{Impact of the CTEs-transition dipole moment expressed through the ratio $a$. Red curve 1 for $a = 0.135$ and the green curve 2 for $a = 0$ ($\delta = 300$ cm$^{-1}$, $\xi_{\rm F} = 0.88$, $\xi = \xi_{\rm F}/2$).}
\label{fig:fg3}
\end{figure}
The value $\delta = 300$ cm$^{-1}$ seems to be more closer to the width of the lowest absorption maximum observed in the MePTCDI crystal \cite{hoffmann00,hoffmann02}. The absorption curves in Figure~2 are calculated at $\delta = 300$ cm$^{-1}$ in the excitonic region (approximately below $17\,800$ cm$^{-1}$), one-phonon vibronic region ($17\,800$--$19\,200$ cm$^{-1}$), two-phonon vibronic region ($19\,200$--$20\,700$ cm$^{-1}$) etc. There exists an obvious similarity between experimental curve (see Ref.~\cite{hoffmann00}) and the green curve 3 in Figure~2 calculated with $\xi = \xi_{\rm F}/2$.
The impact of the CTEs transition dipole moment is illustrated in Figure~3. The two curves -- the red one 1 calculated with the CTEs contribution, and the green one 2 calculated without this contribution [$a \equiv 0$ in formulas (\ref{eq:chixx})--(\ref{eq:chiyy})] -- are relatively close to each other. The strongest impact can be measured near lower excitonic peak generated by the CTEs level $E_{\rm c} = 17\,346$ cm$^{-1}$.
\begin{figure}[!ht]
\centering\includegraphics[height=.25\textheight]{lalov-fg4}
\caption{Comparison of the absorption curves calculated with the data from Table~1 (green curve 1) and the red curve 2 calculated for $E_{\rm F} = 17\,346$ cm$^{-1}$ and $E_{\rm c} = 17\,992$ cm$^{-1}$ (interchange of the excitonic levels) with $\delta = 300$ cm$^{-1}$, $\xi_{\rm F} = 0.88$, $\xi = \xi_{\rm F}/2$.}
\label{fig:fg4}
\end{figure}
The two curves in Figure~4 calculated by using the ``exc'' program for $\delta = 300$ cm$^{-1}$ differ by the mutual positions of the two excitonic levels. The green curve 1 corresponds to a lower position of the CTEs level $E_{\rm c} < E_{\rm F}$, whereas the red curve 2 corresponds to the opposite situation $E_{\rm F} < E_{\rm c}$ (the magnitudes of $E_{\rm c}$ and $E_{\rm F}$ are exchanged). The red curve 2 exhibits a strong domination of the lower maxima associated with the FE level. The two curves can be approximated with five Lorentz maxima in the spectral region of $15\,000$--$22\,000$ cm$^{-1}$ \cite{hoffmann00}. Comparing the lineshapes of these two curves with the experimental curve in Ref.~\cite{hoffmann00}, we cannot make a hypothesis which possibility is more probable. We prefer the fitting from Refs.~\cite{hoffmann00} and \cite{schmidt02} $\left( E_{\rm c} < E_{\rm F} \right)$, however, another choice, $ E_{\rm F} < E_{\rm c}$, is also allowed.
\begin{figure}[!ht]
\centering\includegraphics[height=.25\textheight]{lalov-fg5}
\caption{The impact of the FE--phonon coupling constant $\xi_{\rm F}$ on the linear absorption. The red curve 1 corresponds to $\xi_{\rm F} = 0.88$, and the green curve 2 to $\xi_{\rm F} = 1.1$ ($\xi = \xi_{\rm F}/2$, $\delta = 300$ cm$^{-1}$).}
\label{fig:fg5}
\end{figure}
\begin{figure}[!ht]
\centering\includegraphics[height=.25\textheight]{lalov-fg6}
\caption{Linear absorption spectra of the MePTCDI crystal with FE--CTEs mixing (red curve 1) and pure Frenkel-exciton spectra (blue curve 2 at $\varepsilon_{\rm e}^{\prime \prime} = \varepsilon_{\rm h}^{\prime \prime} = 0$) obtained by the ``exc'' program with $\delta = 300$ cm$^{-1}$, $\xi_{\rm F} = 0.88$, $\xi = \xi_{\rm F}/2$.}
\label{fig:fg6}
\end{figure}
Figure~5 illustrates the impact of the FE--phonon coupling constant $\xi_{\rm F}$ on the absorption spectra. The values $\xi_{\rm F} = 0.88$ and $\xi = \xi_{\rm F}/2$ can be considered as better candidates for the simulation (compared with the experimental absorption lineshape especially in the excitonic region).
The importance of the FE--CTEs mixing can be seen in Figure~6 in which the blue curve 2 represents a pure FE-absorption $\left( \varepsilon_{\rm e}^{\prime \prime} = \varepsilon_{\rm h}^{\prime \prime} = 0 \right)$. The complexity of the experimental absorption curves with five Lorentz maxima in the studied spectral region, see Figures~2 and 3 in Ref.~\cite{hoffmann00}, cannot be understood on the basis of simple Frenkel-exciton model. Contrary, the model of mixed FE--CTEs and their vibronic satellites can be the basis of adequate simulations of the absorption curves (red curve 1).
\begin{figure}[!ht]
\centering\includegraphics[height=.25\textheight]{lalov-fg7}
\caption{One-phonon vibronic spectra (``1p'' program) of the MePTCDI crystal with $\xi_{\rm F} = 0.88$, $\xi = \xi_{\rm F}/\sqrt{2}$, $\delta = 1$ cm$^{-1}$ (red curve 1), $\delta = 2$ cm$^{-1}$ (green curve 2), and $\delta = 300$ cm$^{-1}$ (purple curve 3).}
\label{fig:fg7}
\end{figure}
The absorption curves in the following three Figures~7--9 have been calculated by using formulas (\ref{eq:omega0f})--(\ref{eq:alphaf/h}) for one-phonon vibronic spectra denoted in the following as ``1p'' program. Our calculations based on the Green functions formalism allow to study and simulate the absorption associated with the two types of exciton--phonon states in the vibronic spectra, notably: (a) bound states (one-particle states) corresponding to the propagation of excitons and phonon(s) in the stack as a whole. The lineshape of their absorption maxima is Lorentzian and strongly depends on the excitonic damping $\delta$. (b) Many-particle states (MP) corresponding to the excitation of exciton and phonon(s) on separate molecules of the stack. Their energy lies in the quasicontinuous band(s) and the lineshape of the corresponding absorption maxima depends on the density of the states in the MP band and more weakly on the excitonic damping $\delta$. Unfortunately the big values of $\delta$ in both crystals under consideration mask the MP bands which are totally covered by the absorption maxima. We can simulate the absorption in MP continua by supposing very small values of the damping, for instance, $\delta \sim 1$--$10$ cm$^{-1}$, which is not typical for the perylene derivatives (but occurring in some DA-crystals like An-PMDA \cite{sebastian81,sebastian83,haarer75,brillante80}).
\begin{figure}[ht]
\begin{minipage}[b]{0.5\linewidth}
\centering
\includegraphics[width=5.5cm]{lalov-fg8a}
\end{minipage}
\hspace{-0.2cm}
\begin{minipage}[b]{0.5\linewidth}
\centering
\includegraphics[width=5.5cm]{lalov-fg8b}
\end{minipage}
\caption{(a) MP band in two-phonon vibronic spectra with $\xi_{\rm F} = 0.88$, $\xi = \xi_{\rm F}/\sqrt{2}$, $\delta = 1$ cm$^{-1}$ (red curve 1), $\delta = 10$ cm$^{-1}$ (blue curve 2). (b) Bound exciton--phonon state in one-phonon vibronic spectra with $\xi_{\rm F} = 0.88$, $\xi = \xi_{\rm F}/\sqrt{2}$, $\delta = 1$ cm$^{-1}$ (red curve 1), $\delta = 2$ cm$^{-1}$ (green curve 2), $\delta = 10$ cm$^{-1}$ (blue curve 3).}
\label{fig:fg8}
\end{figure}
\begin{figure}[!ht]
\centering\includegraphics[height=.25\textheight]{lalov-fg9}
\caption{Linear absorption in the vibronic spectra of the MePTCDI crystal (``1p'' program) with FE--CTEs mixing (red curve 1) and pure FE vibronics (green curve 2 at $\varepsilon_{\rm e}^{\prime \prime} = \varepsilon_{\rm h}^{\prime \prime} = 0$) with $\delta = 300$ cm$^{-1}$, $\xi_{\rm F} = 0.88$, $\xi = \xi_{\rm F}/2$.}
\label{fig:fg9}
\end{figure}
The one-phonon vibronic spectra are depicted in Figure~7. The vibronics of FE and CTEs are mixed like their excitons' spectra. Nevertheless, the less intensive lower maximum at $17\,500$--$17\,700$ cm$^{-1}$ can be considered as vibronic replica of CTEs whereas the more intensive maximum near $18\,400$--$18\,700$ cm$^{-1}$ is the replica of the FE. The lower wings of both maxima represent the very weak MP bands (even virtual ones). The main part of absorption intensity is concentrated above those MP bands in the Lorentzian maxima near $17\,600$ cm$^{-1}$ and $18\,650$ cm$^{-1}$. As can be seen in Figure~7, the wide absorption maxima corresponding to the realistic value $\delta = 300$ cm$^{-1}$ (purple curve 3) cover the fine structure of the absorption lineshape.
Figure~8 illustrates the difference between the MP band (left panel (a)) and the bound exciton--phonon maximum (right panel (b)). The line shape of the MP band depends of the density of the states which is strongly modulated by the exciton--phonon coupling. Thus, the lineshape is non-symmetrical. In the case of relatively small damping parameter $\delta$ the lineshape (look at Figure~8(a)) depends very weakly on $\delta$. Contrary, the maximal value of the Lorentzian maximum in Figure~8(b) is proportional to $1/\delta$.
Figure~9 is analog of Figure~6 in the vibronic region (above $17\,500$ cm$^{-1}$). The lineshape of the red curve 1 (with FE--CTEs mixing) is more complicated than the lineshape of the pure FE vibronics. The red curve 1 seems to be a better simulation of the experimental absorption curve \cite{hoffmann00}.
The model of the FE--CTEs mixing reproduces the general structure of the absorption spectra in the excitonic and vibronic regions of the MePTCDI crystal (see Figures~2 and 9). Our calculations confirm the correct choice of the excitonic and vibrational parameters in the fitting procedure implemented in Refs.~\cite{hoffmann00,schmidt02}, especially the value $\xi_{\rm F} = 0.88$ of the constant of the linear FE--phonon coupling. Additionally, we establish as the most probable value of the exciton damping $\delta = 300$ cm$^{-1}$ and CTEs--phonon coupling constant $\xi = \xi_{\rm F}/2$. Our studies show the positions of the bands both of FE$+$phonon and CTEs$+$phonon many-particle states. However, the wide absorption lines cover the virtual MP bands.
\subsection{PTCDA}
\label{subsec:ptcda}
The experimental linear absorption lines for the PTCDA crystal are broader than those of the MePTCDI \cite{hoffmann00}. Thus, the structure of the PTCDA spectra does not exhibit many details generated by the FE--CTEs mixing.
Our calculations show the appearance of the spectral doublets in the excitonic and vibronic spectra associated with the FE--CTEs splitting (red curve 1 in Figure~10). However, the probable value of the excitonic damping parameter is $\delta = 500$ cm$^{-1}$ evaluated using the width of the lowest experimental absorption maximum \cite{hoffmann00}. The blue curve 2 in Figure~10 calculated with $\delta = 500$ cm$^{-1}$ exhibits four Lorentz-type maxima as it is observed in the experimental absorption spectra of the PTCDA crystal \cite{hoffmann00}. The fifth maximum near $17\,000$ cm$^{-1}$ (see Figure~10) manifests itself as a weak peculiarity only of the absorption curve with $\delta = 500$ cm$^{-1}$.
\begin{figure}[!ht]
\centering\includegraphics[height=.25\textheight]{lalov-fg10}
\caption{Linear absorption of the PTCDA crystal calculated using the ``exc'' program with $\xi_{\rm F} = 0.82$, $\xi = \xi_{\rm F}/2$, $\delta = 10$ cm$^{-1}$ (red curve 1), and $\delta = 500$ cm$^{-1}$ (blue curve 2).}
\label{fig:fg10}
\end{figure}
\begin{figure}[!ht]
\centering\includegraphics[height=.25\textheight]{lalov-fg11}
\caption{Linear absorption of the PTCDA crystal (``exc'' program) with $\xi_{\rm F} = 1.1$, $\delta = 500$ cm$^{-1}$,
$\xi = \xi_{\rm F}/2$ (green curve 1) and $\xi = \xi_{\rm F}$ (blue curve 2). The purple curve 3 corresponds to pure FE spectra ($\varepsilon_{\rm e}^{\prime \prime} = \varepsilon_{\rm h}^{\prime \prime} = 0$).}
\label{fig:fg11}
\end{figure}
\begin{figure}[!ht]
\centering\includegraphics[height=.25\textheight]{lalov-fg12}
\caption{Linear absorption of the PTCDA crystal (``exc'' program) with $\xi_{\rm F} = 1.1$, $\xi = \xi_{\rm F}/2$, and $\delta = 500$ cm$^{-1}$. The green curve 1 corresponds to $E_{\rm F} = 17\,992$ cm$^{-1}$, $E_{\rm c} = 17\,346$ cm$^{-1}$, the blue curve 2 to pure FE spectra ($\varepsilon_{\rm e}^{\prime \prime} = \varepsilon_{\rm h}^{\prime \prime} = 0$), and the red curve 3 to $E_{\rm F} = 17\,346$ cm$^{-1}$, $E_{\rm c} = 17\,992$ cm$^{-1}$.}
\label{fig:fg12}
\end{figure}
Moreover, further calculations implemented with $\xi_{\rm F} = 1.1$ (which value seems to be the most suitable to our simulations) demonstrate insensitivity of the absorption spectra on the value of $\xi$ (see the very close green curve 1 and blue curve 2 in Figure~11). We also calculated the absorption curve neglecting the FE--CTEs mixing (purple curve 3). The result consists of some re-distributions of the absorption intensity among four maxima but the absorption lineshape is practically non-changed in the cases with and without FE--CTEs mixing.
The same effect -- the lineshape being not strongly affected by a hypothetical FE--CTEs mixing -- can be seen in Figure~12 (green curve 1 and blue curve 2 there). The mutual replacement of the two excitonic levels (compare the red and green curves) shifts the position of the absorption maxima, however, their lineshapes are very similar.
\begin{figure}[ht]
\begin{minipage}[b]{0.5\linewidth}
\centering
\includegraphics[width=5.3cm]{lalov-fg13a}
\end{minipage}
\hspace{-0.2cm}
\begin{minipage}[b]{0.5\linewidth}
\centering
\includegraphics[width=5.6cm]{lalov-fg13b}
\end{minipage}
\caption{One-phonon vibronic spectra of the PTCDA crystal (``1p'' program) with (a) $\xi_{\rm F} = 0.82$, $\xi = \xi_{\rm F}/2$, $\delta = 1$ cm$^{-1}$ (red curve 1), $\delta = 10$ cm$^{-1}$ (green curve 2) and (b) $\xi_{\rm F} = 0.82$, $\xi = \xi_{\rm F}/2$, and $\varepsilon_{\rm e}^{\prime \prime} = \varepsilon_{\rm h}^{\prime \prime} = 0$ (vibronic spectra of FE). The red curve 1 corresponds to $\delta = 1$ cm$^{-1}$, the blue curve 2 to $\delta = 10$ cm$^{-1}$, and the purple curve 3 to $\delta = 500$ cm$^{-1}$.}
\label{fig:fg13}
\end{figure}
\begin{figure}[!ht]
\centering\includegraphics[height=.25\textheight]{lalov-fg14}
\caption{One-phonon and two-phonon vibronic spectra of the PTCDA crystal (``1p'' program) with $\xi_{\rm F} = 1.1$, $\xi = \xi_{\rm F}/2$. The blue curve 1 corresponds to $\delta = 1$ cm$^{-1}$ and the purple curve 2 to $\delta = 500$ cm$^{-1}$.}
\label{fig:fg14}
\end{figure}
The absorption curves in Figures~13 and 14 have been calculated using the ``1p'' program and they show the positions and the structure of one-phonon and two-phonon vibronic spectra. Surely, the fine structure can be observed in the hypothetical case of small values of $\delta$. Then the structure of the MP bands depends on the FE--CTEs mixing. Figure~13(a) containing two absorption regions, $18\,600$--$19\,200$ cm$^{-1}$ and $19\,400$--$19\,900$ cm$^{-1}$, each one with a weak MP band and a Lorentzian. This picture is rather different in the case of the absence of FE--CTEs mixing (Figure~13(b)). Then we might observe a CTEs vibronic maximum near $18\,800$ cm$^{-1}$ and the vibronic replica of FE with MP band ($18\,900$--$19\,400$ cm$^{-1}$), as well as a Lorentz-type maximum at $19\,600$ cm$^{-1}$. In our model Hamiltonian (\ref{eq:cteshamilt}) the dispersion of pure CTEs has been neglected and thus at $\varepsilon_{\rm e}^{\prime \prime} = \varepsilon_{\rm h}^{\prime \prime} = 0$ the vibronic replica of CTEs consists of one-particle maximum only. In the same time, FEs possess dispersion expressed through the transfer terms with $L$ in expression (\ref{eq:fehamilt}) (see also \cite{hoffmann00,schmidt02}) and in the vibronic absorption spectra of FEs both one-particle maxima and MP band could appear. As in other cases, a strong damping with $\delta = 500$ cm$^{-1}$ masks the whole structure.
The absorption curves in Figure~14 correspond to a higher value of $\xi_{\rm F} = 1.1$. The one-phonon vibronic spectra ($18\,500$--$19\,000$ cm$^{-1}$) and two-phonon vibronic spectra ($19\,700$--$20\,500$ cm$^{-1}$) contain MP bands ($18\,300$--$18\,500$ cm$^{-1}$ and $19\,700$--$20\,000$ cm$^{-1}$) and several other maxima. Their splitting is caused by the relatively big value of the constant $\xi_{\rm F}$ of the linear FE--phonon coupling. In the realistic case of strong damping, $\delta = 500$ cm$^{-1}$, the one-phonon and two-phonon vibronic spectra would contain two broad and asymmetric maxima only.
The main conclusion of our simulations of the linear absorption in the PTCDA crystal is that the hypothesis of a FE--CTEs mixing is not crucial in describing the observed spectra. The wide absorption lines in the excitonic and vibronic spectra cover the effects of the mixing and the hypothesis of Frenkel excitons' spectra and the vibronic of FE could be sufficient (see Ref.~\cite{vragovic03}).
\section{Conclusion}
\label{sec:concl}
Our model for the linear absorption spectra of the one-component charge-transfer molecular crystals includes the following parameters:
\begin{itemize}
\itemsep 0pt \parskip 0pt
\item excitonic levels $E_{\rm F}$ and $E_{\rm c}$, as well as the vibrational frequency $\omega_0$ of the intramolecular mode,
\item parameters $\varepsilon_{\rm e}^{\prime \prime}$ and $\varepsilon_{\rm h}^{\prime \prime}$ of the FE--CTEs mixing,
\item constants $\xi_{\rm F}$ and $\xi$ of the linear exciton--phonon coupling,
\item the width $\delta$ of the excitonic linewidth,
\item angles $2\varphi$, $\gamma$ and the ratio $a$ of the CTEs and FE transition dipole moments.
\end{itemize}
Practically all parameters have been introduced and fitted for MePTCDI and PTCDA crystals in previous papers \cite{henessy99,hoffmann00,hoffmann02,schmidt02}, based primarily on the matrix diagonalization method. In our study we apply the complex vibronic approach based on the Green functions formalism in calculating the linear optical susceptibility and its imaginary part which is a factor in the absorption coefficient in the excitonic and one-phonon vibronic spectra. Higher vibronics -- with two, three phonons -- also have been demonstrated in our calculations. Calculated vibronic spectra consist of one-particle Lorentzian maxima and many-particle (MP) bands which correspond to unbound propagation of the excitons and phonons. Our simulations expose the positions of the MP continua but the wide excitonic and vibronic absorption lines in both crystals cover the fine structure of the vibronic spectra.
The main goal of our model is to simulate the lineshape in the absorption region of $15\,000$--$23\,000$ cm$^{-1}$ of the aforementioned crystals, and to find out better fitted values of the excitonic linewidth, exciton--phonon coupling parameters and so on.
We stress again the main results which concern the two crystals:
\begin{enumerate}
\itemsep 0pt \parskip 0pt
\item The excitonic linewidth of the MePTCDI crystal, according to our simulations, is approximately $\delta \approx 300$ cm$^{-1}$. The FE-phonon linear coupling coefficient $\xi_{\rm F}$ has been estimated correctly in the previous papers \cite{henessy99,hoffmann02} as $\xi_{\rm F} \approx 0.88$ and $\xi = \xi_{\rm F}/2$. The absorption spectra depend strongly on the mutual position of the two excitonic levels $E_{\rm F}$ and $E_{\rm c}$. A supposition for the FE--CTEs mixing \emph{is necessary\/} for an adequate interpretation of the excitonic and vibronic spectra of the MePTCDI crystal.
\item The excitonic linewidth of the PTCDA crystal can be evaluated as $\delta \approx 500$ cm$^{-1}$. That is why the effect of the FE--CTEs mixing, being covered by the wide absorption maxima, are more weakly expressed than in the MePTCDI crystal. However, a very probable conclusion from our calculations may be the stronger linear exciton--phonon coupling ($\xi_{\rm F} \approx 1$ or $1.1$ instead of $0.82$).
\end{enumerate}
Our model can be applied in the interpretation of other one-component molecular stacks (crystals). It can be more effective in the systems with narrower excitonic absorption lines where both types of vibronic states -- one-particle states and MP continua -- calculated by using the Green functions formalism would be seen in the linear absorption.
|
\section*{Abstract (Not appropriate in this style!)}%
\else \small
\begin{center}{\bf Abstract\vspace{-.5em}\vspace{\zeta@}}\end{center}%
\quotation
\fi
}%
}{%
}%
\@ifundefined{endabstract}{\def\endabstract
{\if@twocolumn\else\endquotation\fi}}{}%
\@ifundefined{maketitle}{\def\maketitle#1{}}{}%
\@ifundefined{affiliation}{\def\affiliation#1{}}{}%
\@ifundefined{proof}{\def\proof{\noindent{\bfseries Proof. }}}{}%
\@ifundefined{endproof}{\def\endproof{\mbox{\ \rule{.1in}{.1in}}}}{}%
\@ifundefined{newfield}{\def\newfield#1#2{}}{}%
\@ifundefined{chapter}{\def\chapter#1{\par(Chapter head:)#1\par }%
\newcount\chi@chapter}{}%
\@ifundefined{part}{\def\part#1{\par(Part head:)#1\par }}{}%
\@ifundefined{section}{\def\section#1{\par(Section head:)#1\par }}{}%
\@ifundefined{subsection}{\def\subsection#1%
{\par(Subsection head:)#1\par }}{}%
\@ifundefined{subsubsection}{\def\subsubsection#1%
{\par(Subsubsection head:)#1\par }}{}%
\@ifundefined{paragraph}{\def\paragraph#1%
{\par(Subsubsubsection head:)#1\par }}{}%
\@ifundefined{subparagraph}{\def\subparagraph#1%
{\par(Subsubsubsubsection head:)#1\par }}{}%
\@ifundefined{therefore}{\def\therefore{}}{}%
\@ifundefined{backepsilon}{\def\backepsilon{}}{}%
\@ifundefined{yen}{\def\yen{\hbox{\rm\rlap=Y}}}{}%
\@ifundefined{registered}{%
\def\registered{\relax\ifmmode{}\rho@gistered
\else$\mu@th\rho@gistered$\fi}%
\def\rho@gistered{^{\ooalign
{\hfil\raise.07ex\hbox{$\scriptstyle\rm\RIfM@\expandafter\text@\else\expandafter\mbox\fi{R}$}\hfil\crcr
\mathhexbox20D}}}}{}%
\@ifundefined{Eth}{\def\Eth{}}{}%
\@ifundefined{eth}{\def\eth{}}{}%
\@ifundefined{Thorn}{\def\Thorn{}}{}%
\@ifundefined{thorn}{\def\thorn{}}{}%
\def\TEXTsymbol#1{\mbox{$#1$}}%
\@ifundefined{degree}{\def\degree{{}^{\circ}}}{}%
\newdimen\theight
\@ifundefined{Column}{\def\Column{%
\vadjust{\setbox\zeta@=\hbox{\scriptsize\quad\quad tcol}%
\theight=\ht\zeta@\advance\theight by \dp\zeta@\advance\theight by \lineskip
\kern -\theight \vbox to \theight{%
\rightline{\rlap{\box\zeta@}}%
\vss
}%
}%
}}{}%
\@ifundefined{qed}{\def\qed{%
\ifhmode\unskip\nobreak\fi\ifmmode\ifinner\else\hskip5\psi@\fi\fi
\hbox{\hskip5\psi@\vrule width4\psi@ height6\psi@ depth1.5\psi@\hskip\psi@}%
}}{}%
\@ifundefined{cents}{\def\cents{\hbox{\rm\rlap c/}}}{}%
\@ifundefined{tciLaplace}{\def\tciLaplace{\ensuremath{\mathcal{L}}}}{}%
\@ifundefined{tciFourier}{\def\tciFourier{\ensuremath{\mathcal{F}}}}{}%
\@ifundefined{textcurrency}{\def\textcurrency{\hbox{\rm\rlap xo}}}{}%
\@ifundefined{texteuro}{\def\texteuro{\hbox{\rm\rlap C=}}}{}%
\@ifundefined{euro}{\def\euro{\hbox{\rm\rlap C=}}}{}%
\@ifundefined{textfranc}{\def\textfranc{\hbox{\rm\rlap-F}}}{}%
\@ifundefined{textlira}{\def\textlira{\hbox{\rm\rlap L=}}}{}%
\@ifundefined{textpeseta}{\def\textpeseta{\hbox{\rm P\negthinspace s}}}{}%
\@ifundefined{miss}{\def\miss{\hbox{\vrule height2\psi@ width 2\psi@ depth\zeta@}}}{}%
\@ifundefined{vvert}{\def\vvert{\Vert}}{
\@ifundefined{tcol}{\def\tcol#1{{\baselineskip=6\psi@ \vcenter{#1}} \Column}}{}%
\@ifundefined{dB}{\def\dB{\hbox{{}}}}{
\@ifundefined{mB}{\def\mB#1{\hbox{$#1$}}}{
\@ifundefined{nB}{\def\nB#1{\hbox{#1}}}{
\@ifundefined{note}{\def\note{$^{\dag}}}{}%
\defLaTeX2e{LaTeX2e}
\ifx\fmtnameLaTeX2e
\DeclareOldFontCommand{\rm}{\normalfont\rmfamily}{\mathrm}
\DeclareOldFontCommand{\sf}{\normalfont\sffamily}{\mathsf}
\DeclareOldFontCommand{\tt}{\normalfont\ttfamily}{\mathtt}
\DeclareOldFontCommand{\bf}{\normalfont\bfseries}{\mathbf}
\DeclareOldFontCommand{\it}{\normalfont\itshape}{\mathit}
\DeclareOldFontCommand{\sl}{\normalfont\slshape}{\@nomath\sl}
\DeclareOldFontCommand{\sc}{\normalfont\scshape}{\@nomath\sc}
\fi
\def\alpha{{\Greekmath 010B}}%
\def\beta{{\Greekmath 010C}}%
\def\gamma{{\Greekmath 010D}}%
\def\delta{{\Greekmath 010E}}%
\def\epsilon{{\Greekmath 010F}}%
\def\zeta{{\Greekmath 0110}}%
\def\eta{{\Greekmath 0111}}%
\def\theta{{\Greekmath 0112}}%
\def\iota{{\Greekmath 0113}}%
\def\kappa{{\Greekmath 0114}}%
\def\lambda{{\Greekmath 0115}}%
\def\mu{{\Greekmath 0116}}%
\def\nu{{\Greekmath 0117}}%
\def\xi{{\Greekmath 0118}}%
\def\pi{{\Greekmath 0119}}%
\def\rho{{\Greekmath 011A}}%
\def\sigma{{\Greekmath 011B}}%
\def\tau{{\Greekmath 011C}}%
\def\upsilon{{\Greekmath 011D}}%
\def\phi{{\Greekmath 011E}}%
\def\chi{{\Greekmath 011F}}%
\def\psi{{\Greekmath 0120}}%
\def\omega{{\Greekmath 0121}}%
\def\varepsilon{{\Greekmath 0122}}%
\def\vartheta{{\Greekmath 0123}}%
\def\varpi{{\Greekmath 0124}}%
\def\varrho{{\Greekmath 0125}}%
\def\varsigma{{\Greekmath 0126}}%
\def\varphi{{\Greekmath 0127}}%
\def{\Greekmath 0272}{{\Greekmath 0272}}
\def\FindBoldGroup{%
{\setbox0=\hbox{$\mathbf{x\global\edef\theboldgroup{\the\mathgroup}}$}}%
}
\def\Greekmath#1#2#3#4{%
\if@compatibility
\ifnum\mathgroup=\symbold
\mathchoice{\mbox{\boldmath$\displaystyle\mathchar"#1#2#3#4$}}%
{\mbox{\boldmath$\textstyle\mathchar"#1#2#3#4$}}%
{\mbox{\boldmath$\scriptstyle\mathchar"#1#2#3#4$}}%
{\mbox{\boldmath$\scriptscriptstyle\mathchar"#1#2#3#4$}}%
\else
\mathchar"#1#2#3#
\fi
\else
\FindBoldGroup
\ifnum\mathgroup=\theboldgroup
\mathchoice{\mbox{\boldmath$\displaystyle\mathchar"#1#2#3#4$}}%
{\mbox{\boldmath$\textstyle\mathchar"#1#2#3#4$}}%
{\mbox{\boldmath$\scriptstyle\mathchar"#1#2#3#4$}}%
{\mbox{\boldmath$\scriptscriptstyle\mathchar"#1#2#3#4$}}%
\else
\mathchar"#1#2#3#
\fi
\fi}
\newif\ifGreekBold \GreekBoldfalse
\let\SAVEPBF=\pbf
\def\pbf{\GreekBoldtrue\SAVEPBF}%
\@ifundefined{theorem}{\newtheorem{theorem}{Theorem}}{}
\@ifundefined{lemma}{\newtheorem{lemma}[theorem]{Lemma}}{}
\@ifundefined{corollary}{\newtheorem{corollary}[theorem]{Corollary}}{}
\@ifundefined{conjecture}{\newtheorem{conjecture}[theorem]{Conjecture}}{}
\@ifundefined{proposition}{\newtheorem{proposition}[theorem]{Proposition}}{}
\@ifundefined{axiom}{\newtheorem{axiom}{Axiom}}{}
\@ifundefined{remark}{\newtheorem{remark}{Remark}}{}
\@ifundefined{example}{\newtheorem{example}{Example}}{}
\@ifundefined{exercise}{\newtheorem{exercise}{Exercise}}{}
\@ifundefined{definition}{\newtheorem{definition}{Definition}}{}
\@ifundefined{mathletters}{%
\newcounter{equationnumber}
\def\mathletters{%
\addtocounter{equation}{1}
\edef\@currentlabel{\arabic{equation}}%
\setcounter{equationnumber}{\chi@equation}
\setcounter{equation}{0}%
\edef\arabic{equation}{\@currentlabel\noexpand\alph{equation}}%
}
\def\endmathletters{%
\setcounter{equation}{\value{equationnumber}}%
}
}{}
\@ifundefined{BibTeX}{%
\def\BibTeX{{\rm B\kern-.05em{\sc i\kern-.025em b}\kern-.08em
T\kern-.1667em\lower.7ex\hbox{E}\kern-.125emX}}}{}%
\@ifundefined{AmS}%
{\def\AmS{{\protect\usefont{OMS}{cmsy}{m}{n}%
A\kern-.1667em\lower.5ex\hbox{M}\kern-.125emS}}}{}%
\@ifundefined{AmSTeX}{\def\AmSTeX{\protect\AmS-\protect\TeX\@}}{}%
\def\@@eqncr{\let\@tempa\relax
\ifcase\@eqcnt \def\@tempa{& & &}\or \def\@tempa{& &}%
\else \def\@tempa{&}\fi
\@tempa
\if@eqnsw
\iftag@
\@taggnum
\else
\@eqnnum\stepcounter{equation}%
\fi
\fi
\global\@ifnextchar*{\@tagstar}{\@tag}@false
\global\@eqnswtrue
\global\@eqcnt\zeta@\cr}
\def\@ifnextchar*{\@TCItagstar}{\@TCItag}{\@ifnextchar*{\@TCItagstar}{\@TCItag}}
\def\@TCItag#1{%
\global\@ifnextchar*{\@tagstar}{\@tag}@true
\global\def\@taggnum{(#1)}%
\global\def\@currentlabel{#1}}
\def\@TCItagstar*#1{%
\global\@ifnextchar*{\@tagstar}{\@tag}@true
\global\def\@taggnum{#1}%
\global\def\@currentlabel{#1}}
\def\QATOP#1#2{{#1 \atop #2}}%
\def\QTATOP#1#2{{\textstyle {#1 \atop #2}}}%
\def\QDATOP#1#2{{\displaystyle {#1 \atop #2}}}%
\def\QABOVE#1#2#3{{#2 \above#1 #3}}%
\def\QTABOVE#1#2#3{{\textstyle {#2 \above#1 #3}}}%
\def\QDABOVE#1#2#3{{\displaystyle {#2 \above#1 #3}}}%
\def\QOVERD#1#2#3#4{{#3 \overwithdelims#1#2 #4}}%
\def\QTOVERD#1#2#3#4{{\textstyle {#3 \overwithdelims#1#2 #4}}}%
\def\QDOVERD#1#2#3#4{{\displaystyle {#3 \overwithdelims#1#2 #4}}}%
\def\QATOPD#1#2#3#4{{#3 \atopwithdelims#1#2 #4}}%
\def\QTATOPD#1#2#3#4{{\textstyle {#3 \atopwithdelims#1#2 #4}}}%
\def\QDATOPD#1#2#3#4{{\displaystyle {#3 \atopwithdelims#1#2 #4}}}%
\def\QABOVED#1#2#3#4#5{{#4 \abovewithdelims#1#2#3 #5}}%
\def\QTABOVED#1#2#3#4#5{{\textstyle
{#4 \abovewithdelims#1#2#3 #5}}}%
\def\QDABOVED#1#2#3#4#5{{\displaystyle
{#4 \abovewithdelims#1#2#3 #5}}}%
\def\tint{\msi@int\textstyle\int}%
\def\tiint{\msi@int\textstyle\iint}%
\def\tiiint{\msi@int\textstyle\iiint}%
\def\tiiiint{\msi@int\textstyle\iiiint}%
\def\tidotsint{\msi@int\textstyle\idotsint}%
\def\toint{\msi@int\textstyle\oint}%
\def\tsum{\mathop{\textstyle \sum }}%
\def\tprod{\mathop{\textstyle \prod }}%
\def\tbigcap{\mathop{\textstyle \bigcap }}%
\def\tbigwedge{\mathop{\textstyle \bigwedge }}%
\def\tbigoplus{\mathop{\textstyle \bigoplus }}%
\def\tbigodot{\mathop{\textstyle \bigodot }}%
\def\tbigsqcup{\mathop{\textstyle \bigsqcup }}%
\def\tcoprod{\mathop{\textstyle \coprod }}%
\def\tbigcup{\mathop{\textstyle \bigcup }}%
\def\tbigvee{\mathop{\textstyle \bigvee }}%
\def\tbigotimes{\mathop{\textstyle \bigotimes }}%
\def\tbiguplus{\mathop{\textstyle \biguplus }}%
\newtoks\temptoksa
\newtoks\temptoksb
\newtoks\temptoksc
\def\msi@int#1#2{%
\def\@temp{{#1#2\the\temptoksc_{\the\temptoksa}^{\the\temptoksb}}
\futurelet\@nextcs
\@int
}
\def\@int{%
\ifx\@nextcs\limits
\typeout{Found limits}%
\temptoksc={\limits}%
\let\@next\@intgobble%
\else\ifx\@nextcs\nolimits
\typeout{Found nolimits}%
\temptoksc={\nolimits}%
\let\@next\@intgobble%
\else
\typeout{Did not find limits or no limits}%
\temptoksc={}%
\let\@next\msi@limits%
\fi\fi
\@next
}%
\def\@intgobble#1{%
\typeout{arg is #1}%
\msi@limits
}
\def\msi@limits{%
\temptoksa={}%
\temptoksb={}%
\@ifnextchar_{\@limitsa}{\@limitsb}%
}
\def\@limitsa_#1{%
\temptoksa={#1}%
\@ifnextchar^{\@limitsc}{\@temp}%
}
\def\@limitsb{%
\@ifnextchar^{\@limitsc}{\@temp}%
}
\def\@limitsc^#1{%
\temptoksb={#1}%
\@ifnextchar_{\@limitsd}{\@temp
}
\def\@limitsd_#1{%
\temptoksa={#1}%
\@temp
}
\def\dint{\msi@int\displaystyle\int}%
\def\diint{\msi@int\displaystyle\iint}%
\def\diiint{\msi@int\displaystyle\iiint}%
\def\diiiint{\msi@int\displaystyle\iiiint}%
\def\didotsint{\msi@int\displaystyle\idotsint}%
\def\doint{\msi@int\displaystyle\oint}%
\def\dsum{\mathop{\displaystyle \sum }}%
\def\dprod{\mathop{\displaystyle \prod }}%
\def\dbigcap{\mathop{\displaystyle \bigcap }}%
\def\dbigwedge{\mathop{\displaystyle \bigwedge }}%
\def\dbigoplus{\mathop{\displaystyle \bigoplus }}%
\def\dbigodot{\mathop{\displaystyle \bigodot }}%
\def\dbigsqcup{\mathop{\displaystyle \bigsqcup }}%
\def\dcoprod{\mathop{\displaystyle \coprod }}%
\def\dbigcup{\mathop{\displaystyle \bigcup }}%
\def\dbigvee{\mathop{\displaystyle \bigvee }}%
\def\dbigotimes{\mathop{\displaystyle \bigotimes }}%
\def\dbiguplus{\mathop{\displaystyle \biguplus }}%
\if@compatibility\else
\RequirePackage{amsmath}
\fi
\def\makeatother\endinput{\makeatother\endinput}
\bgroup
\ifx\ds@amstex\relax
\message{amstex already loaded}\aftergroup\makeatother\endinput
\else
\@ifpackageloaded{amsmath}%
{\if@compatibility\message{amsmath already loaded}\fi\aftergroup\makeatother\endinput}
{}
\@ifpackageloaded{amstex}%
{\if@compatibility\message{amstex already loaded}\fi\aftergroup\makeatother\endinput}
{}
\@ifpackageloaded{amsgen}%
{\if@compatibility\message{amsgen already loaded}\fi\aftergroup\makeatother\endinput}
{}
\fi
\egroup
\typeout{TCILATEX defining AMS-like constructs in LaTeX 2.09 COMPATIBILITY MODE}
\let\DOTSI\relax
\def\RIfM@{\relax\ifmmode}%
\def\FN@{\futurelet\next}%
\newcount\intno@
\def\iint{\DOTSI\intno@\tw@\FN@\ints@}%
\def\iiint{\DOTSI\intno@\thr@@\FN@\ints@}%
\def\iiiint{\DOTSI\intno@4 \FN@\ints@}%
\def\idotsint{\DOTSI\intno@\zeta@\FN@\ints@}%
\def\ints@{\findlimits@\ints@@}%
\newif\iflimtoken@
\newif\iflimits@
\def\findlimits@{\limtoken@true\ifx\next\limits\limits@true
\else\ifx\next\nolimits\limits@false\else
\limtoken@false\ifx\ilimits@\nolimits\limits@false\else
\ifinner\limits@false\else\limits@true\fi\fi\fi\fi}%
\def\multint@{\int\ifnum\intno@=\zeta@\intdots@
\else\intkern@\fi
\ifnum\intno@>\tw@\int\intkern@\fi
\ifnum\intno@>\thr@@\int\intkern@\fi
\int
\def\multintlimits@{\intop\ifnum\intno@=\zeta@\intdots@\else\intkern@\fi
\ifnum\intno@>\tw@\intop\intkern@\fi
\ifnum\intno@>\thr@@\intop\intkern@\fi\intop}%
\def\intic@{%
\mathchoice{\hskip.5em}{\hskip.4em}{\hskip.4em}{\hskip.4em}}%
\def\negintic@{\mathchoice
{\hskip-.5em}{\hskip-.4em}{\hskip-.4em}{\hskip-.4em}}%
\def\ints@@{\iflimtoken@
\def\ints@@@{\iflimits@\negintic@
\mathop{\intic@\multintlimits@}\limits
\else\multint@\nolimits\fi
\eat@
\else
\def\ints@@@{\iflimits@\negintic@
\mathop{\intic@\multintlimits@}\limits\else
\multint@\nolimits\fi}\fi\ints@@@}%
\def\intkern@{\mathchoice{\!\!\!}{\!\!}{\!\!}{\!\!}}%
\def\plaincdots@{\mathinner{\cdotp\cdotp\cdotp}}%
\def\intdots@{\mathchoice{\plaincdots@}%
{{\cdotp}\mkern1.5mu{\cdotp}\mkern1.5mu{\cdotp}}%
{{\cdotp}\mkern1mu{\cdotp}\mkern1mu{\cdotp}}%
{{\cdotp}\mkern1mu{\cdotp}\mkern1mu{\cdotp}}}%
\def\RIfM@{\relax\protect\ifmmode}
\def\RIfM@\expandafter\text@\else\expandafter\mbox\fi{\RIfM@\expandafter\RIfM@\expandafter\text@\else\expandafter\mbox\fi@\else\expandafter\mbox\fi}
\let\nfss@text\RIfM@\expandafter\text@\else\expandafter\mbox\fi
\def\RIfM@\expandafter\text@\else\expandafter\mbox\fi@#1{\mathchoice
{\textdef@\displaystyle\phi@size{#1}}%
{\textdef@\textstyle\tf@size{\firstchoice@false #1}}%
{\textdef@\textstyle\sf@size{\firstchoice@false #1}}%
{\textdef@\textstyle \ssf@size{\firstchoice@false #1}}%
\glb@settings}
\def\textdef@#1#2#3{\hbox{{%
\everymath{#1}%
\let\phi@size#2\selectfont
#3}}}
\newif\iffirstchoice@
\firstchoice@true
\def\Let@{\relax\iffalse{\fi\let\\=\cr\iffalse}\fi}%
\def\vspace@{\def\vspace##1{\crcr\noalign{\vskip##1\relax}}}%
\def\multilimits@{\bgroup\vspace@\Let@
\baselineskip\fontdimen10 \scriptfont\tw@
\advance\baselineskip\fontdimen12 \scriptfont\tw@
\lineskip\thr@@\fontdimen8 \scriptfont\thr@@
\lineskiplimit\lineskip
\vbox\bgroup\ialign\bgroup\hfil$\mu@th\scriptstyle{##}$\hfil\crcr}%
\def\Sb{_\multilimits@}%
\def\endSb{\crcr\egroup\egroup\egroup}%
\def\Sp{^\multilimits@}%
\let\endSp\endSb
\newdimen\ex@
\ex@.2326ex
\def\rightarrowfill@#1{$#1\mu@th\mathord-\mkern-6mu\cleaders
\hbox{$#1\mkern-2mu\mathord-\mkern-2mu$}\hfill
\mkern-6mu\mathord\rightarrow$}%
\def\leftarrowfill@#1{$#1\mu@th\mathord\leftarrow\mkern-6mu\cleaders
\hbox{$#1\mkern-2mu\mathord-\mkern-2mu$}\hfill\mkern-6mu\mathord-$}%
\def\leftrightarrowfill@#1{$#1\mu@th\mathord\leftarrow
\mkern-6mu\cleaders
\hbox{$#1\mkern-2mu\mathord-\mkern-2mu$}\hfill
\mkern-6mu\mathord\rightarrow$}%
\def\overrightarrow{\mathpalette\overrightarrow@}%
\def\overrightarrow@#1#2{\vbox{\ialign{##\crcr\rightarrowfill@#1\crcr
\noalign{\kern-\ex@\nointerlineskip}$\mu@th\hfil#1#2\hfil$\crcr}}}%
\let\overarrow\overrightarrow
\def\overleftarrow{\mathpalette\overleftarrow@}%
\def\overleftarrow@#1#2{\vbox{\ialign{##\crcr\leftarrowfill@#1\crcr
\noalign{\kern-\ex@\nointerlineskip}$\mu@th\hfil#1#2\hfil$\crcr}}}%
\def\overleftrightarrow{\mathpalette\overleftrightarrow@}%
\def\overleftrightarrow@#1#2{\vbox{\ialign{##\crcr
\leftrightarrowfill@#1\crcr
\noalign{\kern-\ex@\nointerlineskip}$\mu@th\hfil#1#2\hfil$\crcr}}}%
\def\underrightarrow{\mathpalette\underrightarrow@}%
\def\underrightarrow@#1#2{\vtop{\ialign{##\crcr$\mu@th\hfil#1#2\hfil
$\crcr\noalign{\nointerlineskip}\rightarrowfill@#1\crcr}}}%
\let\underarrow\underrightarrow
\def\underleftarrow{\mathpalette\underleftarrow@}%
\def\underleftarrow@#1#2{\vtop{\ialign{##\crcr$\mu@th\hfil#1#2\hfil
$\crcr\noalign{\nointerlineskip}\leftarrowfill@#1\crcr}}}%
\def\underleftrightarrow{\mathpalette\underleftrightarrow@}%
\def\underleftrightarrow@#1#2{\vtop{\ialign{##\crcr$\mu@th
\hfil#1#2\hfil$\crcr
\noalign{\nointerlineskip}\leftrightarrowfill@#1\crcr}}}%
\def\qopnamewl@#1{\mathop{\operator@font#1}\nlimits@}
\let\nlimits@\displaylimits
\def\setboxz@h{\setbox\zeta@\hbox}
\def\varlim@#1#2{\mathop{\vtop{\ialign{##\crcr
\hfil$#1\mu@th\operator@font lim$\hfil\crcr
\noalign{\nointerlineskip}#2#1\crcr
\noalign{\nointerlineskip\kern-\ex@}\crcr}}}}
\def\rightarrowfill@#1{\mu@th\setboxz@h{$#1-$}\ht\zeta@\zeta@
$#1\copy\zeta@\mkern-6mu\cleaders
\hbox{$#1\mkern-2mu\box\zeta@\mkern-2mu$}\hfill
\mkern-6mu\mathord\rightarrow$}
\def\leftarrowfill@#1{\mu@th\setboxz@h{$#1-$}\ht\zeta@\zeta@
$#1\mathord\leftarrow\mkern-6mu\cleaders
\hbox{$#1\mkern-2mu\copy\zeta@\mkern-2mu$}\hfill
\mkern-6mu\box\zeta@$}
\def\qopnamewl@{proj\,lim}{\qopnamewl@{proj\,lim}}
\def\qopnamewl@{inj\,lim}{\qopnamewl@{inj\,lim}}
\def\mathpalette\varlim@\rightarrowfill@{\mathpalette\varlim@\rightarrowfill@}
\def\mathpalette\varlim@\leftarrowfill@{\mathpalette\varlim@\leftarrowfill@}
\def\mathpalette\varliminf@{}{\mathpalette\mathpalette\varliminf@{}@{}}
\def\mathpalette\varliminf@{}@#1{\mathop{\underline{\vrule\@depth.2\ex@\@width\zeta@
\hbox{$#1\mu@th\operator@font lim$}}}}
\def\mathpalette\varlimsup@{}{\mathpalette\mathpalette\varlimsup@{}@{}}
\def\mathpalette\varlimsup@{}@#1{\mathop{\overline
{\hbox{$#1\mu@th\operator@font lim$}}}}
\def\stackunder#1#2{\mathrel{\mathop{#2}\limits_{#1}}}%
\begingroup \catcode `|=0 \catcode `[= 1
\catcode`]=2 \catcode `\{=12 \catcode `\}=12
\catcode`\\=12
|gdef|@alignverbatim#1\end{align}[#1|end[align]]
|gdef|@salignverbatim#1\end{align*}[#1|end[align*]]
|gdef|@alignatverbatim#1\end{alignat}[#1|end[alignat]]
|gdef|@salignatverbatim#1\end{alignat*}[#1|end[alignat*]]
|gdef|@xalignatverbatim#1\end{xalignat}[#1|end[xalignat]]
|gdef|@sxalignatverbatim#1\end{xalignat*}[#1|end[xalignat*]]
|gdef|@gatherverbatim#1\end{gather}[#1|end[gather]]
|gdef|@sgatherverbatim#1\end{gather*}[#1|end[gather*]]
|gdef|@gatherverbatim#1\end{gather}[#1|end[gather]]
|gdef|@sgatherverbatim#1\end{gather*}[#1|end[gather*]]
|gdef|@multilineverbatim#1\end{multiline}[#1|end[multiline]]
|gdef|@smultilineverbatim#1\end{multiline*}[#1|end[multiline*]]
|gdef|@arraxverbatim#1\end{arrax}[#1|end[arrax]]
|gdef|@sarraxverbatim#1\end{arrax*}[#1|end[arrax*]]
|gdef|@tabulaxverbatim#1\end{tabulax}[#1|end[tabulax]]
|gdef|@stabulaxverbatim#1\end{tabulax*}[#1|end[tabulax*]]
|endgroup
\def\align{\@verbatim \frenchspacing\@vobeyspaces \@alignverbatim
You are using the "align" environment in a style in which it is not defined.}
\let\endalign=\endtrivlist
\@namedef{align*}{\@verbatim\@salignverbatim
You are using the "align*" environment in a style in which it is not defined.}
\expandafter\let\csname endalign*\endcsname =\endtrivlist
\def\alignat{\@verbatim \frenchspacing\@vobeyspaces \@alignatverbatim
You are using the "alignat" environment in a style in which it is not defined.}
\let\endalignat=\endtrivlist
\@namedef{alignat*}{\@verbatim\@salignatverbatim
You are using the "alignat*" environment in a style in which it is not defined.}
\expandafter\let\csname endalignat*\endcsname =\endtrivlist
\def\xalignat{\@verbatim \frenchspacing\@vobeyspaces \@xalignatverbatim
You are using the "xalignat" environment in a style in which it is not defined.}
\let\endxalignat=\endtrivlist
\@namedef{xalignat*}{\@verbatim\@sxalignatverbatim
You are using the "xalignat*" environment in a style in which it is not defined.}
\expandafter\let\csname endxalignat*\endcsname =\endtrivlist
\def\gather{\@verbatim \frenchspacing\@vobeyspaces \@gatherverbatim
You are using the "gather" environment in a style in which it is not defined.}
\let\endgather=\endtrivlist
\@namedef{gather*}{\@verbatim\@sgatherverbatim
You are using the "gather*" environment in a style in which it is not defined.}
\expandafter\let\csname endgather*\endcsname =\endtrivlist
\def\multiline{\@verbatim \frenchspacing\@vobeyspaces \@multilineverbatim
You are using the "multiline" environment in a style in which it is not defined.}
\let\endmultiline=\endtrivlist
\@namedef{multiline*}{\@verbatim\@smultilineverbatim
You are using the "multiline*" environment in a style in which it is not defined.}
\expandafter\let\csname endmultiline*\endcsname =\endtrivlist
\def\arrax{\@verbatim \frenchspacing\@vobeyspaces \@arraxverbatim
You are using a type of "array" construct that is only allowed in AmS-LaTeX.}
\let\endarrax=\endtrivlist
\def\tabulax{\@verbatim \frenchspacing\@vobeyspaces \@tabulaxverbatim
You are using a type of "tabular" construct that is only allowed in AmS-LaTeX.}
\let\endtabulax=\endtrivlist
\@namedef{arrax*}{\@verbatim\@sarraxverbatim
You are using a type of "array*" construct that is only allowed in AmS-LaTeX.}
\expandafter\let\csname endarrax*\endcsname =\endtrivlist
\@namedef{tabulax*}{\@verbatim\@stabulaxverbatim
You are using a type of "tabular*" construct that is only allowed in AmS-LaTeX.}
\expandafter\let\csname endtabulax*\endcsname =\endtrivlist
\def\endequation{%
\ifmmode\ifinner
\iftag@
\addtocounter{equation}{-1}
$\hfil
\displaywidth\linewidth\@taggnum\egroup \endtrivlist
\global\@ifnextchar*{\@tagstar}{\@tag}@false
\global\@ignoretrue
\else
$\hfil
\displaywidth\linewidth\@eqnnum\egroup \endtrivlist
\global\@ifnextchar*{\@tagstar}{\@tag}@false
\global\@ignoretrue
\fi
\else
\iftag@
\addtocounter{equation}{-1}
\eqno \hbox{\@taggnum}
\global\@ifnextchar*{\@tagstar}{\@tag}@false%
$$\global\@ignoretrue
\else
\eqno \hbox{\@eqnnum
$$\global\@ignoretrue
\fi
\fi\fi
}
\newif\iftag@ \@ifnextchar*{\@tagstar}{\@tag}@false
\def\@ifnextchar*{\@TCItagstar}{\@TCItag}{\@ifnextchar*{\@TCItagstar}{\@TCItag}}
\def\@TCItag#1{%
\global\@ifnextchar*{\@tagstar}{\@tag}@true
\global\def\@taggnum{(#1)}%
\global\def\@currentlabel{#1}}
\def\@TCItagstar*#1{%
\global\@ifnextchar*{\@tagstar}{\@tag}@true
\global\def\@taggnum{#1}%
\global\def\@currentlabel{#1}}
\@ifundefined{tag}{
\def\@ifnextchar*{\@tagstar}{\@tag}{\@ifnextchar*{\@tagstar}{\@tag}}
\def\@tag#1{%
\global\@ifnextchar*{\@tagstar}{\@tag}@true
\global\def\@taggnum{(#1)}}
\def\@tagstar*#1{%
\global\@ifnextchar*{\@tagstar}{\@tag}@true
\global\def\@taggnum{#1}}
}{}
\def\tfrac#1#2{{\textstyle {#1 \over #2}}}%
\def\dfrac#1#2{{\displaystyle {#1 \over #2}}}%
\def\binom#1#2{{#1 \choose #2}}%
\def\tbinom#1#2{{\textstyle {#1 \choose #2}}}%
\def\dbinom#1#2{{\displaystyle {#1 \choose #2}}}%
\makeatother
\endinput
\section{Introduction}
Let $\left( M,\omega \right) $ be a compact symplectic manifold.
This is a sequel to \cite{oh-zhu} in which we study the adiabatic
degeneration of maps $u:{\mathbb{R}}\times S^{1}\rightarrow M$ satisfying
the following 1-parameter ($0<\varepsilon <\varepsilon _{0}$) family of
Floer equations:
\begin{equation}
(du+P_{K_{\varepsilon }}(u))_{J_{\varepsilon }}^{(0,1)}=0\quad
\mbox{ or
equivalently }\,{\overline{\partial }}_{J_{\varepsilon
}}(u)+(P_{K_{\varepsilon }})_{J_{\varepsilon }}^{(0,1)}(u)=0, \label{eq:KJe}
\end{equation}%
We refer to section \ref{sec:floer} for detailed description of the
one-parameter family Hamiltonian $K_{\varepsilon }$ and almost complex
structure $J_{\varepsilon }$, and the equation \eqref{eq:KJe}, the invariant
form of the Floer equation. The expression of the degenerating Hamiltonian $%
K_{\varepsilon }:{\mathbb{R}}\times S^{1}\times M\rightarrow {\mathbb{R}}$
has the form
\begin{equation}
K_{\varepsilon }(\tau ,t,x)=%
\begin{cases}
\kappa _{\varepsilon }^{+}(\tau )\cdot H(t,x)\quad & \mbox{for }\,\tau \geq
R(\varepsilon ) \\
\kappa_{\varepsilon }^0(\tau )\cdot \varepsilon f(x)\quad & \mbox{for }%
\,|\tau |\leq R(\varepsilon ) \\
\kappa _{\varepsilon }^{-}(\tau )\cdot H(t,x)\quad & \mbox{for }\,\tau \leq
-R(\varepsilon )%
\end{cases}
\label{eq:KR}
\end{equation}%
where $\kappa _{\varepsilon }^{\pm }$ and $\kappa_{\varepsilon }^0$ are
suitable cut-off functions (See subsection \ref{sec:floer} for the precise
definition.)
Roughly speaking, the adiabatic degeneration occurs because $K_{\varepsilon
} $ restricts to Morse function $\varepsilon f$ on longer and longer
cylinder $[-R(\varepsilon ),R(\varepsilon )]\times S^{1}$ in ${\mathbb{R}}%
\times S^{1}$. A basic assumption that we had put on \cite{oh-zhu} was that $%
R(\varepsilon )$ satisfies
\begin{equation}
\lim_{\varepsilon \rightarrow 0}\varepsilon R(\varepsilon )=0.
\label{eq:length=0}
\end{equation}
The main purpose of the present paper is to prove a gluing result when we
have the \emph{non-zero} limit
\begin{equation*}
\lim_{\varepsilon \rightarrow 0}\varepsilon R(\varepsilon )=\ell
\end{equation*}%
for $\ell >0$ as promised in \cite{oh-zhu}. Under this assumption, it is
proved in \cite{oh:dmj} and \cite{MT} that as $\varepsilon \rightarrow 0$, a
degenerating sequence of Floer trajectories converges to a \textquotedblleft
\emph{disk-flow-disk\textquotedblright } configuration denoted by $%
(u_{-},\chi ,u_{+})$ where $u_{\pm }$ satisfy the equation similar to %
\eqref{eq:KJe} with $K_{\varepsilon }$ is replaced by $K_{\pm }$ defined by
\begin{equation}
K_+ =
\begin{cases}
0 & \RIfM@\expandafter\text@\else\expandafter\mbox\fi{near}\;o_{+} \\
H_{+}(t,x)\,dt & \RIfM@\expandafter\text@\else\expandafter\mbox\fi{near}\;e_{+}%
\end{cases}
\label{eq:Kplus}
\end{equation}
and
\begin{equation}
K_- =
\begin{cases}
0 & \RIfM@\expandafter\text@\else\expandafter\mbox\fi{near}\;o_{-} \\
H_{-}(t,x)\,dt & \RIfM@\expandafter\text@\else\expandafter\mbox\fi{near}\;e_{-}%
\end{cases}
\label{eq:Kminus}
\end{equation}%
Here $H_{\pm }:S^{1}\times M\rightarrow {\mathbb{R}}$ is a Hamiltonian
function independent of the variable $\tau $, and $\chi $ is a gradient
trajectory of $f$ that satisfies $\dot{\chi}+\RIfM@\expandafter\text@\else\expandafter\mbox\fi{grad}f(\chi )=$ on $%
[0,\ell ]$ such that
\begin{equation}
u_{-}(o_{-})=\chi (-l),\quad \chi (l)=u_{+}(o_{+}).
\label{eq:nodalmathching}
\end{equation}
Let $\dot{\Sigma}_+$ be the Riemann sphere with one marked point $o_+$ and
one positive puncture $e_+$. Choose analytical charts at $o_+$ and at $e_+$
on some neighborhoods $O_+$ and $E_+$ respectively, so that conformally $%
O_+\backslash o_+ \cong (-\infty,0]\times S^1$, and $E_+\backslash e_+\cong
[0,+\infty)\times S^1$. We use $t$ for the $S^1$ coordinate and $\tau$ for
the ${\mathbb{R}}$ coordinate. Then $\{-\infty\}\times S^1$ and $%
\{+\infty\}\times S^1$ correspond to $o_+$ and $e_+$ respectively.
Let $z_\pm:S^1 \to M$ be a nondegenerate periodic orbit of $H_\pm$ and
consider a finite energy solution $u_\pm: \dot \Sigma \to M$ of the Floer
equation \eqref{eq:Kplus}, \eqref{eq:Kminus} associated to $K_\pm$
respectively. By the finite energy condition and since $K_\pm \equiv 0$ near
$o_\pm$, $u_\pm$ extend smoothly across $o_\pm$ and can be regarded as a
smooth map defined on ${\mathbb{C}}$ that is holomorphic near the origin $0
\in {\mathbb{C}}$.
Now we consider the lifting $[z_\pm,w_\pm]$ of $z_\pm$ and introduce the
main moduli spaces of our interest in a precise term.
Let $J_{\pm }$ be a pair of domain-dependent almost complex structures on $M$
and let $u_{\pm }:{\mathbb{R}}\times S^{1}\rightarrow M$ be solutions of %
\eqref{eq:KJe} with $K_{\varepsilon }$ replaced by $K_{\pm }$ given in %
\eqref{eq:Kplus}, \eqref{eq:Kminus} in class $A_{\pm }\in \pi _{2}(z_{\pm })$
with a marked point $o_{\pm }\in S^{2}$ respectively, and $\chi
:[-l,l]\rightarrow M$ is a gradient segment of the Morse function $f$
connecting the two points $u_{+}(o_{+})$ and $u_{-}\left( o_{-}\right) $. We
assume $(J_{\pm })$ satisfy $J_{\pm }\equiv J_{0}$ near the punctures $%
o_{\pm }$ respectively and generic in that $u_{\pm }$ are Fredholm regular.
We also assume that the pair $(J_{0},f)$ is generic in that $\chi $ are
Fredholm regular and the configuration $(u_{-},\chi ,u_{+})$ satisfies the
\textquotedblleft disc-flow-disc\textquotedblright\ transversality defined
in \cite{oh-zhu} Proposition 5.2. We will recall this definition below.
Define the moduli space
\begin{eqnarray*}
\mathcal{M}\left( K_{\pm },J_{\pm };z_{\pm };A_{\pm }\right) &=&\Big\{\left(
u_{\pm },o_{\pm }\right) |u_{\pm }:S^{2}\setminus \{e_{\pm },o_{\pm
}\}\rightarrow M \\
&{}&\quad \mbox{satisfying \eqref{eq:Kplus}, \eqref{eq:Kminus}
respectively}\Big\}
\end{eqnarray*}%
and the evaluation map%
\begin{equation*}
ev_{\pm }:\mathcal{M}\left( K_{\pm },J_{\pm };z_{\pm };A_{\pm }\right)
\rightarrow M,\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ \ }u_{\pm }\rightarrow ev_{\pm }\left( o_{\pm }\right) .
\end{equation*}
\begin{defn}
\label{dfd-trans}The configuration $(u_{-},\chi ,u_{+})$ satisfies the
\textquotedblleft disc-flow-disc\textquotedblright\ transversality if the
map
\begin{eqnarray*}
\phi _{f}^{2l}\circ ev_{-}\times ev_{+} &:&\mathcal{M}\left(
K_{-},J_{-};z_{-};A_{-}\right) \times \mathcal{M}\left(
K_{+},J_{+};z_{+};A_{+}\right) \rightarrow M\times M \\
\left( u_{-},u_{+}\right) &\rightarrow &\left( \phi _{f}^{2l}u_{-}\left(
o_{-}\right) ,u_{+}\left( o_{+}\right) \right)
\end{eqnarray*}%
is transversal to the diagonal $\triangle \subset M\times M$, where $\phi
_{f}^{2l}$ is the time-$2l$ flow of the Morse function $f$.
\end{defn}
The \textquotedblleft disc-flow-disc\textquotedblright\ transversality can
be achieved by a generic choice of the pair $(J_{0},f)$ (Corollary \ref%
{achieve-dfd-trans}).
We note that for each given pair $(A_{-},A_{+})\in \pi _{2}(z_{-})\times \pi
_{2}(z_{+})$ we consider and \emph{fix} a class $B \in \pi _{2}(z_{-},z_{+})$
such that
\begin{equation} \label{eq:A-AA+}
A_-\# B \# A_+ = 0 \quad \mbox{in }\, \pi_2(M).
\end{equation}
(Such $B \in \pi _{2}(z_{-},z_{+})$ is unique modulo the action of $\pi_1$).
We define
\begin{equation*}
{\mathcal{M}}^{dfd}(K_{-},J_{-};f;K_{+},J_{+};B)
\end{equation*}%
to be the fiber product
\begin{equation}
\mathcal{M}\left( K_{-},J_{-};z_{-};A_{-}\right) {}_{\phi _{f}^{l}\circ
ev_{-}}\times _{ev_{+}}\mathcal{M}\left( K_{+},J_{+};z_{+};A_{+}\right) .
\label{eq:K-fK+}
\end{equation}%
We denote by
\begin{equation*}
{\mathcal{M}}^{\varepsilon }:={\mathcal{M}}(K_{\varepsilon },J_{\varepsilon
};z_{-},z_{+};B)
\end{equation*}%
the moduli space of solutions to \eqref{eq:KJe} and consider
\begin{equation*}
{\mathcal{M}}_{(0,\varepsilon _{0}]}^{para}:=\bigcup_{0<\varepsilon \leq
\varepsilon _{0}}{\mathcal{M}}^{\varepsilon }.
\end{equation*}%
In this paper, we prove the following main theorem:
\begin{thm}
${\mathcal{M}}_{(0,\varepsilon _{0})}^{para}$ can be embedded into a
manifold with boundary $\overline{\mathcal{M}}_{[0,\varepsilon _{0})}^{para}$
whose boundary is diffeomorphic to ${\mathcal{M}}%
(K_{-},J_{-};f;K_{+},J_{+};B)$, i.e., there is a diffeomorphism
\begin{equation*}
Glue:[0,\varepsilon _{0})\times {\mathcal{M}}%
^{dfd}(K_{-},J_{-};f;K_{+},J_{+};B)\rightarrow \overline{\mathcal{M}}%
_{[0,\varepsilon _{0})}^{para}
\end{equation*}%
such that the following diagram commutes:
\begin{equation*}
\xymatrix{{\mathcal M}^{para}_{(0,\varepsilon_0)} \ar@{^{(}->}[r] \ar[d] &
\overline{\mathcal M}^{para}_{[0,\varepsilon_0)} \ar[r]^(.25){Glue^{-1}}\ar[d] & [0,\varepsilon_0]
\times {\mathcal M}^{dfd}(K_-,J_-;f;K_+,J_+;A_-\# A_+) \ar[dl]\\ (0,\varepsilon_0)
\ar@{^{(}->}[r] & [0,\varepsilon_0) &{} }.
\end{equation*}
\end{thm}
The statement of this theorem is the analog to the corresponding theorem,
Theorem 10.19 \cite{oh-zhu} for the case $\ell =0$. However the gluing
analysis and its off-shell framework are of nature very different from those
of \cite{oh-zhu}: In \cite{oh-zhu}, we need to \emph{rescale the target}
near the nodal point of nodal Floer trajectories, while in the present paper
we do not need to rescale the target but we need to \emph{renormalize} the
domain variable $(\tau ,t)$ to $(\tau /\varepsilon ,t/\varepsilon )$ to
obtain the limiting gradient flow of $f$ arising from the thin part of the
degeneration of ${\mathcal{M}}^{\varepsilon }$ as $\varepsilon \rightarrow 0$
as done in \cite{FO}. The length $\ell $ of the limiting gradient trajectory
is determined by the limit $\lim_{\varepsilon \rightarrow 0}\varepsilon
R(\varepsilon )=\ell $, i.e, by how fast the length of thin part of
degenerating Floer trajectories increases relative to the speed as $%
\varepsilon \rightarrow 0$. We do not need to rescale the target $M$ unlike
the case in \cite{oh-zhu}. This is because the presence of gradient line
with \emph{non-zero} length carries enough information to recover all the
solutions in ${\mathcal{M}}^{\varepsilon }$ \emph{with immersed join points}
nearby the elements from ${\mathcal{M}}(K_{-}|f|K_{+};A_{-}\#A_{+})$ in
Gromov-Hausdorff topology, as long as $\varepsilon $ is sufficiently small.
On the other hand, as $l=0$ the Sobolev constant in present paper blows up
which obstructs application of the same scheme which led us to blowing up
target appears to be necessary as in \cite{oh-zhu}. \emph{In both cases, we
need to assume that the Floer trajectories either at the nodal points or at
the join point of the gradient trajectories are immersed.} It appears that
without this assumption gluing analysis is much more delicate. The immersion
condition can be achieved for a generic Floer datum under the small index
conditions (Proposition \ref{immersed-joint}) which naturally enter in the
construction of various operations in Floer theory.
Incidentally we would like to point out that the main theorem can be also
used to give another proof of isomorphism property of the PSS map which uses
a piecewise smooth cobordism different from that of \cite{oh-zhu}. Namely we
directly go from \textquotedblleft disk-flow-disk\textquotedblright\
configurations to resolved Floer trajectories by the above mentioned
adiabatic gluing theorem \emph{without passing through the nodal Floer
trajectories} unlike the one proposed in \cite{PSS} and realized in \cite%
{oh-zhu}. Unlike the case of \cite{oh-zhu}, we do not need to change the
target this time. We also remark that the full gradient trajectory (i.e. $%
l=\infty $) case can be treated by the same method of this paper, which is
in fact easier, because we do not need to put the power weight $\rho \left(
\tau \right) $ on $\chi $ (see Remark \ref{weights}).
Some remarks on the relationship with other relevant literature are now in
order.
We would like to point out that our gluing analysis in the present paper and
in \cite{oh-zhu} can be applied to many other contexts where similar
thick-thin decomposition occurs (including $l=0$ case) during the adiabatic
degeneration arises. One of the very first instances where importance of
this kind of gluing analysis was mentioned is the papers \cite{fukaya}, \cite%
{oh:newton}. An analogous analysis was also carried out by Fukaya and the
senior author in \cite{FO} when there is no non-constant bubbles around.
The gluing analysis with non-constant bubbles around for the equation
\begin{equation*}
\frac{\partial u}{\partial \tau} + J_0 \left(\frac{\partial u}{\partial t} -
X_{\varepsilon f}(u)\right) = 0
\end{equation*}
as $\varepsilon > 0$ small was pursued by Fukaya and the senior author at
that time but left unfinished \cite{fukaya-oh2} since then, partly because
the analysis turned out to be much more nontrivial than they originally
expected. The gluing analysis given in the present paper can be similarly
applied also to this case \emph{under the assumption that the join points
are immersed}. Section \ref{sec:variants} of the present paper briefly
outlines the necessary modification for this case in the setting of both
Hamiltonian and Lagrangian contexts. One outcome of this analysis and the
dimension counting argument proves the following equivalence theorem which
was expected in \cite{oh:newton}. We state the theorem only for the case
where we do not need any additional argument except some immediate
translations of our gluing analysis presented in this paper without using
the virtual machinery.
\begin{thm}
Let $(M,\omega )$ be a monotone symplectic manifold and $L\subset (M,\omega
) $ be a monotone Lagrangian submanifold. Fix a Darboux neighborhood $U$ of $%
L$ and identify $U $ with a neighborhood of the zero section in $T^{\ast }L$%
. Consider $k+1$ Hamiltonian deformations of $L$ by autonomous Hamiltonian
functions $F_{0},\ldots ,F_{k}:M\rightarrow {\mathbb{R}}$ such that
\begin{equation*}
F_{i}=\chi f_{i}\circ \pi
\end{equation*}%
where $f_0,f_1,\ldots f_k: L\rightarrow {\mathbb{R}}$ are generic Morse
functions, $\chi $ is a cut-off function such that $\chi=1$ on $U$ and
supported nearby U. Now consider
\begin{equation*}
L_{i,\varepsilon }=\operatorname{Graph}(\varepsilon df_{i})\subset U\subset M,\quad
i=0,\ldots ,k.
\end{equation*}%
Assume transversality of $L_{i}$'s of the type given in \cite{FO}. Consider
the intersections $p_{i}\in L_{i}\cap L_{i+1}$ such that
\begin{equation*}
\mathop{\kern0pt{\rm dim}}\nolimits{\mathcal{M}}(L_{0},\ldots
,L_{k};p_{0},\ldots ,p_{k})=0,\,1.
\end{equation*}%
Then when $\varepsilon $ is sufficiently small, the moduli space ${\mathcal{M%
}}(L_{0,\varepsilon },\ldots ,L_{k,\varepsilon };p_{0},\ldots ,p_{k})$ is
diffeomorphic to the moduli space of pearl complex defined in \cite%
{oh:newton,oh:imrn}, \cite{biran-cornea}.
\end{thm}
The following is an immediate corollary of this theorem.
\begin{cor}
Under the same hypothesis on $(M,\omega )$ and $L$, the $A_\infty$-structure
proposed in \cite{fukaya,oh:newton,oh:imrn}, which also coincides with the
quantum structure defined in \cite{biran-cornea}, is isomorphic to the $%
A_{\infty }$-structure defined in \cite{FOOO}.
\end{cor}
In \cite{BO}, Bourgeois and Oancea studied the Floer equation of autonomous
Hamiltonian $H$ under the assumption that its 1-periodic orbits are
transversally nondegenerate on a symplectic manifold with contact type
boundary, in relation to the study of the linearized contact homology of a
fillable contact manifold and symplectic homology of its filling. Their
construction is based on Morse-Bott techniques for Floer trajectories where
a function of the type $\varepsilon f$ is used on the Morse-Bott set of
\emph{parameterized }periodic orbits of $H$ to break the $S^{1}$ symmetry of
the orbit space, whose loci consist of \emph{isolated}, \emph{unparameterized%
}, \emph{nonconstant} periodic orbits. The Hamiltonian pearl complex case
treated in subsection \ref{subsec:ham-pearl} can be put in a similar
Morse-Bott setting this time to break the $M$-symmetry of constant orbits of
Hamiltonian $0$ by adding a small Morse function $\varepsilon f$ on $M$
itself.
\textbf{Acknowledgment:} The first named author would like to thank Kenji
Fukaya for having much discussion on this kind of gluing analysis some years
ago after \cite{FO} was completed when they attempted to push the gluing
analysis further to include the thick part but stopped working on it \cite%
{fukaya-oh2}.
The second named author would like to thank the math department of Chinese
University of Hong Kong for its nice research environment, where major part
of his research on this paper was carried out. He also thanks Xiaowei Wang
for constant encouragement.
\section{Invariant set-up of the Floer equation}
\label{sec:invariant}
In this section, we formulate the set-up for the general Floer's perturbed
Cauchy-Riemann equation on compact Riemann surface with a finite number of
punctures. This requires a coordinate-free framework of the equation.
\subsection{Punctures with analytic coordinates}
\label{subsec:punctures}
We start with the description of positive and negative \emph{punctures}. Let
$\Sigma$ be a compact Riemann surface with a marked point $p \in \Sigma$.
Consider the corresponding punctured Riemann surface $\dot \Sigma$ with an
analytic coordinates $z: D\setminus \{p\} \to {\mathbb{C}}$ on a
neighborhood $D \setminus \{p\} \subset \dot \Sigma$. By composing $z$ with
a linear translation of ${\mathbb{C}}$, we may assume $z(p) = 0$.
We know that $D \setminus \{p\}$ is conformally isomorphic to both $%
[0,\infty) \times S^1$ and $(-\infty,0] \times S^1$.
\begin{enumerate}
\item We say that the pair $(p;(D,z))$ has a \emph{incoming cylindrical end}
(with analytic chart) if we have
\begin{equation*}
D = z^{-1}(D^2(1))
\end{equation*}
and are given by the biholomorphism
\begin{equation*}
(\tau, t) \in S^1 \times (-\infty,0] \mapsto e^{2\pi(\tau + it)} \in D^2(1)
\setminus \{0\} \mapsto z^{-1} \in D\setminus \{p\}.
\end{equation*}
We call the corresponding puncture $p \in \Sigma$ a \emph{positive puncture}.
\item We say that the pair $(p;(D,z))$ has a \emph{outgoing cylindrical end}
(with analytic chart) if we have
\begin{equation*}
D = z^{-1}(D^2(1))
\end{equation*}
and are given by the biholomorphism
\begin{equation*}
(\tau,t) \in S^1 \times [0,\infty) \mapsto e^{-2\pi(\tau + it)} \in D^2(1)
\setminus \{0\} \mapsto z^{-1} \in D \setminus \{p\}.
\end{equation*}
In this case, we call the corresponding puncture $(p;(D,z))$ a \emph{%
negative puncture} (with analytic chart).
\end{enumerate}
\subsection{Hamiltonian perturbations}
\label{eq:perturb}
Now we describe the Hamiltonian perturbations in a coordinate free fashion.
Let $\Sigma$ be a compact Riemann surface and $\dot \Sigma$ denote $\Sigma$
with a finite number of punctures and analytic coordinates. We denote by ${%
\mathcal{J}}_{0,\omega}$ the set of almost complex structures that are
cylindrical near the puncture with respect to the given analytic charts $z =
e^{\pm(2\pi(\tau + it)}$. Define ${\mathcal{J}}_\Sigma$ or ${\mathcal{J}}%
_{\dot\Sigma}$ to be the set of maps $J: \Sigma, \, \dot \Sigma \to {%
\mathcal{J}}_{0,\omega}$ respectively.
\begin{defn}
We call $K \in \Omega^1(\Sigma,C^\infty(M))$ \emph{cylindrical} at the
puncture $p \in \Sigma$ with analytic chart $(D,z)$, if it has the form
\begin{equation*}
K(\tau,t) = H(t)\, dt
\end{equation*}
in $D \setminus \{p\}$. We denote by ${\mathcal{K}}_{\dot\Sigma}$ the set of
such $K$'s.
\end{defn}
One important quantity associated to the one-form $K$ is a two-form, denoted
by $R_K$, and defined by
\begin{equation} \label{eq:R_K}
R_K\left(\xi_1,\xi_2\right) = \xi_1[K(\xi_2)] - \xi_2[K(\xi_1)] -
\left\{K(\xi_2), K(\xi_1)\right\}
\end{equation}
for two vector fields $\xi_1, \, \xi_2$, where $\xi_1[(K(\xi_2)]$ denotes
directional derivative of the function $K(\xi_2)(z,x)$ with respect to the
vector field $\xi_1$ as a function on $\Sigma$, holding the variable $x \in
M $ fixed. It follows from the expression that $R_K$ is tensorial on $\Sigma$%
.
The Hamiltonian-perturbed Cauchy-Riemann equation has the form
\begin{equation} \label{eq:KJ}
(du + P_K(u))^{(0,1)}_J = 0 \quad\mbox{ or equivalently }\, {\overline
\partial}_J(u) + (P_K)_J^{(0,1)}(u) = 0
\end{equation}
on $\Sigma$ in general.
For each given such pair $(K,J)$, it defines a perturbed Cauchy-Riemann
operator by
\begin{equation*}
{\overline \partial}_{(K,J)} u := {\overline \partial}_Ju + P_K(u)^{(0,1)}_J
= (du +P_K(u))^{(0,1)}_J.
\end{equation*}
Let $(\mathfrak{p, q)}$ be a given set of positive punctures $\mathfrak{p =
\{p_1, \cdots, p_k\}}$ and with negative punctures $\mathfrak{q = \{ q_1,
\cdots, q_\ell\}}$ on $\Sigma$. For each given Floer datum $(K,J)$ and a
collection $\vec z = \{z_*\}_{* \in \mathfrak{p \cup q}}$ of asymptotic
periodic orbits $z_*$ attached to the punctures $* = p_i$ or $* = q_j$, we
consider the perturbed Cauchy-Riemann equation
\begin{equation} \label{eq:KJ-asymp}
\begin{cases}
{\overline \partial}_{(K,J)}(u) = 0 \\
u(\infty_*,t) = z_*(t).%
\end{cases}%
\end{equation}
Our main interest will lie in the case where $(|\mathfrak{p}|, |\mathfrak{q}%
|)$ is either $(1,0), \, (0,1)$ or $(1,1)$ in the present paper.
One more ingredient we need to give the definition of the
Hamiltonian-perturbed moduli space is the choice of an appropriate energy of
the map $u$. For this purpose, we fix a metric $h_\Sigma$ which is
compatible with the structure of the Riemann surface and which has the
cylindrical ends with respect to the given cylindrical coordinates near the
punctures, i.e., $h_\Sigma$ has the form
\begin{equation} \label{eq:gSigma}
h_\Sigma = d\tau^2 + dt^2
\end{equation}
on $D_* \setminus \{*\}$. We denote by $dA_\Sigma$ the corresponding area
element on $\Sigma$.
Here is the relevant energy function
\begin{defn}[Energy]
For a given asymptotically cylindrical pair $(K,J)$, we define
\begin{equation*}
E_{(K,J)}(u) = \frac{1}{2}\int_\Sigma |du - P_K(u)|_J^2\, dA_\Sigma
\end{equation*}
where $|\cdot|_J$ is the norm of $\Lambda^{(0,1)}(u^*TM) \to \Sigma$ induced
by the metrics $h_\Sigma$ and $g_J: = \omega(\cdot, J \cdot)$.
\end{defn}
Note that this energy depends only on the conformal class of $h_\Sigma$,
i.e., depends only on the complex structure $j$ of $\Sigma$ and restricts to
the standard energy for the usual Floer trajectory moduli space given by
\begin{equation*}
E_{(H,J)} = \frac{1}{2}\int_{C_*} \left(\left|{\frac{\partial u}{\partial
\tau}}\right|_J^2 + \left|{\frac{\partial u}{\partial t}} -
X_H(u)\right|_J^2\right) \, dt\, d\tau
\end{equation*}
in the cylindrical coordinates $(\tau,t)$ on the cylinder $C_*$
corresponding to the puncture $*$. $E_{(K,J)}(u)$ can be bounded by a more
topological quantity depending only on the asymptotic orbits, or more
precisely their liftings to the \emph{universal covering space} of ${%
\mathcal{L}}_0(M)$, where the latter is the contractible loop space of $M$.
As usual, we denote such a lifting of a periodic orbit $z$ by $[z,w]$ where $%
w:D^2 \to M$ is a disc bounding the loop $z$.
We also consider the \emph{real blow-up} of $\dot \Sigma \subset \Sigma$ at
the punctures and denote it by $\overline\Sigma$ which is a compact Riemann
surface with boundary
\begin{equation*}
\partial \overline \Sigma = \coprod_{* \in \mathfrak{p} \cup \mathfrak{q}}
S^1_*
\end{equation*}
where $S^1_*$ is the exceptional circle over the point $*$. We note that
since there is given a preferred coordinates near the point $*$, each circle
$S^1_*$ has the canonical identification
\begin{equation*}
\theta_*: S^1_* \to {\mathbb{R}}/{\mathbb{Z}} = [0,1] \mod 1.
\end{equation*}
We note that for a given asymptotic orbits $\vec z$, one can define the
space of maps $u : \dot \Sigma \to M $ which can be extended to $\overline
\Sigma$ such that $u \circ \theta_* = z_*(t)$ for $* \in \mathfrak{p \cup q}$%
. Each such map defines a natural homotopy class $B$ relative to the
boundary. We denote the corresponding set of homotopy classes by $\pi_2(\vec
z)$. When we are given the additional data of bounding discs $w_*$ for each $%
z_*$, then we can form a natural homology (in fact a homotopy class),
denoted by $B \# \left(\coprod_{* \in \mathfrak{p} \cup \mathfrak{q}}
[w_*]\right) \in H_2(M)$, by `capping-off' the boundary components of $B$
using the discs $w_*$ respectively.
\begin{defn}
Let $\{[z_*,w_*]\}_{* \in \mathfrak{p \cup q}}$ be given. We say $B \in
\pi(\vec z)$ is \emph{admissible} if it satisfies
\begin{equation} \label{eq:Bsharpws}
B \# \left(\coprod_{* \in \mathfrak{p} \cup \mathfrak{q}} [w_*]\right) = 0
\quad \mbox{in }\, H_2(M,{\mathbb{Z}})
\end{equation}
\end{defn}
where
\begin{equation*}
\# : \pi_2(\vec z) \times \prod_{* \in \mathfrak{p} \cup \mathfrak{q}}
\pi_2(z_*) \to H_2(M,{\mathbb{Z}})
\end{equation*}
is the natural gluing operation of the homotopy class from $\pi_2(\vec z)$
and those from $\pi_2(z_*)$ for $* \in \mathfrak{p} \cup \mathfrak{q}$. Now
we are ready to give the definition of the Floer moduli spaces.
\begin{defn}
Let $(K,J)$ be a Floer datum over $\Sigma$ with punctures $\mathfrak{p, \, q}
$, and let $\{[z_*,w_*]\}_{* \in \mathfrak{p\cup q}}$ be the given
asymptotic orbits. Let $B \in \pi_2(\vec z)$ be a homotopy class admissible
to $\{[z_*,w_*]\}_{* \in \mathfrak{p\cup q}}$. We define the moduli space
\begin{equation} \label{eq:Mz*w*}
{\mathcal{M}}(K,J;\{[z_*,w_*]\}_*) = \{u: \dot \Sigma \to M \mid u \,
\mbox{ satisfies (\ref{eq:KJ-asymp}) and
$[u]\# (\coprod_{* \in \mathfrak{p} \cup \mathfrak{q}} [w_*]) =0$ }\}.
\end{equation}
\end{defn}
We note that the moduli space ${\mathcal{M}}(K,J;\{[z_*,w_*]\}_*)$ is a
finite union of the moduli spaces
\begin{equation*}
{\mathcal{M}}(K,J;\vec z; B); \quad B \# \left(\coprod_{* \in \mathfrak{p}
\cup \mathfrak{q}} [w_*]\right) = 0.
\end{equation*}
\section{Adiabatic family of Floer moduli spaces}
\label{sec:floer}
In this section, we give the precise description of one-parameter family of
perturbed Cauchy-Riemann equations in a coordinate-free form parameterized
by $\varepsilon > 0$ such that as $\varepsilon \to 0$ the equation becomes
degenerate in a suitable sense which we will make precise.
We first consider the cases $\mathfrak{p}_\pm = \{e_\pm\},\, \mathfrak{q}%
_\pm = \{o_\pm\}$ with one negative and one positive punctures. Then we
consider the one-form $K_\pm$ such that
\begin{equation} \label{eq:Kpm}
K_\pm(\tau,t,x) = \kappa^\pm(\tau) H_t(x)
\end{equation}
on $U_\pm$ in terms of the given analytic coordinates.
Now we consider a one-parameter family $(K_\varepsilon,J_\varepsilon)$ with $%
R = R(\varepsilon) \to \infty$,
\begin{equation} \label{eq:length}
\varepsilon R(\varepsilon) \to \ell
\end{equation}
with $\ell \geq 0$ as $\varepsilon \to 0$. More precise description of $%
(K_\varepsilon, J_\varepsilon)$ is in order.
To define this family, we fix any two continuous functions $R(\varepsilon )$
and $S(\varepsilon )$ that satisfies
\begin{equation}
\varepsilon R(\varepsilon )\rightarrow \ell ,\,\varepsilon S(\varepsilon
)\rightarrow 0 \label{eq:RSe}
\end{equation}%
as $\varepsilon \rightarrow 0$. For example, we can choose
\begin{equation}
R(\varepsilon )=\frac{l}{\varepsilon },\quad S(\varepsilon )=\frac{1}{2\pi }%
\ln (1+\frac{l}{\varepsilon }) \label{eq:ReSe}
\end{equation}%
and introduce
\begin{equation*}
\tau \left( \varepsilon \right) =R\left( \varepsilon \right) +\frac{p-1}{%
\delta }S\left( \varepsilon \right)
\end{equation*}%
for the convenience of our exposition later, which will appear frequently in
our calculations. Here we fix any $p>2$ and $0<\delta <1$. The $p>2$ is for $%
W^{1,p}\hookrightarrow C^{0}$ Sobolev embedding on Riemann surface, and $%
0<\delta <1$ is to get rid of the $0$-spectrum of $i\frac{\partial }{%
\partial t}$ on $S^{1}$. Any such choice of $p$ and $\delta $ will be
suffice for our analysis. The choice of $\tau \left( \varepsilon \right) $
is not canonical as it depends on $p$ and $\delta $, but when $p\rightarrow
2 $ and $\delta \rightarrow 1$, $\left\vert \tau \left( \varepsilon \right)
-R\left( \varepsilon \right) \right\vert $ is close to $S\left( \varepsilon
\right) $.
Then we decompose ${\mathbb{R}}$ into
\begin{equation*}
-\infty <-\tau \left( \varepsilon \right) -1<-\tau \left( \varepsilon
\right) <-R(\varepsilon )<R(\varepsilon )<\tau \left( \varepsilon \right)
<\tau \left( \varepsilon \right) +1<\infty .
\end{equation*}
Let $\dot\Sigma_\pm$ be two compact surfaces each with two positive puncture
(resp. one negative puncture) with analytic coordinates. Let $e_\pm, o_\pm
\in \dot \Sigma_\pm$ two marked points and denote by $(\tau,t)$ with $\pm
\tau \geq 0$ the cylindrical charts of $\dot\Sigma_\pm \setminus \{o_\pm\}$
such that $z = e^{\pm 2\pi(\tau + it)}$.
We choose neighborhoods $U_{\pm }$ of $e_{\pm }$ and analytic charts $%
\varphi _{\pm }:U_{\pm }\rightarrow {\mathbb{C}}$ and the associated
coordinates $z=e^{2\pi (\tau +it)}$ so that
\begin{equation*}
\varphi _{+}(U_{+}\setminus \{e_{+}\})\cong \lbrack 0,\infty )\times
S^{1},\,\quad \varphi _{-}(U_{-}\setminus \{e_{-}\})\cong (-\infty ,0]\times
S^{1},
\end{equation*}%
we fix a function
\begin{equation}
\kappa ^{+}(\tau )=%
\begin{cases}
0 \quad & 0 \leq \tau \leq 1 \\
1\quad & {} \tau \geq 2%
\end{cases}
\label{eq:beta}
\end{equation}%
and let $\kappa ^{-}(\tau )=\kappa ^{+}(-\tau )$.
We define a glued surface with a cylindrical coordinates, denoted by
\begin{equation*}
\dot\Sigma: = \Sigma_+ \# \Sigma_-
\end{equation*}
Then $\dot \Sigma$ carries two natural marked points $(e_-,e_+)$. We pick an
embedded path passing a point $o \in \Sigma_+ \cap \Sigma_-$ and connecting $%
e_-, e_+$. We fix the unique cylindrical coordinates $\dot\Sigma \cong {%
\mathbb{R}} \times S^1$ mapping the path to ${\mathbb{R}} \times \{0\}$ and $%
o$ to $(0,0)$.
With this cylindrical coordinates, we consider $\varepsilon >0$ and a family
of cut-off functions defined by $\kappa _{\varepsilon }^{+}(\tau )=\kappa
^{+}(\tau -\tau (\varepsilon )+1)$ and $\kappa _{\varepsilon }^{-}(\tau
)=\kappa _{\varepsilon }^{+}(-\tau )$. It is easy to see
\begin{equation}
\kappa _{\varepsilon }^{+}(\tau )=%
\begin{cases}
1\,\quad \, & \mbox{for }\,\tau \geq \tau (\varepsilon )+1 \\
0\quad \, & \mbox{for }\,0\leq \tau \leq \tau (\varepsilon )%
\end{cases}%
,\quad \kappa _{\varepsilon }^{-}=%
\begin{cases}
1\,\quad & \mbox{for }\,\tau \leq -\tau (\varepsilon )-1 \\
0\quad & \mbox{for }\,\tau (\varepsilon )\leq \tau \leq 0%
\end{cases}
\label{eq:betaR}
\end{equation}%
We then define $(K_{\varepsilon },J_{\varepsilon })$ to be the glued family
\begin{eqnarray}
K_{\varepsilon }(\tau ,t) &=&%
\begin{cases}
\kappa _{\varepsilon }^{+}(\tau )\cdot H_{t}\quad & \tau \geq R(\varepsilon )
\\
\kappa _{\varepsilon }^{0}(\tau )\cdot \varepsilon f\quad & |\tau |\leq
R(\varepsilon ) \\
\kappa _{\varepsilon }^{-}(\tau )\cdot H_{t}\quad & \tau \leq -R(\varepsilon
).%
\end{cases}
\label{eq:Kepsilon} \\
J_{\varepsilon }^{\pm }(\tau ,t,x) &=&%
\begin{cases}
J^{\kappa _{\varepsilon }^{+}(\tau )}(t,x)\quad & \tau \geq R(\varepsilon )
\\
J_{0}(x)\quad & |\tau |\leq R(\varepsilon ) \\
J^{\kappa _{\varepsilon }^{-}(\tau )}(t,x)\quad & \tau \leq -R(\varepsilon )%
\end{cases}
\label{eq:Jepsilon}
\end{eqnarray}%
associated to $\kappa _{\varepsilon }^{\pm }$ respectively, where $\kappa
_{\varepsilon }^{0}(\tau )$ is a smooth cut-off function such that
\begin{equation}
\kappa _{\varepsilon }^{0}(\tau )=%
\begin{cases}
1\quad &
\mbox{for $\left\vert \tau \right\vert \leq R\left( \varepsilon
\right) $} \\
0\quad &
\mbox{for $\left\vert \tau \right\vert \geq R\left(
\varepsilon \right) +1$},%
\end{cases}
\label{eq:beta0}
\end{equation}%
$\left\vert \kappa _{\varepsilon }^{0}(\tau )\right\vert \leq 1,\left\vert
(\kappa _{\varepsilon }^{0})^{\prime }(\tau )\right\vert \leq 2$.
We will vary $R=R(\varepsilon)$ depending on $\varepsilon$ and study the
family of equation
\begin{equation} \label{eq:duPK}
(du + P_{K_\varepsilon}(u))^{(0,1)}_J = 0
\end{equation}
as $\varepsilon \to 0$.
\section{Adiabatic convergence}
\label{sec:adiabatic}
By definition of $K_{\varepsilon }$ and $J_{\varepsilon }$, as $\varepsilon
\rightarrow 0$, on the domain
\begin{equation*}
\lbrack -R(\varepsilon ),R(\varepsilon )]\times S^{1}
\end{equation*}%
we have $K_{\varepsilon }(\tau ,t)\equiv \varepsilon f$ and $J_{R}(\tau
,t)\equiv J_{0}$, and so \eqref{eq:KJe} becomes
\begin{equation*}
\frac{\partial u}{\partial \tau }+J_{0}\left( \frac{\partial u}{\partial t}%
-\varepsilon X_{f}(u)\right) =0.
\end{equation*}%
Furthermore $K_{\varepsilon }^{\pm }(\tau ,t)\equiv H_{t}\,dt$, $%
J_{\varepsilon }^{\pm }(\tau ,t)\equiv J_{t}$ on
\begin{equation*}
{\mathbb{R}}\times S^{1}\setminus \lbrack -\tau (\varepsilon )-1,\tau
(\varepsilon )+1]\times S^{1}
\end{equation*}%
(\ref{eq:duPK}) is cylindrical at infinity, i.e., invariant under the
translation in $\tau $-direction at infinity.
Note that on any fixed compact set $B \subset {\mathbb{R}} \times S^1$, we
will have
\begin{equation*}
B \subset [-R(\varepsilon),R(\varepsilon)]\times S^1
\end{equation*}
for all sufficiently small $\varepsilon$. And as $\varepsilon \to 0$, $%
K_\varepsilon \to 0$ on $B$ in $C^\infty$-topology, and hence the equation (%
\ref{eq:duPK}) converges to ${\overline \partial}_{J_0} u = 0$ on $B$ in
that $J \to J_0$ and $K_\varepsilon \to 0$ in $C^\infty$-topology. On the
other hand, after translating the region $(-\infty, -(R(\varepsilon) - \frac{%
1}{3}]$ to the right (resp. $[R(\varepsilon) -\frac{1}{3},\infty)$ to the
left) by $2R(\varepsilon) - \frac{2}{3}$ in $\tau$-direction, (\ref{eq:duPK}%
) converges to
\begin{equation*}
\frac{\partial u}{\partial \tau} + J^+_\varepsilon \left(\frac{\partial u}{%
\partial t} - X_H(u) \right) = 0
\end{equation*}
on $(-\infty, 0] \times S^1$ (resp. on $[0, \infty) \times S^1$) and ${%
\overline \partial}_{J_0}u = 0$ on $[0,R(\varepsilon) - \frac{1}{3}] \times
S^1$ (resp. on $[-R(\varepsilon) + \frac{1}{3},0] \times S^1$).
Now we are ready to state the meaning of the \emph{adiabatic convergence}
for a sequence $u_{n}$ of solutions $(du_{n}+P_{K_{\varepsilon
_{n}}})_{J_{\varepsilon _{n}}}^{(0,1)}=0$ as $n\rightarrow \infty $. After
taking away bubbles, we can assume that we have the derivative bound
\begin{equation}
|du_{n}|<C<\infty \label{eq:|du|<C}
\end{equation}%
where we take the norm $|du_{n}|$ with respect to the given metric $g$ on $M$%
.
\emph{From now on, we will always assume that we have this derivative bound,
i.e., that bubble does not occur as $\varepsilon \to 0$.}
We denote
\begin{equation*}
\Theta_\varepsilon = \left[-R(\varepsilon), R(\varepsilon)\right] \times S^1
\end{equation*}
and consider the local energy
\begin{defn}
We define the \emph{local energy} of $u$ on $B$ by
\begin{equation*}
E_{J,B}(u) := \int_B |du|_J^2 dt \, d\tau.
\end{equation*}
\end{defn}
There are two cases to consider :
\begin{enumerate}
\item there exists a subsequence $n_{i}$ and $c>0$ such that $E_{J,\Theta
_{\varepsilon _{n_{i}}}}(u_{n_{i}})>c>0$ for all sufficiently large $i$,
\item $\limsup_{n \to \infty} E_{J,\Theta_{\varepsilon_n}}(u_n) = 0$.
\end{enumerate}
For the case (1), standard argument produces a non-constant $J_0$%
-holomorphic sphere. To describe the situation (2) in a precise manner, we
introduce the following notion of adiabatic convergence to a disc-flow-disc
trajectory $(u_-,\chi,u_+)$. This would describe a neighborhood basis of $%
(u_-,\chi,u_+)$ in a suitable completion of ${\mathcal{F}}%
^{1,p}(K_{\varepsilon_n},J_{\varepsilon_n};z_-,z_+;B)$.
We first recall the definition of Hausdorff distance of subsets $A, \, B
\subset X$
\begin{equation*}
d_{\RIfM@\expandafter\text@\else\expandafter\mbox\fi{\textrm{H}}}(A,B) = \sup_{x \in A}d(x,B) + \sup_{y \in B} d(A,y).
\end{equation*}
\subsection{Adiabatic deformations of domain}
\label{subsec:domain}
To provide a rigorous definition of the adiabatic convergence, one has to
give a precise way of degenerating the punctured sphere equipped with the
cylindrical structure at the punctures to the union of two spheres joined by
a line segment of length $2\ell $ parameterized by the interval $[-\ell
,\ell ]$.
For $U_{\pm }\subset \Sigma _{\pm }$, in terms of analytic charts $\varphi
_{\pm }:U_{\pm }\rightarrow D^{2}\subset {\mathbb{C}}$ and the associated
coordinates $z$, we identify $U_{\pm }\setminus \{o_{\pm }\}$ with the open
punctured unit disc $D^{2}\setminus \{0\}$. We first consider the annular
domain of $U_{\pm }\setminus \{o_{\pm }\}$:
\begin{equation*}
\RIfM@\expandafter\text@\else\expandafter\mbox\fi{\textrm{Ann}}_{\alpha }:=\{z\in D^{2}\mid |\alpha |^{3/4}\leq |z|\leq
|\alpha |^{1/4}\}=\{z\in D^{2}\mid |R_{\alpha }|^{-3/2}\leq |z|\leq
|R_{\alpha }|^{-1/2}\}.
\end{equation*}%
The choice of the power is dictated by that of Fukaya-Ono's deformation
given in \cite{fukaya-ono} in which $|\alpha |=R_{\alpha }^{-2}$ for $%
|\alpha |$ sufficiently small.
We note that the conformal modulus of $\RIfM@\expandafter\text@\else\expandafter\mbox\fi{\textrm{Ann}}_{\alpha }$ is $%
\Vert \alpha \Vert ^{-1/2}=R_{\alpha }$. For each given $\varepsilon >0$, we
choose $\alpha (\varepsilon )$ so that
\begin{equation*}
\frac{p-1}{\delta }S(\varepsilon )=\ln \Vert \alpha (\varepsilon )\Vert
^{-1/2}.
\end{equation*}%
We recall the choice of $S(\varepsilon )$
\begin{equation*}
S(\varepsilon )=\frac{1}{2\pi }\ln \left( 1+\frac{\ell }{\varepsilon }%
\right) .
\end{equation*}%
Then we choose a biholomorphism
\begin{equation*}
\varphi _{\varepsilon }^{-}:\RIfM@\expandafter\text@\else\expandafter\mbox\fi{\textrm{Ann}}_{\alpha (\varepsilon
)}\rightarrow \lbrack -\tau (\varepsilon ),-R(\varepsilon )]\times S^{1}.
\end{equation*}%
Similarly we define
\begin{equation*}
\varphi _{\varepsilon }^{+}:\RIfM@\expandafter\text@\else\expandafter\mbox\fi{\textrm{Ann}}_{\alpha (\varepsilon
)}\rightarrow \lbrack R(\varepsilon ),\tau (\varepsilon )]\times S^{1}.
\end{equation*}%
Using this, we define a family of glued surfaces
\begin{equation*}
\Sigma _{\varepsilon }=\left( \Sigma _{-}\setminus D^{2}(|\alpha
(\varepsilon )|^{3/4})\right) \cup _{\varphi _{\varepsilon }^{-}}\left(
[-\tau (\varepsilon ),\tau (\varepsilon )]\times S^{1}\right) {}_{(\varphi
_{\varepsilon }^{+})^{-1}}\cup \left( \Sigma _{+}\setminus D^{2}(|\alpha
(\varepsilon )|^{3/4})\right) .
\end{equation*}%
We denote
\begin{equation*}
C(\varepsilon )=[-\tau (\varepsilon ),\tau (\varepsilon )]\times S^{1}
\end{equation*}%
with the standard metric $g_{C(\varepsilon )}$. Then through the
identifications $\varphi _{\pm }:U_{\pm }\rightarrow D^{2}$, this family
give rise to the following $\varepsilon $-parameterized family of resolved
cylinders $(\Sigma _{\varepsilon }^{adi},g_{\varepsilon }^{adi})$ equipped
with the metric provided by
\begin{equation*}
g_{\varepsilon }^{adi}=%
\begin{cases}
g_{+}\quad & \mbox{on }\,\Sigma _{+}\setminus U_{+}(|\alpha (\varepsilon )|)
\\
g_{C(\varepsilon )}\quad & \mbox{on }\,C(\varepsilon ) \\
g_{-}\quad & \mbox{on }\,\Sigma _{-}\setminus U_{-}(|\alpha (\varepsilon )|)%
\end{cases}%
\end{equation*}%
and suitably interpolated in between. Note that the conformal structure of $%
\Sigma _{\varepsilon }$ is degenerating on the annuli region $C(\varepsilon
) $ when $\varepsilon \rightarrow 0$, and any given compact subset $K\subset
\Sigma _{\pm }\setminus \{o_{\pm }\}$ is covered by $\Sigma _{\pm }\setminus
U_{\pm }(\delta )$ for a sufficiently small $\delta >0$.
\subsection{Definition of adiabatic convergence}
Now we involve maps defined on the resolved domains $(\Sigma_%
\varepsilon^{adi},g_\varepsilon^{adi})$ and provide a precise definition of
adiabatic convergence or of adiabatic topology near the disc-flow-disc
moduli space
\begin{equation*}
{\mathcal{M}}^{dfd}_{\ell}(z_-;f;z_+;B)
\end{equation*}
inside the off-shell spaces
\begin{equation*}
\overline {\mathcal{F}}^{para}(z_-;f;z_+;B) = \bigcup_{\ell \geq \ell_0}{%
\mathcal{F}}^\ell(z_-;f;z_+,;B).
\end{equation*}
\begin{defn}
\label{defn:adiabatic} Let $\{\varepsilon_j\}$ be sequence with $%
\varepsilon_j \to 0$ as $j \to \infty$.
We say a sequence $u_j$ of maps in ${\mathcal{F}}^{1,p}(K_{%
\varepsilon_j},J_{\varepsilon_j};z_-,z_+;B)$ \emph{$\{\varepsilon_j\}$%
-adiabatically converges} to a disc-flow-disc trajectory $(u_-,\chi,u_+)$ if
the following hold:
\begin{enumerate}
\item $\lim_{j\rightarrow \infty }E_{J,\Theta _{\varepsilon _{j}}}(u_{j})=0.$
\item $\lim_{n\rightarrow \infty }d_{\RIfM@\expandafter\text@\else\expandafter\mbox\fi{\textrm{H}}}(u_{j}([-R(%
\varepsilon _{j}),R(\varepsilon _{j})]\times S^{1}),\chi ([-l,l])=0$, where $%
d_{\RIfM@\expandafter\text@\else\expandafter\mbox\fi{\textrm{H}}}$ is the Gromov-Hausdorff metric.
\item $u_{j}|_{\Sigma _{\pm }\setminus U_{\pm }(\zeta )}\rightarrow u_{\pm }$
in $C^{\infty }$ for any given $0<\zeta <1$, or equivalently,
$u_{j}(\cdot \pm \tau \left( \varepsilon _{j}\right) ,\cdot )\rightarrow
u_{\pm }$ in $C^{\infty }$ on any domain $\pm \lbrack \frac{1}{2\pi }\ln
\zeta ,\infty )\times S^{1}.$
\item $\lim_{\zeta \rightarrow 0}$ $\lim_{j\rightarrow \infty }\operatorname{diam}%
\left( u_{j}|_{\varphi ^{\pm }\left( \RIfM@\expandafter\text@\else\expandafter\mbox\fi{\textrm{Ann}}_{\alpha
(\varepsilon _{j})}\right) \setminus \pm \left[ \frac{1}{2\pi }\ln \zeta
,\tau \left( \varepsilon \right) \right] \times S^{1}}\right) =0$, or
equivalently,
$\lim_{\zeta \rightarrow 0}$ $\lim_{j\rightarrow \infty }\operatorname{diam}\left(
u_{j}\left( \pm \left[ R\left( \varepsilon _{j}\right) ,\tau \left(
\varepsilon _{j}\right) +\frac{1}{2\pi }\ln \zeta \right] \times
S^{1}\right) \right) =0.$
\end{enumerate}
\end{defn}
In fact, we can turn this adiabatic convergence into a topology by
describing a neighborhood basis of the topology at $\varepsilon =0$. For any
$\varepsilon ,\zeta >0$ we define
\begin{eqnarray}
&{}&d_{adia}^{\varepsilon ,\zeta }(u,(u_{-},\chi ,u_{+})) \notag \\
:= &&\max \Big\{E_{J,\Theta _{\varepsilon }}(u),d_{\RIfM@\expandafter\text@\else\expandafter\mbox\fi{\textrm{H}}%
}(u([-R(\varepsilon ),R(\varepsilon )]\times S^{1}),\chi ([l,l]), \notag \\
&&\operatorname{diam}\left( u_{j}\left( \pm \left[ R\left( \varepsilon \right) ,\tau
\left( \varepsilon \right) +\frac{1}{2\pi }\ln \zeta \right] \times
S^{1}\right) \right) , \notag \\
&{}&d_{C_{\Sigma _{\pm }\setminus U_{\pm }(\zeta )}^{\infty }}(u(\cdot -\tau
\left( \varepsilon \right) ,u_{-}),d_{C_{\Sigma _{\pm }\setminus U_{\pm
}(\zeta )}^{\infty }}(u(\cdot +\tau \left( \varepsilon \right) ,\cdot
),u_{+})\Big\} \label{eq:adiadist}
\end{eqnarray}%
Then the sequence $u_{j}$ \emph{$\{\varepsilon _{j}\}$-adiabatically
converges} to a disc-flow-disc trajectory $(u_{-},\chi ,u_{+})$ if and only
if
\begin{equation*}
\lim_{\zeta \rightarrow 0}\lim_{j\rightarrow \infty }{}d_{adia}^{\varepsilon
_{j},\zeta }(u_{j},(u_{-},\chi ,u_{+}))=0.
\end{equation*}%
We define the open set $V_{\zeta ,\delta }^{\varepsilon }$ in ${\mathcal{F}}%
^{1,p}(K_{\varepsilon },J_{\varepsilon };z_{-},z_{+};B)$ by%
\begin{equation}
V_{\zeta ,\delta }^{\varepsilon }=\left\{ u\in {\mathcal{F}}%
^{1,p}(K_{\varepsilon },J_{\varepsilon
};z_{-},z_{+};B)|d_{adia}^{\varepsilon ,\zeta }(u,(u_{-},\chi
,u_{+}))<\delta \right\} \RIfM@\expandafter\text@\else\expandafter\mbox\fi{.} \label{eq:Ve}
\end{equation}
For a given sequence $u_{j}$ satisfying (2), we consider the
reparameterization
\begin{equation*}
\overline{u}_{j}(\tau ,t)=u_{j}\left( \frac{\tau }{\varepsilon _{j}},\frac{t%
}{\varepsilon _{j}}\right)
\end{equation*}%
on the domain $[-\varepsilon _{j}R(\varepsilon _{j}),\varepsilon
_{j}R(\varepsilon _{j})]\times {\mathbb{R}}/2\pi \varepsilon _{j}{\mathbb{Z}}
$. A straightforward calculation shows that $\overline{u}_{j}$ satisfies
\begin{equation*}
\frac{\partial \overline{u}_{j}}{\partial \tau }+J_{0}\left( \frac{\partial
\overline{u}_{j}}{\partial t}-X_{f}(\overline{u}_{j})\right) =0
\end{equation*}%
or equivalently
\begin{equation*}
\frac{\partial \overline{u}_{j}}{\partial \tau }+J_{0}\frac{\partial
\overline{u}_{j}}{\partial t}+\operatorname{grad}_{J_{0}}f(\overline{u}_{j})=0
\end{equation*}%
on $[-\varepsilon _{j}R(\varepsilon _{j}),\varepsilon _{j}R(\varepsilon
_{j})]\times {\mathbb{R}}/2\pi \varepsilon _{j}{\mathbb{Z}}$. For the
simplicity of notation, we will sometimes denote
\begin{equation*}
R_{j}=R(\varepsilon _{j}).
\end{equation*}%
The following result was proved in Part II of \cite{oh:dmj}. A similar
result was also obtained by Mundet i Riera and Tian. (See Theorem 1.3 \cite%
{MT}.)
\begin{thm}[\protect\cite{oh:dmj}, \protect\cite{MT}]
\label{thm:centrallimit} Suppose
\begin{equation*}
\ell = \lim_{j \to \infty}\varepsilon_jR(\varepsilon_j), \quad \lim_{j \to
\infty}E_{J,\Theta_{R_j}}(u_j) = 0.
\end{equation*}
Then there exists a subsequence, again denoted by $u_j$, such that the
reparameterized maps $\overline u_j$ satisfy the following:
\begin{enumerate}
\item Consider the supremum
\begin{equation*}
\operatorname{width} \overline u_j|_{[-\ell,\ell] \times S^1]}: = \sup_{\tau \in
[-\ell,\ell]} \operatorname{diam} \operatorname{Im} \overline u_j|_{\{\tau\} \times S^1}.
\end{equation*}
Then $\operatorname{width} \overline u_j|_{[-\ell,\ell] \times S^1]} \to 0$ and in
particular the center of mass of $\overline u_j:[-\ell,\ell] \times S^1$
defines a smooth path
\begin{equation*}
\operatorname{cm}(\overline u_j): [-\ell,\ell] \to M
\end{equation*}
and we can uniquely write
\begin{equation*}
\overline u_j(\tau,t) = \exp_{\operatorname{cm}(\overline u_j)}(\tau) \xi_j(\tau,t)
\end{equation*}
so that $\int_{S^1} \xi_j(\tau,t)\,dt = 0$ for all $\tau \in [-\ell,\ell]$.
\item The path $\operatorname{cm}(\overline u_j)$ converges to a gradient trajectory
$\chi:[-\ell,\ell] \to M$ satisfying $\dot \chi + \operatorname{grad}_J f(\chi) = 0$
and $\xi_j \to 0$ in $C^\infty$-topology.
\end{enumerate}
\end{thm}
Under the assumption $\lim_{j\rightarrow \infty }E_{J,\Theta _{Rj}}(u_{j})=0$%
, after taking away bubbles, on $(-\infty ,K]\times S^{1}$ of any fixed $K$,
the translated sequences $u_{j}(\cdot -\tau \left( \varepsilon _{j}\right)
,\cdot ):{\mathbb{R}}\times S^{1}\rightarrow M$ of solutions $u_{j}$ of (\ref%
{eq:duPK}) as above converge to $u_{-}:{\mathbb{R}}\times S^{1}\rightarrow M$
that satisfies the equation
\begin{equation*}
\frac{\partial u_{-}}{\partial \tau }+J\left( \frac{\partial u_{-}}{\partial
t}-X_{H_{-}}(u_{-})\right) =0
\end{equation*}%
in compact $C^{\infty }$-topology, where $H_{\pm }$ are the Hamiltonians
\begin{equation*}
H_{\pm }(\tau ,t,x)=\kappa ^{\pm }(\tau )H(t,x).
\end{equation*}%
Similar statement holds for $u_{j}(\cdot +R(\varepsilon _{j})+S\left(
\varepsilon _{j}\right) ,\cdot )$ at $+\infty $.
\section{Fredholm theory of Floer trajectories near gradient segments}
\label{sec:Qinmiddle}In this section, we study Floer trajectories near a
gradient segment $\chi $. Since $\chi $ itself is a Floer trajectory with $%
S^{1}$ symmetry, transversality is hard to achieve. We will set up
approprite Banach manifold hosting $\chi $ such that $\chi $ is a
transversal Floer trajectory. During this section we let $J$ be a $t$%
-independent almost complex structure compatible with $\omega $.
\subsection{The Banach manifold set up}
For the gluing purpose, we treat $\chi $ as a $t$-independent Floer
trajectory $u_{\chi }:[-l,l]\times S^{1}\rightarrow M$ of the equation
\begin{equation}
\frac{\partial u}{\partial \tau }+J(u)\left( \frac{\partial u}{\partial t}%
-X_{f}(u)\right) =0. \label{floer}
\end{equation}%
We choose a connection such that ${\Greekmath 0272} J=0$, and a trivialize $u_{\chi
}^{\ast }TM$ using the parallel transport. We denote the corresponding
trivialization by $\Phi :u_{\chi }^{\ast }TM\rightarrow \left[ -l,l\right]
\times \mathbb{C}^{n}$ over $[-l,l]$. In this trivialization, the
linearization of the above equation has the form
\begin{equation}
D\xi =\frac{\partial \xi }{\partial \tau }+J_{0}\frac{\partial \xi }{%
\partial t}+A\left( \tau \right) \xi \label{L-floer}
\end{equation}%
for any vector field $\xi :[-l,l]\times S^{1}\rightarrow {\mathbb{C}}^{n}$
where $J_{0}$ is the standard complex structure on ${\mathbb{C}}^{n}$,
independent of $(\tau ,t)$, and
\begin{equation*}
A=:\xi \rightarrow \left( {\Greekmath 0272} _{\chi ^{\prime }}\Phi +{\Greekmath 0272} _{\Phi
}{\Greekmath 0272} f(\chi (\tau ))\right) \xi
\end{equation*}%
is a 0-th order linear differential operator. If we change $f$ to $%
\varepsilon f$, then $\chi $ is changed to $\chi _{\varepsilon }:=\chi
\left( \varepsilon \cdot \right) $, $\Phi $ is changed to $\Phi
_{\varepsilon }:=\Phi \left( \varepsilon \cdot ,\cdot \right) $, and $A$ is
changed to $A_{\varepsilon }:=\varepsilon A$. So without loss of any
generality, we may assume that $A$ is as small as we want by considering the
Morse function $\varepsilon _{0}f$ for some small $\varepsilon _{0}$. We
denote the linearization operator of \eqref{floer} for the function $%
\varepsilon f$ by $D^{\varepsilon }$ which has the form
\begin{equation*}
D^{\varepsilon }\xi =\frac{\partial \xi }{\partial \tau }+J_{0}\frac{%
\partial \xi }{\partial t}+A_{\varepsilon }\left( \tau \right) \xi
\end{equation*}
We are going to construct a Banach manifold hosting Floer trajectories
nearby $u_{\chi }$ that matches $u_{\pm }$, on which good estimate of the
right inverse $Q$ of $D$ can be obtained. The idea is to use Fourier series
to decompose the variation vector field $\xi $ into the \textquotedblleft
zero mode" part and the \textquotedblleft higher mode" part. For the zero
mode, it is tied with the Fredholm theory of the linearized gradient
operator $L$ of Morse function $f$ that%
\begin{equation*}
L\xi =\frac{\partial \xi }{\partial \tau }+A\left( \tau \right) \xi .
\end{equation*}%
For the higher mode part, it has no zero spectrum so uniform bound of the
right inverse can be obtained. We remark that only the boundary value of the
\textquotedblleft zero mode" is compatible with the \textquotedblleft
disc-flow-disc" transversality defined in definition \ref{dfd-trans} (also
in \cite{oh-zhu}), while that of the \textquotedblleft higher modes" is not
needed thanks to its smaller norms compared to that of zero modes.
\begin{rem}
It is important to assume $J$ to be $t$-independent hence after
trivialization of $\chi ^{\ast }TM\rightarrow \lbrack -l,l]$, the resulted $%
J_{0}$ is $\left( \tau ,t\right) $-independent. Otherwise the Fourier
decomposition of the above linearized Floer equation $\left( \ref{L-floer}%
\right) $ with respect to $t\in S^{1}$ becomes difficult. On the other hand,
if we use $t$-dependent $J$, the transversality of $\chi $ as a Floer
trajectory is easier to achieve. This is the case for the transformed Floer
equation $\left( \ref{LH-Floer-eq-e}\right) $ for Lagrangian pearl complex
and will be discussed later.
\end{rem}
We define
\begin{equation}
{\mathcal{B}}_{\chi }=\{u\in W^{1,p}([-l,l]\times S^{1},M)\mid \RIfM@\expandafter\text@\else\expandafter\mbox\fi{dist}%
\left( u\left( \tau ,t\right) ,\chi \left( \tau \right) \right) <d\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ for
all }\tau ,t.\}, \label{B-mfd:FixEnds}
\end{equation}%
where $d>0$ is a small constant that for any loop $w:S^{1}\rightarrow \left(
M,g\right) $ whose image has diameter less than $d$, the center of mass of
the loop $w$ is well defined, and is denoted by cm$\left( w\right) $.
Therefore for any $u\in {\mathcal{B}}_{\chi }$, its \emph{center of mass
curve}
\begin{equation*}
\RIfM@\expandafter\text@\else\expandafter\mbox\fi{cm}\left( u\right) :\left[ -l,l\right] \rightarrow M
\end{equation*}
is well defined and is close to $\chi \left( \tau \right) $. Here
\begin{equation*}
\tau \mapsto \int_{S^1}u(\tau,t) \, dt
\end{equation*}
is the \emph{center of mass} of the loops $t \mapsto u(\tau ,t)$, and is
well defined for $u$ close enough to $\chi $: The center of mass of a loop $%
\gamma:S^1 \to M$ is defined to be the unique point $m_\gamma \in M$ such
that
\begin{equation*}
\int_0^1 \operatorname{dist}^2(m,\gamma(t)) \, dt
\end{equation*}
is the minimum. (See \cite{katcher} for the detailed exposition on the
center of mass in general).
\begin{rem}
The center of mass of a curve in a Riemannian manifold is well defined
whenever the diameter of the curve is sufficiently small. In particular the
condition $\RIfM@\expandafter\text@\else\expandafter\mbox\fi{diam}(u(\tau,t))<C\varepsilon$ enables us to define the
center of mass of the curve $u(\tau,t) \; (t\in S^1)$ when $\varepsilon$ is
sufficiently small. Therefore we can also use a Darboux chart containing the
image of $u(\tau,t)$ and may identify the curve $t \mapsto u(\tau,t)$ one in
${\mathbb{C}}^n$. With this understood, we will sometimes denote the center
of mass $\overline u(0)$ just as the average $\int_{S^1} u(\tau,t)\, dt$.
\end{rem}
Along the gradient flow $\chi \left( \tau \right) $, for any section $\xi
\in \Gamma (W^{1,p}(u_{\chi }^{\ast }TM)$), we let the average vector field
to be
\begin{equation}
\xi _{0}(\tau ):=\int_{S^{1}}\xi (\tau ,t)dt \label{VecField:flow}
\end{equation}%
and let the reduced vector field be
\begin{equation}
\tilde{\xi}(\tau ,t):=\xi (\tau ,t)-\xi _{0}(\tau ).
\label{VecField:reduced}
\end{equation}%
Then we define the Banach norm of $\xi \in T_{u_{\chi }}{\mathcal{B}}_{\chi
} $ to be
\begin{equation}
\Vert \xi \Vert _{\left( W_{\rho}^{1,p}\left[ -l,l\right] \times
S^{1}\right) }=\left\Vert \xi _{0}\right\Vert _{W^{1,p}([-l,l])}+\Vert
\tilde{\xi}\Vert _{W_{\rho }^{1,p}([-l,l]\times S^{1})}, \label{B-norm}
\end{equation}%
where
\begin{eqnarray*}
\left\Vert \xi _{0}\right\Vert _{W^{1,p}([-l,l])}^{p} &=&\int_{-l}^{l}\left(
|\xi _{0}|^{p}+|{\Greekmath 0272} _{\tau }\xi _{0}|^{p}\right) d\tau , \\
\Vert \tilde{\xi}\Vert _{W_{\rho }^{1,p}([-l,l]\times S^{1})}^{p}
&=&\int_{S^{1}}\int_{-l}^{l}\left( |\tilde{\xi}|^{p}+|{\Greekmath 0272} \tilde{\xi}%
|^{p}\right) \left( 1+\left\vert \tau \right\vert \right) ^{\delta }d\tau dt
\end{eqnarray*}%
We call $\rho \left( \tau \right) :=\left( 1+\left\vert \tau \right\vert
\right) ^{\delta }$ the weighting function for the above weighted $W_{\rho
}^{1,p}$\ norm. \ We remark that the norm $\Vert \tilde{\xi}\Vert _{W_{\rho
}^{1,p}([-l,l]\times S^{1})}$ is equivalent to the norm $\Vert \left(
1+\left\vert \tau \right\vert \right) ^{\frac{\delta }{p}}\tilde{\xi}\Vert
_{W^{1,p}([-l,l]\times S^{1})}$.
Similarly for any section $\eta \in \Gamma (W^{1,p}(u_{\chi }^{\ast
}TM)\otimes _{J}\Lambda ^{0,1}\left( [-l,l]\times S^{1}\right) $, we let%
\begin{eqnarray*}
\eta _{0}(\tau ) &:&=\int_{S^{1}}\eta (\tau ,t)dt, \\
\tilde{\eta}(\tau ,t) &:&=\eta (\tau ,t)-\eta _{0}(\tau ).
\end{eqnarray*}%
and define
\begin{eqnarray*}
\left\Vert \eta _{0}\right\Vert _{L^{p}([-l,l])}^{p} &=&\int_{-l}^{l}|\eta
_{0}|^{p}d\tau , \\
\left\Vert \tilde{\eta}\right\Vert _{L_{\rho }^{p}([-l,l]\times S^{1})}^{p}
&=&\left\Vert \left( 1+\left\vert \tau \right\vert \right) ^{\frac{\delta }{p%
}}\tilde{\eta}\left( \tau ,t\right) \right\Vert _{L^{p}([-l,l]\times S^{1})}
\\
\left\Vert \eta \right\Vert _{L_{\rho}^{p}([-l,l]\times S^{1})}
&=&\left\Vert \eta _{0}\right\Vert _{L^{p}([-l,l])}+\left\Vert \tilde{\eta}%
\right\Vert _{L_{\rho }^{p}([-l,l]\times S^{1})}
\end{eqnarray*}
Along general elements $u\in {\mathcal{B}}_{\chi }$ that are close enough to
$\chi $, (which are \textquotedblleft thin\textquotedblright\ cylinders), we
define the norm $\Vert \xi \Vert _{W_{\rho}^{1,p}}$ for $\xi \in
W^{1,p}\left( \Gamma \left( u^{\ast }TM\right) \right) $ as the following:
let
\begin{equation*}
\bar{\xi}\left( \tau ,t\right) =Pal_{\RIfM@\expandafter\text@\else\expandafter\mbox\fi{cm}\left( u\right) }^{u}\xi
\left( \tau ,t\right) ,
\end{equation*}%
where $Pal_{\RIfM@\expandafter\text@\else\expandafter\mbox\fi{cm}\left( u\right) }^{u}$ is the parallel transport from $%
u\left( \tau ,t\right) $ to cm$\left( u\right) \left( \tau \right) $ along
the shortest geodesic. Then along cm$\left( u\right) $, similar to $\chi
\left( \tau \right) $ case, we decompose
\begin{equation*}
\bar{\xi}\left( \tau ,t\right) =\left( \bar{\xi}\right) _{0}\left( \tau
\right) +\widetilde{\bar{\xi}}\left( \tau ,t\right)
\end{equation*}%
and pull back to define
\begin{equation*}
\xi _{0}=Pal_{u}^{\RIfM@\expandafter\text@\else\expandafter\mbox\fi{cm}\left( u\right) }\left( \bar{\xi}\right) _{0}%
\RIfM@\expandafter\text@\else\expandafter\mbox\fi{, \ }\tilde{\xi}=Pal_{u}^{\RIfM@\expandafter\text@\else\expandafter\mbox\fi{cm}\left( u\right) }\widetilde{\bar{\xi%
}}
\end{equation*}%
where $Pal_{u}^{\RIfM@\expandafter\text@\else\expandafter\mbox\fi{cm}\left( u\right) }$ is the parallel transport from cm%
$\left( u\right) \left( \tau \right) $ to back $u\left( \tau ,t\right) $
along the shortest geodesic. Therefore we have well defined decomposition%
\begin{equation*}
\xi \left( \tau ,t\right) =\xi _{0}\left( \tau \right) +\tilde{\xi}\left(
\tau ,t\right) .
\end{equation*}%
for $W^{1,p}\left( \Gamma \left( u^{\ast }TM\right) \right) $. Then we
define
\begin{eqnarray*}
\left\Vert \xi _{0}\right\Vert _{W^{1,p}\left( \left[ -l,l\right] \right) }
&=&\left\Vert \left( \bar{\xi}\right) _{0}\right\Vert _{W^{1,p}\left( \left[
-l,l\right] \right) } \\
\left\Vert \tilde{\xi}\right\Vert _{W_{\rho }^{1,p}\left( \left[ -l,l\right]
\times S^{1}\right) } &=&\left\Vert \widetilde{\bar{\xi}}\right\Vert
_{W_{\rho }^{1,p}\left( \left[ -l,l\right] \times S^{1}\right) } \\
\left\Vert \xi \right\Vert _{W_{\rho}^{1,p}\left( \left[ -l,l\right] \times
S^{1}\right) } &=&\left\Vert \xi _{0}\right\Vert _{W^{1,p}\left( \left[ -l,l%
\right] \right) }+\left\Vert \tilde{\xi}\right\Vert _{W_{\rho }^{1,p}\left( %
\left[ -l,l\right] \times S^{1}\right) }.
\end{eqnarray*}%
For $\eta \in L^{p}(u^{\ast }TM\otimes _{J}\Lambda ^{0,1}([-l,l]\times
S^{1}))$, the norm
\begin{equation*}
\left\Vert \eta \right\Vert _{L_{\rho}^{p}([-l,l]\times S^{1})}=\left\Vert
\eta _{0}\right\Vert _{L^{p}([-l,l])}+\left\Vert \tilde{\eta}\right\Vert
_{L_{\rho }^{p}([-l,l]\times S^{1})}
\end{equation*}%
is defined similarly by using parallel transport to cm$\left( u\right) $ and
decomposition $\eta =\eta _{0}+\tilde{\eta}$. Let
\begin{equation*}
L_{\rho}^{p}(u^{\ast }TM\otimes _{J}\Lambda ^{0,1}([-l,l]\times
S^{1}))=\left\{ \eta |\eta \in L^{p}\left( [-l,l]\times S^{1}\right)
,\left\Vert \eta \right\Vert _{L_{\rho}^{p}([-l,l]\times S^{1})}<\infty
\right\} .
\end{equation*}
We define the Banach bundle
\begin{equation*}
{\mathcal{L}}_{\chi }=\bigcup\limits_{u\in {\mathcal{B}}_{\chi
}}L_{\rho}^{p}(u^{\ast }TM\otimes _{J}\Lambda ^{0,1}([-l,l]\times S^{1})).
\end{equation*}
Then the Floer operator
\begin{equation*}
{\overline{\partial }}_{J_{0},f}:u\mapsto \left( {\partial }_{\tau }u+J_{0}{%
\partial }_{t}u+A\left( \tau \right) u\right) \otimes \left( d\tau -it\right)
\end{equation*}
gives a Fredholm section $\mathcal{F}$ of the Banach bundle ${\mathcal{L}}%
_{\chi }\rightarrow {\mathcal{B}}_{\chi }$.
\subsection{$L^{2}$ estimate of the right inverse}
For any section $\xi \in W^{1,2}(u_{\chi }^{\ast }TM)$, we take the Fourier
expansion
\begin{equation*}
\xi (\tau ,t)=\sum_{-\infty }^{\infty }a_{k}(\tau )e^{2\pi ikt},
\end{equation*}%
where $a_{k}(\tau )$ are vectors in $T_{\chi (\tau )}M$. It is easy to see
\begin{equation*}
\xi _{0}(\tau )=a_{0}(\tau ),\qquad \tilde{\xi}\left( \tau \right)
=\sum_{k\neq 0}a_{k}(\tau )e^{2\pi ikt}.
\end{equation*}%
Let $V_{0}$ and $\widetilde{V}$ be the $L^{2}$-completions of the spans of
zero Fourier mode and of higher Fourier modes respectively, then we have the
$L^{2}$-decomposition
\begin{equation*}
W^{1,2}(u_{\chi }^{\ast }TM)=V_{0}\oplus \widetilde{V}
\end{equation*}%
where we still denote by $V_{0},\,\widetilde{V}$ the intersections
\begin{equation*}
V_{0}\cap W^{1,2}(u_{\chi }^{\ast }TM),\,\widetilde{V}\cap W^{1,2}(u_{\chi
}^{\ast }TM)
\end{equation*}%
respectively. We observe that $V_{0}$ and $\widetilde{V}$ are invariant
subspaces of operator $D=\frac{\partial }{\partial \tau }+J_{0}\frac{%
\partial }{\partial t}+A(\tau )$ and so $D$ splits into
\begin{equation*}
D=D_{0}\oplus \widetilde{D}:V_{0}\oplus \widetilde{V}\rightarrow V_{0}\oplus
\widetilde{V}\RIfM@\expandafter\text@\else\expandafter\mbox\fi{,}
\end{equation*}%
where $\widetilde{D}=D|_{\widetilde{V}}:V_{0}\rightarrow V_{0}$, and $%
D_{0}=D|_{V_{0}}:\widetilde{V}\rightarrow \widetilde{V}$. Notice that $D_{0}=%
\frac{\partial }{\partial \tau }+A\left( \tau \right) $ is exactly the
linearized gradient operator $L$ of $f$.
For the construction of the right inverse $Q$ of $D$, we use Fourier
expansions of $\eta $. For any given $\eta $, we write
\begin{equation*}
\eta (\tau ,t)=\sum_{-\infty }^{\infty }b_{k}(\tau )e^{i2\pi kt}.
\end{equation*}%
Now the equation $D\xi =\eta $, i.e. $\partial _{\tau }\xi +J_{0}\partial
_{t}\xi +A(\tau )\xi =\eta $ splits into
\begin{equation}
a_{k}^{\prime }(\tau )+\left( A(\tau )-2\pi k\right) a_{k}\left( \tau
\right) =b_{k}(\tau )\qquad \RIfM@\expandafter\text@\else\expandafter\mbox\fi{for all }\;k\in {\mathbb{Z}}
\label{rightinverse}
\end{equation}%
Especially when $k=0$, it becomes
\begin{equation}
a_{0}^{\prime }\left( \tau \right) +A(\tau )a_{0}\left( \tau \right)
=b_{0}\left( \tau \right) . \label{LGradFlow}
\end{equation}
We note $\left( \ref{LGradFlow}\right) ~$is exactly the linearized gradient
flow equation. We can always solve $\left( \ref{rightinverse}\right) $
\begin{equation}
a_{k}(\tau )=e^{-\int_{0}^{\tau }\left( A(\mu )-2\pi k\right) d\mu }\left[
\int_{0}^{\tau }b_{k}(s)e^{\int_{0}^{s}\left( A(\mu )-2\pi k\right) d\mu
}ds+C_{k}\right] \label{sol}
\end{equation}%
by the variation of constants with $C_{k}$ arbitrary constant. Any choice of
$C_{k}$ will produce a right inverse of $D$.
However for the resulting right inverse to carry uniform bound independent
of $\varepsilon > 0$, we need to impose a good boundary condition on $a_k$,
which in turn requires us to make a good choice of the free constants $C_{k}$%
.
For $k=0$, we put the boundary condition
\begin{equation}
a_{0}(\pm l)= \xi_{0}(\pm l)\in d(ev_{\pm })\left( T_{(u_{\pm },o_{\pm })}%
\mathcal{M}_{2}\left( K_{\pm },J_{\pm };A_{\pm }\right) \right) ,
\label{BdryCond:0Mode}
\end{equation}%
where $ev_{\pm }:{\mathcal{M}}_{2}\left( K_{\pm },J_{\pm };A_{\pm }\right)
\rightarrow M$ are the evaluation maps $,ev_{\pm }(u_{\pm },o_{\pm })=u_{\pm
}(o_{\pm })$. The disc-flow-disc transversality condition, Proposition \ref%
{family-dfd-trans}, enables us to solve this two point boundary problem %
\eqref{BdryCond:0Mode}.
For $k\neq 0$, we impose one point boundary condition
\begin{equation}
a_{k}(l)=0\;\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ if }k>0;\qquad a_{k}(-l)=0\;\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ if }k<0.
\label{BdryCond:HigherMode}
\end{equation}%
The boundary condition can be always satisfied since the equation %
\eqref{LGradFlow} is a first order linear ODE or one can choose $C_{k}$
arbitrarily. In fact this one point condition uniquely determines $a_{k}$
for $k\neq 0$:
\begin{equation}
a_{k}(\tau )=%
\begin{cases}
-\int_{\tau }^{l}b_{k}(s)e^{\int_{\tau }^{s}\left( A(\mu )-2\pi k\right)
d\mu }ds\; & \RIfM@\expandafter\text@\else\expandafter\mbox\fi{ if }k>0, \\
\int_{-l}^{\tau }b_{k}(s)e^{\int_{\tau }^{s}\left( A(\mu )-2\pi k\right)
d\mu }ds\; & \RIfM@\expandafter\text@\else\expandafter\mbox\fi{ if }k<0.%
\end{cases}
\label{sol:ak}
\end{equation}
This constructs the right inverse $Q$ of $D$ and we denote it by $Q(\eta) =:
\xi$. From above we see $b_{0}\left( \tau \right) =0$ if and only if $%
a_{0}\left( \tau \right) =0$, thus $Q$ also splits into
\begin{equation} \label{eq:Q}
Q = Q_0 \oplus \widetilde Q: V_0 \oplus \widetilde V \to V_0 \oplus
\widetilde V
\end{equation}
where $Q_{0}=Q|_{V_{0}}$ and $\widetilde{Q}=Q|_{\widetilde{V}}$.
In fact, the above discussion shows that if the image of $Q$ becomes the
subspace
\begin{equation*}
W_{0}^{1,2}(u_{\chi }^{\ast }TM):=\{\xi \in W_{0}^{1,2}(u_{\chi }^{\ast
}TM)\mid a_{k}(\ell )=0\,\mbox{for }\,k>0,\quad a_{k}(-\ell )=0,,\mbox{for }%
\,k<0\}
\end{equation*}%
then restriction of $D$ to the subspace is an isomorphism with its inverse
given by $Q$.
\begin{rem}
The boundary condition \eqref{BdryCond:HigherMode} is geared more for $L^{2}$
estimate of $Q$ rather than matching with the $J$-holomorphic spheres $%
u_{\pm }$. We can't put two-point boundary condition for each $a_{k}(k\neq
0) $, since we have only one free constant $C_{k}$ in \eqref{sol}. There are
lots of choices of the right inverse $Q$, with various operator norm bound,
but for uniform $L^{2}$ estimate of $Q$ our choice seems to be the most
optimal one, which can be seen in the following estimate \eqref{KeyIdentify}.
\end{rem}
We estimate $\Vert \tilde{\xi}\Vert _{2}$. For $k\neq 0$,
\begin{eqnarray}
&&\left( 2\pi k\right) ^{2}\int_{-l}^{l}\left\vert a_{k}(\tau )\right\vert
^{2}d\tau \notag \\
&\leq &\int_{-l}^{l}(\left\vert a_{k}^{\prime }\right\vert ^{2}+\left( 2\pi
k\right) ^{2}\left\vert a_{k}(\tau )\right\vert ^{2}d\tau \notag \\
&=&\int_{-l}^{l}\left\vert a_{k}^{\prime }-2\pi ka_{k}\right\vert ^{2}d\tau
+2\pi k\int_{-l}^{l}2a_{k}\cdot a_{k}^{\prime }d\tau \notag \\
&=&\int_{-l}^{l}\left\vert b_{k}\left( \tau \right) -A(\tau )a_{k}\left(
\tau \right) \right\vert ^{2}d\tau +2\pi k\int_{-l}^{l}\frac{d}{d\tau }%
\left\vert a_{k}(\tau )\right\vert ^{2}d\tau \notag \\
&=&\int_{-l}^{l}\left\vert b_{k}\left( \tau \right) -A(\tau )a_{k}\left(
\tau \right) \right\vert ^{2}d\tau +2\pi k(\left\vert a_{k}(l)\right\vert
^{2}-\left\vert a_{k}(-l)\right\vert ^{2}) \notag \\
&\leq &2\left( \int_{-l}^{l}\left\vert b_{k}\right\vert ^{2}d\tau +\delta
^{2}\int_{-l}^{l}\left\vert a_{k}\right\vert ^{2}d\tau \right) +2\pi
k(\left\vert a_{k}(l)\right\vert ^{2}-\left\vert a_{k}(-l)\right\vert ^{2})
\label{KeyIdentify}
\end{eqnarray}%
provided $\left\vert A\left( \tau \right) \right\vert _{\infty }<\delta $.
By the boundary condition \eqref{BdryCond:HigherMode} of $a_{k}$, the second
summand of the last inequality is never positive, and noting $\delta
<1-1/p<1 $, so we get
\begin{equation*}
\int_{-l}^{l}\left\vert a_{k}\left( \tau \right) \right\vert ^{2}d\tau \leq
\frac{2}{\left( 2\pi k\right) ^{2}-2\delta ^{2}}\int_{-l}^{l}\left\vert
b_{k}\left( \tau \right) \right\vert ^{2}d\tau \leq \int_{-l}^{l}\left\vert
b_{k}\left( \tau \right) \right\vert ^{2}d\tau \RIfM@\expandafter\text@\else\expandafter\mbox\fi{ when }k\neq 0,
\end{equation*}%
Summing over $k\neq 0$ we get
\begin{equation}
\Vert \tilde{\xi}\Vert _{L^{2}\left[ -l,l\right] }^{2}\leq \Vert \tilde{\eta}%
\Vert _{L^{2}\left[ -l,l\right] }^{2}. \label{Estimate:L2}
\end{equation}
>From \eqref{BdryCond:HigherMode} and \eqref{KeyIdentify}, we also get
\begin{equation*}
0\leq (\left( 2\pi k\right) ^{2}-2\delta ^{2})\int_{-l}^{l}\left\vert
a_{k}\right\vert ^{2}\leq 2\int_{-l}^{l}\left\vert b_{k}\right\vert
^{2}+2\pi k(0-\left\vert a_{k}(-l)\right\vert ^{2})
\end{equation*}%
for $k>0$ and similar inequality for $k<0$. Hence
\begin{equation*}
|a_{k}(-l)|\leq \sqrt{\frac{1}{k\pi }}\Vert b_{k}\Vert _{L^{2}([-l,l])}\;%
\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ if }k>0;\qquad |a_{k}(l)|\leq \sqrt{\frac{-1}{k\pi }}\Vert b_{k}\Vert
_{L^{2}([-l,l])}\;\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ if }k<0.
\end{equation*}%
Squaring and summing over $k\neq 0$ we also have
\begin{equation}
\sum_{k>0}ka_{k}^{2}(-l)\leq \frac{1}{\pi }\Vert \tilde{\eta}\Vert
_{2}^{2},\qquad \sum_{k<0}|k|a_{k}^{2}(l)\leq \frac{1}{\pi }\Vert \tilde{\eta%
}\Vert _{2}^{2}. \label{eq:sumkak}
\end{equation}
\begin{rem}
It would be nice if we can get $C^{0}$ estimate $\tilde{\xi}(\pm l,t)$ in
terms of $\Vert \tilde{\eta}\Vert _{2}$, but it seems that we can at most
get the $W^{\frac{1}{2},2}$ norm estimate of $\tilde{\xi}(\pm l,t)$ by the
above summation inequalities. However, we will later derive the $C^{0}$
estimate of $\tilde{\xi}$ by $W^{1,p}$ estimate and Sobolev embedding.
\end{rem}
\begin{rem}
\bigskip If we extend $\eta $ to be $0$ outside $\left[ -l,l\right] $, then
we can think $\eta $ is on the full gradient trajectory $\chi ,$
corresponding to the case where $l=\infty $. Then we can still use the above
method to construct $\widetilde{\xi }$ with slightly different boundary
condition%
\begin{equation*}
a_{k}(\infty )=0\;\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ if }k>0;\qquad a_{k}(-\infty )=0\;\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ if }k<0.
\end{equation*}%
(Here $a_{k}(\pm \infty )$ makes sense, since in their defining integrals $%
b_{k}\left( \tau \right) $ is compactly supported). For such $\widetilde{\xi
}$ we gain stronger inequality
\begin{equation*}
\Vert \tilde{\xi}\Vert _{L^{2}\left( -\infty ,\infty \right) }^{2}\leq \Vert
\tilde{\eta}\Vert _{L^{2}\left[ -l,l\right] }^{2}.
\end{equation*}%
We choose this $\tilde{\xi}$ in the remaining part of our paper, because the
above stronger inequality is needed to construct the approximate right
inverse.
\end{rem}
Last we estimate $|\xi _{0}|_{W^{1,p}}$. We notice $\xi _{0}$ satisfies the
linearized gradient flow equation%
\begin{equation*}
\frac{\partial }{\partial \tau }\xi _{0}\left( \tau \right) +A(\tau )\xi
_{0}\left( \tau \right) =\eta _{0}\left( \tau \right) .
\end{equation*}%
Note the gradient flow $\chi $ can always be extended to a full gradient
flow connecting two critical points, whose Fredholm theory and
transversality has been well established. Therefore by extending $\eta $ to
be $0$ outside $\left[ -l,l\right] $ and using the right inverse for $%
L_{\chi }$ on the full gradient trajectory $\chi $, $\xi _{0}$ is always
solvable and%
\begin{equation*}
\left\Vert \xi _{0}\right\Vert _{L\left( -\infty ,\infty \right) }\leq
C_{p}\left\Vert \eta _{0}\right\Vert _{L^{p}[-l,l]},
\end{equation*}%
especially
\begin{equation}
\left\Vert \xi _{0}\right\Vert _{W^{1,p}[-l,l]}\leq C_{p}\left\Vert \eta
_{0}\right\Vert _{L^{p}[-l,l]} \label{eq:xi0leta0}
\end{equation}%
for some constant $C_{p}\left( p\geq 2\right) .$
\subsection{$L^{p}$ estimate of the right inverse}
Now assume $p\geq 2$. Without loss of generality we assume $l$ is an integer
greater than $1$ (The general case $l\geq l_{0}>0$ is similar). Let $%
B_{m}=[m-1/2,m+1/2]\times S^{1}$ and $B_{m}^{+}=[m-1,m+1]\times S^{1}$.
Given $\eta $ on $[-l,l]$, we extend it on $[-l-1,l+1]$ by letting $\eta =0$
outside $[-l,l]$. Using the method in the above section we construct $Q(\eta
)$, then restrict it on $[-l,l]$.
By the elliptic regularity we can improve $L^{2}$ estimate to $L^{p}$
estimate, if we enlarge the region a bit. In our case the inequality is of
the form
\begin{equation}
\Vert \tilde{\xi}\Vert _{W^{1,p}(B_{m})}\leq C(\Vert D\tilde{\xi}\Vert
_{L^{p}(B_{m}^{+})}+\Vert \tilde{\xi}\Vert _{L^{2}(B_{m}^{+})}) \label{L2Lp}
\end{equation}%
where the constant $C$ is independent on $m$. Summing $m$ from $-l$ to $l$,
we get
\begin{eqnarray}
\Vert \tilde{\xi}\Vert _{W^{1,p}([-l,l]\times S^{1})} &\leq &C(\Vert D\tilde{%
\xi}\Vert _{L^{p}([-l-1,l+1]\times S^{1})}+\Vert \tilde{\xi}\Vert
_{L^{2}([-l-1,l+1]\times S^{1}}) \notag \label{L2LpS} \\
&\leq &C(\Vert \tilde{\eta}\Vert _{L^{p}([-l,l]\times S^{1})}+\Vert \tilde{%
\eta}\Vert _{L^{p}([-l,l]\times S^{1})}) \notag \\
&=&2C\Vert \tilde{\eta}\Vert _{L^{p}([-l,l]\times S^{1})}
\end{eqnarray}%
where the second inequality holds because of $\tilde{\eta}=0$ on $\pm
\lbrack l,l+1]$ and the $L^{2}$ estimate \eqref{Estimate:L2}. Here $C$ is
independent of $l$.
For $\xi_{0}$, we have \eqref{eq:xi0leta0}. Altogether we get the right
inverse
\begin{equation*}
Q:L^{p}\left( [-l,l]\times S^{1}\right) \to W^{1,p}\left( [-l,l]\times
S^{1}\right) ; \, \eta \rightarrow \xi
\end{equation*}
with uniform bound
\begin{equation*}
\left\Vert Q\eta \right\Vert _{W^{1,p}\left( [-l,l]\times S^{1}\right) }\leq
C\left\Vert \eta \right\Vert _{L^{p}\left( [-l,l]\times S^{1}\right)}
\end{equation*}
with $C$ independent of the size $l$.
\begin{rem}
\label{Donaldson-Method}Since the construction of the higher mode of $\xi $
is geared for the uniform right inverse bound, not for the matching
condition, we can obtain it by an easier method: We first extend $\eta $
outside $\left[ -l,l\right] $ trivially by letting $\eta =0$ there, then we
can apply the method in chapter 3 of \cite{Don} to construct the right
inverse of higher mode part of $\frac{\partial }{\partial \tau }+J_{0}\frac{%
\partial }{\partial t}:C^{\infty }\left( \mathbb{R\times }S^{1},\mathbb{C}%
^{n}\right) \rightarrow C^{\infty }\left( \mathbb{R\times }S^{1},\mathbb{C}%
^{n}\right) $, since $J_{0}\frac{\partial }{\partial t}:C^{\infty }\left(
S^{1},\mathbb{C}^{n}\right) \rightarrow C^{\infty }\left( S^{1},\mathbb{C}%
^{n}\right) $ is $\tau $-independent operator and has nonzero spectrum when
restricted on higher mode subspace . Then we simply restrict the obtained $%
\xi =Q\eta $ on $\left[ -l,l\right] $. Then the $L^{2}$ and $L^{p}$ estimate
for $\left\Vert Q\right\Vert $ has already been obtained in \cite{Don}. For
the operator $D^{\varepsilon }=\frac{\partial }{\partial \tau }+J_{0}\frac{%
\partial }{\partial t}+A_{\varepsilon }(\tau )$, it is a $\varepsilon $%
-small perturbation of $\frac{\partial }{\partial \tau }+J_{0}\frac{\partial
}{\partial t}$, hence its right inverse bound is inherited from $Q$. In the
next section the right inverse bound can be obtained in the same way, since
for small $\delta $, the $\left( 1+\left\vert \tau \right\vert \right)
^{\delta }$ weight amounts to a $\delta $-perturbation of $\frac{\partial }{%
\partial \tau }+J_{0}\frac{\partial }{\partial t}$.
\end{rem}
\subsection{$L_{\protect\rho }^{p}$ estimate of the right inverse}
On a fixed domain $[-l,l]\times S^{1}$, the $W^{1,p}$ norm and the $W_{\rho
}^{1,p}$ norm are equivalent, they defines the same function space. But when
$l\rightarrow \infty $, the weighted norm gives better control of the
\textquotedblleft Morse-Bott" variation. This is a soft technique to get
around the point estimate of $\tilde{\xi}(\pm l/\varepsilon ,t)$ that we are
lacking.
Choose $0<\delta <1$. By conjugating with the multiplication of $\rho \left(
\tau \right) ^{\frac{1}{p}}$ where $\rho \left( \tau \right) =\left(
1+\left\vert \tau \right\vert \right) ^{\delta }$ is the weighting function,
the operator $D:W_{\rho }^{1,p}\rightarrow L_{\rho }^{p}$ is equivalent to $%
D_{\rho }:W^{1,p}\rightarrow L^{p}$, with%
\begin{equation*}
D_{\rho }=D-\frac{\left( \rho \left( \tau \right) ^{\frac{1}{p}}\right)
^{\prime }}{\left( \rho \left( \tau \right) \right) ^{\frac{1}{p}}}=D-\frac{%
\delta /p}{1+\left\vert \tau \right\vert }.
\end{equation*}%
See the following diagram%
\begin{equation*}
\xymatrix{ W^{1,p} \ar[r]^{D_{\rho }} & L^{p} \\ W_{\rho }^{1,p} \ar[r]^D
\ar[u]^{\rho \left( \tau \right) ^{\frac{1}{p}}} & L_{\rho }^{p}
\ar[u]_{\rho \left( \tau \right) ^{\frac{1}{p}}}}
\end{equation*}%
For the restriction $D=\frac{\partial }{\partial \tau }+J_{0}\frac{\partial
}{\partial t}+A(\tau )$ to the higher modes $\widetilde{V}$, $D-\frac{\delta
/p}{1+\left\vert \tau \right\vert }$ is also invertible since the
restriction $J_{0}\partial _{t}+A(\tau )$ on
\begin{equation*}
\bigoplus_{k\neq 0}E_{k}\subset L^{2}(S^{1},{\mathbb{R}}^{n})
\end{equation*}%
has its spectrum outside $(-1,1)\subset {\mathbb{R}}$ and we have choose
that $0<\frac{\delta }{p}<1$. Similarly to \eqref{L2LpS}, we obtain
\begin{equation}
\Vert \tilde{\xi}\Vert _{W_{\rho }^{1,p}([-l,l]\times S^{1})}\leq 2C\Vert
\tilde{\eta}\Vert _{L_{\rho }^{p}([-l,l]\times S^{1})} \label{L2LpSw}
\end{equation}%
where $C$ is independent on $l$.
The gain of this $W_{\rho }^{1,p}$-estimate of $\widetilde{\xi }$ is the
following pointwise decay estimate:
\begin{equation}
|\tilde{\xi}\left( \tau ,t\right) |_{C^{0}}\leq \frac{1}{\left\vert \tau
\right\vert ^{\delta }}\Vert \tilde{\xi}\Vert _{_{W_{\rho
}^{1,p}([-l,l]\times S^{1})}}\leq \frac{2C}{\left\vert \tau \right\vert
^{\delta }}\Vert \tilde{\eta}\Vert _{L_{\rho }^{p}([-l,l]\times S^{1})}
\end{equation}%
through by Sobolev embedding $W^{1,p}\hookrightarrow C^{0}$
For the zero-mode, by Sobolev embedding $W^{1,p}\left( \left[ -l,l\right]
\right) \hookrightarrow C^{\gamma }\left( \left[ -l,l\right] \right) $ $%
\left( l\geq l_{0}>0\right) $ with $\gamma =1-\frac{1}{p}$, we have
\begin{eqnarray}
|\xi _{0}\left( \tau \right) -\xi _{0}(\pm l)|_{C^{0}} &\leq &C\left\vert
\tau \pm l\right\vert ^{\gamma }\Vert \xi _{0}\Vert _{W^{1,p}([-l,l])}
\notag \\
&\leq &2C\left\vert \tau \pm l\right\vert ^{\delta }\Vert \eta _{0}\Vert
_{L^{p}([-l,l])} \label{Stablization}
\end{eqnarray}%
when $\left\vert \tau \pm l\right\vert <1$ and $0<\delta <1-1/p$.
\begin{rem}
\label{weights}The power weight $\rho \left( \tau \right) =\left(
1+\left\vert \tau \right\vert \right) ^{\delta }$ takes care of the decay of
the high modes. We choose this weight because the \emph{gradient} \emph{%
segment} $\chi $ converges to its noncritical endpoints $p_{\pm }$ in linear
order, not in exponential order. If the $\chi $ is a \emph{full gradient
trajectory} connecting two nondegenerate Morse critical points, we do not
need any weight because the higher mode $\tilde{\xi}$ with finite $W^{1,p%
\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ }}$norm automatically vanishes at $\tau =\pm \infty $. If one of the
critical end points of $\chi $ is Morse-Bott, we need to put the exponential
weight to capture the correct convergence rate. The power weight and
exponential weight can be unified by one function, say $\left(
\int_{0}^{\left\vert \tau \right\vert }\left\vert \frac{1}{{\Greekmath 0272} f\left(
\chi \right) }\right\vert d\tau \right) ^{\delta }+1$.
\end{rem}
\subsection{$\protect\varepsilon $-reparametrization of the gradient segment
and adiabatic weight}
When $\varepsilon \rightarrow 0$, we reparametrize the gradient segment $%
\chi \left( \tau \right) $ to $\chi \left( \varepsilon \tau \right) $ for
gluing. Along $\chi \left( \varepsilon \tau \right) $ we define $\varepsilon
$-family of Banach manifolds as following:%
\begin{equation*}
\Vert \xi \Vert _{W_{\rho _{\varepsilon }}^{1,p}([-l/\varepsilon
,l/\varepsilon ]\times S^{1}),}=\left\Vert \xi _{0}\left( \tau \right)
\right\Vert _{W_{\varepsilon }^{1,p}([-l/\varepsilon ,l/\varepsilon
])}+\Vert \tilde{\xi}\left( \tau \right) \Vert _{W_{\rho _{\varepsilon
}}^{1,p}([-l/\varepsilon ,l/\varepsilon ]\times S^{1}),}
\end{equation*}%
and
\begin{equation*}
\Vert \eta \Vert _{L_{\rho _{\varepsilon }}^{p}([-l/\varepsilon
,l/\varepsilon ]\times S^{1})}=\left\Vert \eta _{0}\left( \tau \right)
\right\Vert _{L_{\varepsilon }^{p}([-l/\varepsilon ,l/\varepsilon ])}+\Vert
\tilde{\eta}\left( \tau \right) \Vert _{L_{\rho _{\varepsilon
}}^{p}([-l/\varepsilon ,l/\varepsilon ]\times S^{1})}.
\end{equation*}%
Here for $\xi _{0}$ and $\eta _{0}$, the so called geometric $\varepsilon $%
-weighted norm $W_{\varepsilon }^{1,p}$ and $L_{\varepsilon }^{p}$ are
\begin{equation*}
\left\Vert \xi _{0}\right\Vert _{W_{\varepsilon }^{1,p}[-l/\varepsilon
,l/\varepsilon ]}^{p}=\int_{-l/\varepsilon }^{l/\varepsilon }\left(
\varepsilon \left\vert \xi _{0}\right\vert ^{p}+\varepsilon ^{1-p}\left\vert
{\Greekmath 0272} \xi _{0}\right\vert ^{p}\right) d\tau
\end{equation*}%
and
\begin{equation*}
\left\Vert \eta _{0}\right\Vert _{L_{\varepsilon }^{p}[-l/\varepsilon
,l/\varepsilon ]}^{p}=\int_{-l/\varepsilon }^{l/\varepsilon }\varepsilon
^{1-p}\left\vert \eta _{0}\right\vert ^{p}d\tau .
\end{equation*}%
The geometric $\varepsilon $-weight is useful in the adiabatic limit
problems, for example in \cite{FO}. One loses uniform right inverse bound of
$\frac{\partial }{\partial \tau }+\varepsilon {\Greekmath 0272} $grad$f$ on $\chi
_{\varepsilon }$ if just use usual $W^{1,p}$ norm, since the spectrum of $%
\varepsilon {\Greekmath 0272} $grad$f$ goes to $0$ as $\varepsilon \rightarrow 0$. The
norm $\left\Vert \cdot \right\Vert _{W_{\varepsilon }^{1,p}[-l/\varepsilon
,l/\varepsilon ]}$ is just the usual Sobolev norm after the
reparameterization
\begin{equation*}
R_{\varepsilon }:\chi _{\varepsilon }\rightarrow \chi ,\chi _{\varepsilon
}\left( \tau ^{\prime }\right) \rightarrow \chi \left( \tau \right) ,\RIfM@\expandafter\text@\else\expandafter\mbox\fi{
where }\tau ^{\prime }=\varepsilon \tau ,\tau ^{\prime }\in \left[
-l/\varepsilon ,l/\varepsilon \right] .
\end{equation*}%
Indeed, letting
\begin{equation*}
\widehat{\xi _{0}}\left( \tau \right) =\left( R_{\varepsilon }\right) ^{\ast
}\xi _{0}=\xi _{0}\left( \varepsilon \tau \right) ,
\end{equation*}%
then one can verify
\begin{equation*}
\left\Vert \xi _{0}\right\Vert _{W_{\varepsilon }^{1,p}[-l/\varepsilon
,l/\varepsilon ]}=\left\Vert \widehat{\xi _{0}}\right\Vert _{W^{1,p}[-l,l]}.
\end{equation*}%
In other words, under this norm
\begin{equation*}
R_{\varepsilon }:W_{\varepsilon }^{1,p}\left( \left[ -l/\varepsilon
,l/\varepsilon \right] ,M\right) \rightarrow W^{1,p}\left( \left[ -l,l\right]
,M\right)
\end{equation*}%
is an isometry between two Banach manifolds. Driven by the change of $L^{p}$%
-norm
\begin{eqnarray*}
&&\int_{-l}^{l}\left\vert \frac{\partial \chi }{\partial \tau }\left( \tau
\right) +\RIfM@\expandafter\text@\else\expandafter\mbox\fi{grad}f\left( \chi (\tau )\right) \right\vert ^{p}d\tau \\
&=&\int_{-l/\varepsilon }^{l/\varepsilon }\left\vert \frac{1}{\varepsilon }%
\frac{\partial \chi _{\varepsilon }}{\partial \tau ^{\prime }}(\tau ^{\prime
})+\RIfM@\expandafter\text@\else\expandafter\mbox\fi{grad}f\left( \chi _{\varepsilon }(\tau ^{\prime })\right)
\right\vert ^{p}\varepsilon d\tau ^{\prime } \\
&=&\int_{-l/\varepsilon }^{l/\varepsilon }\left\vert \frac{\partial \chi
_{\varepsilon }}{\partial \tau ^{\prime }}\left( \tau ^{\prime }\right)
+\varepsilon \RIfM@\expandafter\text@\else\expandafter\mbox\fi{grad}f\left( \chi _{\varepsilon }(\tau ^{\prime })\right)
\right\vert ^{p}\varepsilon ^{1-p}d\tau ^{\prime },
\end{eqnarray*}%
under the reparameterization, we define
\begin{equation*}
\left\Vert \eta _{0}\right\Vert _{L_{\varepsilon }^{p}[-l/\varepsilon
,l/\varepsilon ]}^{p}=\int_{-l/\varepsilon }^{l/\varepsilon }\varepsilon
^{1-p}\left\vert \eta _{0}\right\vert ^{p}d\tau ,
\end{equation*}%
because then the two sections
\begin{eqnarray*}
\widehat{\eta _{0}} &=&\frac{\partial }{\partial \tau }+\RIfM@\expandafter\text@\else\expandafter\mbox\fi{grad}%
f:W^{1,p}\left( \left[ -l,l\right] ,M\right) \rightarrow
\bigcup\limits_{\chi \in W^{1,p}\left( \left[ -l,l\right] ,M\right)
}L^{p}\left( \left[ -l,l\right] ,\chi ^{\ast }TM\right) \\
\eta _{0} &=&\frac{\partial }{\partial \tau ^{\prime }}+\varepsilon \RIfM@\expandafter\text@\else\expandafter\mbox\fi{%
grad}f:W_{\varepsilon }^{1,p}\left( \left[ -l/\varepsilon ,l/\varepsilon %
\right] ,M\right) \rightarrow \bigcup\limits_{\chi _{\varepsilon }\in
W_{\varepsilon }^{1,p}\left( \left[ -l/\varepsilon ,l/\varepsilon \right]
,M\right) }L_{\varepsilon }^{p}\left( \left[ -l/\varepsilon ,l/\varepsilon %
\right] ,\chi _{\varepsilon }^{\ast }M\right)
\end{eqnarray*}%
in the two Banach bundles are isometrically conjugate to each other. The
relation between the two sections is
\begin{equation}
\eta _{0}\left( \chi _{\varepsilon }\left( \tau ^{\prime }\right) \right)
=\varepsilon \widehat{\eta _{0}}\left( \chi \left( \tau \right) \right) .
\label{e-eta-relation}
\end{equation}%
It is easy to check that under these norms, the right inverse $%
Q_{0}^{\varepsilon }$ of the linearized gradient operator
\begin{equation*}
D_{0}^{\varepsilon }=\frac{\partial }{\partial \tau }+\varepsilon {\Greekmath 0272}
\RIfM@\expandafter\text@\else\expandafter\mbox\fi{grad}f:W_{\varepsilon }^{1,p}\rightarrow L_{\varepsilon }^{p}
\end{equation*}%
has uniform operator bounds over $\varepsilon >0$, since we have
\begin{equation}
\left\Vert Q_{0}^{\varepsilon }\right\Vert =\left\Vert Q_{0}\right\Vert
\label{eq:Qe0Q}
\end{equation}%
for all $\varepsilon $. The Sobolev constant
\begin{equation*}
c_{p}^{\varepsilon }:=\sup_{\xi _{0}\neq 0}\frac{\left\vert \xi
_{0}\right\vert }{\left\Vert \xi _{0}\right\Vert _{W_{\varepsilon
}^{1,p}\left( \left[ -l/\varepsilon ,l/\varepsilon \right] \right) }}%
=\sup_{\xi _{0}\neq 0}\frac{\left\vert \widehat{\xi _{0}}\right\vert }{%
\left\Vert \widehat{\xi _{0}}\right\Vert _{W^{1,p}\left( \left[ -l,l\right]
\right) }}
\end{equation*}%
is also uniform for all $\varepsilon $, since it is equal to the $W^{1,p}$
Sobolev constant of $\left\Vert \widehat{\xi _{0}}\right\Vert $on $\left[
-l,l\right] $, and we have assumed $l\geq l_{0}>0$. \
The power weight $\rho _{\varepsilon }\left( \tau \right) $ here is
transformed to
\begin{equation*}
\rho _{\varepsilon }\left( \tau \right) =\varepsilon ^{1-p}\rho \left( \tau
\right) =\varepsilon ^{1-p}\left( 1+\left\vert \tau \right\vert \right)
^{\delta },
\end{equation*}%
and it is used to define
\begin{eqnarray*}
\Vert \tilde{\xi}\left( \tau \right) \Vert _{W_{\rho _{\varepsilon
}}^{1,p}([-l/\varepsilon ,l/\varepsilon ]\times S^{1})} &=&\Vert \tilde{\xi}%
\left( \tau \right) \left( \rho _{\varepsilon }\left( \tau \right) \right) ^{%
\frac{1}{p}}\Vert _{W^{1,p}([-l/\varepsilon ,l/\varepsilon ]\times S^{1}),}%
\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ \ } \\
\Vert \tilde{\eta}\left( \tau \right) \Vert _{L_{\rho _{\varepsilon
}}^{p}([-l/\varepsilon ,l/\varepsilon ]\times S^{1})} &=&\Vert \tilde{\eta}%
\left( \tau \right) \left( \rho _{\varepsilon }\left( \tau \right) \right) ^{%
\frac{1}{p}}\Vert _{L^{p}([-l/\varepsilon ,l/\varepsilon ]\times S^{1})}.
\end{eqnarray*}%
The $\varepsilon ^{1-p}$\ factor in $\rho _{\varepsilon }\left( \tau \right)
$ is to make the norms $W_{\rho _{\varepsilon }}^{1,p}$ and $W_{\varepsilon
}^{1,p}$ comparable, which is important later in our right inverse estimate
via the weight comparison. By conjugation with the multiplication operator
by the weight function $\rho _{\varepsilon }\left( \tau \right) $,
\begin{equation*}
D^{\varepsilon }=\frac{\partial }{\partial \tau }+J_{0}\frac{\partial }{%
\partial t}+\varepsilon {\Greekmath 0272} f(\chi _{\varepsilon }^{\prime
}):W^{1,p}([-l/\varepsilon ,l/\varepsilon ]\times S^{1})\rightarrow
L^{p}([-l/\varepsilon ,l/\varepsilon ]\times S^{1})
\end{equation*}%
is equivalent to $D_{\rho _{\varepsilon }}:W_{\rho _{\varepsilon
}}^{1,p}\rightarrow L_{\rho _{\varepsilon }}^{p}$, with
\begin{equation*}
D_{\rho _{\varepsilon }}=D_{\varepsilon }-\frac{\left( \rho _{\varepsilon
}\left( \tau \right) ^{\frac{1}{p}}\right) ^{\prime }}{\rho _{\varepsilon
}\left( \tau \right) ^{\frac{1}{p}}}=D_{\varepsilon }-\frac{\delta /p}{%
1+\left\vert \tau \right\vert }.
\end{equation*}%
Since
\begin{equation*}
\frac{\Vert \tilde{\xi}\left( \tau \right) \Vert _{W_{\rho _{\varepsilon
}}^{1,p}([-l/\varepsilon ,l/\varepsilon ]\times S^{1})}}{\Vert \tilde{\eta}%
\left( \tau \right) \Vert _{L_{\rho _{\varepsilon }}^{p}([-l/\varepsilon
,l/\varepsilon ]\times S^{1})}}=\frac{\Vert \tilde{\xi}\left( \tau \right)
\Vert _{W_{\rho }^{1,p}([-l/\varepsilon ,l/\varepsilon ]\times S^{1})}}{%
\Vert \tilde{\eta}\left( \tau \right) \Vert _{L_{\rho }^{p}([-l/\varepsilon
,l/\varepsilon ]\times S^{1})}}
\end{equation*}%
the right inverse bound obtained in $W_{\rho }^{1,p}$ setting passes to the $%
W_{\rho _{\varepsilon }}^{1,p}$ setting.
\bigskip
We remark the following useful inequality for $\xi _{0}$ on $\chi \left(
\varepsilon \tau \right) $, by the Sobolev embedding $W^{1,p}\left( \left[
0,l\right] \right) \hookrightarrow C^{\gamma }\left( \left[ 0,l\right]
\right) $ with $\gamma =1-\frac{1}{p}$:
\begin{eqnarray}
\left\vert \xi _{0}\left( \tau \right) -\xi _{0}\left( \pm l/\varepsilon
\right) \right\vert &=&\left\vert \widehat{\xi _{0}}\left( \varepsilon \tau
\right) -\widehat{\xi _{0}}\left( \pm l\right) \right\vert \notag \\
&\leq &C\left( l\right) \left\vert \varepsilon \tau -\pm l\right\vert
^{\gamma }\left\Vert \widehat{\xi _{0}}\right\Vert _{W^{1,p}[0,l]} \notag \\
&=&C\left( l\right) \left\vert \varepsilon \tau \pm l\right\vert ^{\gamma
}\left\Vert \xi _{0}\right\Vert _{W_{\varepsilon }^{1,p}[-l/\varepsilon
,l/\varepsilon ]} \label{Schauder-Convergence}
\end{eqnarray}%
Especially for $\tau =\pm \left( l/\varepsilon -T\left( \varepsilon \right)
\right) ,$ where $T\left( \varepsilon \right) =\frac{1}{3}\frac{p-1}{\delta }%
S\left( \varepsilon \right) $, we have%
\begin{eqnarray*}
\left\vert \xi _{0}\left( \tau \right) -\xi _{0}\left( \pm l/\varepsilon
\right) \right\vert &\leq &C\left( l\right) \left( \varepsilon T\left(
\varepsilon \right) \right) ^{\gamma }\left\Vert \xi _{0}\right\Vert
_{W_{\varepsilon }^{1,p}[-l/\varepsilon ,l/\varepsilon ]} \\
&\leq &C\left( l\right) \left\vert \varepsilon \ln \varepsilon \right\vert
^{\gamma }\left\Vert \xi _{0}\right\Vert _{W_{\varepsilon
}^{1,p}[-l/\varepsilon ,l/\varepsilon ]} \\
&\leq &C\left( l\right) \varepsilon ^{\widetilde{\gamma }}\left\Vert \xi
_{0}\right\Vert _{W_{\varepsilon }^{1,p}[-l/\varepsilon ,l/\varepsilon ]}
\end{eqnarray*}%
for any $0<\widetilde{\gamma }<1-\frac{1}{p}$, if $\varepsilon $ is
sufficiently small.
\section{Moduli space of \textquotedblleft disc-flow-disc\textquotedblright\
configurations}
\label{subsec:disc-flow-disc}
This subsection is the first stage of the deformation of the parameterized
moduli space entering in the construction of the chain homotopy map between $%
\Psi \circ \Phi $ and the identity on $HF(H,J)$ in \cite{PSS}. The material
in the first half of this section is largely taken from section 5.1 \cite%
{oh-zhu} with slight modifications.
A ``disk-flow-disk" configuration consists of two perturbed $J$-holomorphic
discs joined by a gradient flow line between their marked points. In this
section we will define the moduli space of such configurations.
For notation brevity, we just denote
\begin{equation*}
{\mathcal{M}}^{l}(K^{\pm },J^{\pm };[z_{-},w_{-}];f;[z_{+},w_{+}];A_{\pm })={%
\mathcal{M}}^{l}([z_{-},w_{-}];f;[z_{+},w_{+}];A_{\pm }),
\end{equation*}%
omitting the Floer datum $(K^{\pm },J^{\pm })$, as long as it does not cause
confusion.
Given the two moduli spaces ${\mathcal{M}}([z_{-},w_{-}];A_{-})$ and ${%
\mathcal{M}}([z_{+},w_{+}];A_{+})$ and the Morse function $f$, we define the
moduli space
\begin{equation*}
{\mathcal{M}}^{l}([z_{-},w_{-}];f;[z_{+},w_{+}];A_{\pm })
\end{equation*}%
by the set consisting of \textquotedblleft \textquotedblleft disk-flow-disk"
configurations $(u_{-},\chi ,u_{+})$ of \emph{flow time} $2l$ such that
\index{${\mathcal{M}}^{l}([z_{-},w_{-}];f;[z_{+},w_{+}];A_{\pm })$}
\begin{equation}
{\mathcal{M}}^{l}([z_{-},w_{-}];f;[z_{+},w_{+}];A_{\pm })=\left\{
(u_{-},\chi ,u_{+})\left\vert
\begin{array}{c}
u_{\pm }\in {\mathcal{M}}(K^{\pm },J^{\pm };%
\vec{z}_{\pm };A_{\pm }),\, \\
\left[ u_{\pm }\#w_{\pm }\right] =A_{\pm },\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ }\chi :[-l,l]\rightarrow M,
\\
\dot{\chi}-{\Greekmath 0272} f(\chi )=0, \\
u_{-}(o_{-})=\chi (-l),u_{+}(o_{+})=\chi (l).%
\end{array}%
\right. \right\}
\end{equation}%
Then the moduli space of \textquotedblleft disk-flow-disk" configurations is
defined to be
\index{${\mathcal{M}}^{para}([z_{-},w_{-}];f;[z_{+},w_{+}];A_{\pm })$}
\begin{equation}
{\mathcal{M}}^{para}([z_{-},w_{-}];f;[z_{+},w_{+}];A_{\pm }):=\bigcup_{l\geq
l_{0}}{\mathcal{M}}^{l}([z_{-},w_{-}];f;[z_{+},w_{+}];A_{\pm }).
\end{equation}
We now provide the off-shell formulation of the \textquotedblleft
disk-flow-disk" moduli spaces. We first provide the Banach manifold hosting $%
{\mathcal{M}}^{l}([z_{-},w_{-}];f;[z_{+},w_{+}];A_{\pm })$. We define
\begin{multline}
{\mathcal{B}}_{l}^{dfd}(z_{-},z_{+}):=\{(u_{-},\chi ,u_{+})\mid u_{\pm }\in
W^{1,p}(%
\dot{\Sigma},M;z_{\pm }), \label{Bmfd} \\
\chi \in W^{1,p}([-l,l],M),u_{-}(o_{-})=\chi (-l),\;u_{+}(o_{+})=\chi (l)\}
\end{multline}%
for $p>2$. For any $u\in {\mathcal{B}}_{l}^{dfd}(z_{-},z_{+})$, the tangent
space is%
\begin{equation*}
T_{u}{\mathcal{B}}_{l}^{dfd}(z_{-},z_{+})=\Big\{(\xi _{-},a,\xi _{+})\mid
\xi _{\pm }\in W^{1,p}(u_{\pm }^{\ast }TM),\xi _{-}(o_{-})=a(-l),\,\xi
_{+}(o_{+})=a(l)\Big\}.
\end{equation*}%
Then for each $u=(u_{-},\chi ,u_{+})\in {\mathcal{B}}_{l}^{dfd}(z_{-},z_{+})$%
, we define
\begin{equation*}
L_{u}^{p}(z_{-},z_{+})=L^{p}(\Lambda ^{(0,1)}u^{\ast }TM)
\end{equation*}%
and form the Banach bundle
\begin{equation*}
{\mathcal{L}}_{l}^{dfd}=\bigcup_{u\in {\mathcal{B}}%
_{l}^{dfd}(z_{-},z_{+})}L_{u}^{p}(z_{-},z_{+})
\end{equation*}%
over ${\mathcal{B}}_{l}^{dfd}(z_{-},z_{+})$. Here the superscript `dfd'
stands for `disk-flow-disk'. We refer to \cite{Fl} for a more detailed
description of the asymptotic behavior of the elements in ${\mathcal{B}}%
_{l}^{dfd}(z_{-},z_{+})$ in the context of Floer moduli spaces.
For $u=(u_-,\chi,u_+)\in {\mathcal{B}}^{dfd}_{l}(z_-,z_+)$, its tangent
space $T_u{\mathcal{B}}^{dfd}_{l}$ consists of $\xi=(\xi_-,a,\xi_+)$, where $%
\xi_{\pm}\in W^{1,p}(u_{\pm}^*TM)$, $a\in W^{1,p}(\chi^*TM)$, with the
matching condition
\begin{equation} \label{match}
\xi_-(o_-)=a(-l),\quad \xi_+(o_+)=a(l)
\end{equation}
We denote the set of such $\xi$ as
\index{$W^{1,p}_u(z_-,z_+)$} $W^{1,p}_u(z_-,z_+)$.
We let
\index{${\mathcal{B}}^{dfd}(z_-,z_+)$}
\begin{equation*}
{\mathcal{B}}^{dfd}(z_-,z_+)= \bigcup_{l\ge l_0} {\mathcal{B}}%
^{dfd}_{l}(z_-,z_+) \quad
\RIfM@\expandafter\text@\else\expandafter\mbox\fi{and} \quad {\mathcal{L}}^{dfd}(z_-,z_+)=\bigcup_{l\ge l_0} {\mathcal{L}%
}^{dfd}_{l}(z_-,z_+)
\end{equation*}
\begin{rem}
If we regard $u$ in ${\mathcal{B}}^{dfd}(z_{-},z_{+})$ instead of ${\mathcal{%
B}}_{l}^{dfd}(z_{-},z_{+})$, then
\begin{equation*}
T_{u}{\mathcal{B}}^{dfd}(z_{-},z_{+})\simeq T_{u}{\mathcal{B}}%
_{l}^{dfd}(z_{-},z_{+})\times T_{l}\mathbb{R},
\end{equation*}%
namely the tangent space of $u$ in ${\mathcal{B}}^{dfd}(z_{-},z_{+})$
consists of $\xi =(\xi _{-},a,\xi _{+},\mu )$, where $\xi _{\pm }\in
W^{1,p}(u_{\pm }^{\ast }TM)$, $a\in W^{1,p}(\chi ^{\ast }TM),$ and $\mu \in
T_{l}{\mathbb{R}}\cong {\mathbb{R}}$, with the matching condition
\begin{equation}
\xi _{-}(o_{-})=a(-l)-\frac{\mu }{l}\dot{\chi}(-l),\quad \xi
_{+}(o_{+})=a(l)+\frac{\mu }{l}\dot{\chi}(l). \label{p-match}
\end{equation}%
Here the $\mu $ comes from the variation of the length $l$ of the domain of
gradient flows. For $\mu \in T_{l}\mathbb{R}$, the induced path in ${%
\mathcal{B}}^{dfd}(z_{-},z_{+})$, starting from $u=(u_{-},\chi ,u_{+})\in {%
\mathcal{B}}_{l}^{dfd}(z_{-},z_{+})$, is $u_{s}=(u_{-},\chi _{s},u_{+})\in {%
\mathcal{B}}_{\left( l-s\mu \right) }^{dfd}(z_{-},z_{+})$, where $\chi
_{s}(\tau ^{\prime })$ is the reparameterization of $\chi (\tau )$,%
\begin{equation*}
\chi _{s}(\tau ^{\prime }):=\chi \left( \frac{l\tau ^{\prime }}{l-s\mu }%
\right) ,\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ \ \ \ \ \ }\tau ^{\prime }\in \left[ -\left( l-s\mu \right)
,l-s\mu \right] ,
\end{equation*}%
for $s$ nearby $0$. There is a canonical way to associate points on $\chi $
to points on $\chi _{s}:$%
\begin{equation*}
\chi \left( \tau \right) \longleftrightarrow \chi _{s}\left( \tau ^{\prime
}\right) ,\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ \ \ \ \ \ where }\tau =\frac{l\tau ^{\prime }}{l-s\mu }.
\end{equation*}%
They all correspond to the same point on the image of $\chi $.
\end{rem}
Now we fix $l>0$ and consider a natural section
\begin{equation}
e:{\mathcal{B}}_{l}^{dfd}(z_{-},z_{+})\rightarrow {\mathcal{L}}%
_{l}^{dfd}(z_{-},z_{+})
\end{equation}%
such that $e(u)\in L_{u}^{p}(z_{-},z_{+})$ is given by
\begin{equation*}
e(u)=({\overline{\partial }}_{(K_{-},J_{-})}u_{-},\dot{\chi}-{\Greekmath 0272} f(\chi ),%
{\overline{\partial }}_{(K_{+},J_{+})}u_{+})
\end{equation*}%
where the $u_{\pm }$ and $\chi $ satisfy the matching condition in %
\eqref{Bmfd}. The linearization of $e$ at $u\in e^{-1}(0)={\mathcal{M}}%
^{l}([z_{-},w_{-}];f;[z_{+},w_{+}];A_{\pm })$ induces a linear operator
\begin{equation}
E(u):=D_{u}e:T_{u}{\mathcal{B}}_{l}^{dfd}(z_{-},z_{+})\rightarrow
L_{u}^{p}(z_{-},z_{+})
\end{equation}%
where we have
\begin{equation*}
T_{u}{\mathcal{B}}_{l}^{dfd}(z_{-},z_{+})=\Big\{(\xi _{-},a,\xi _{+})\mid
\xi _{\pm }\in W^{1,p}(u_{\pm }^{\ast }TM),\xi _{-}(o_{-})=a(-l),\,\xi
_{+}(o_{+})=a(l)\Big\}
\end{equation*}%
and the value $D_{u}e(\xi )=:\eta $ has the expression
\begin{equation*}
\eta =(\eta _{-},b,\eta _{+})=\left( D_{u_{-}}{\overline{\partial }}%
_{(K_{-},J_{-})}(\xi _{-}),\,\frac{Da}{d\tau }-{\Greekmath 0272} _{a}\operatorname{grad}%
(f),D_{u_{+}}{\overline{\partial }}_{(K_{+},J_{+})}(\xi _{+})\,\right)
\end{equation*}%
for $\xi =(\xi _{-},a,\xi _{+})$. We remark that if we regard $u\in {%
\mathcal{B}}^{dfd}(z_{-},z_{+})$, then for $(\xi _{-},a,\xi _{+},\mu )\in $ $%
T_{u}{\mathcal{B}}^{dfd}(z_{-},z_{+}),$ $D_{u}e(\xi )=:\eta $ has the
expression $\eta =(\eta _{-},b,\eta _{+})$ where
\begin{equation}
b=\frac{D}{d\tau }a-{\Greekmath 0272} _{a}\operatorname{grad}(f)+\frac{\mu }{l}\dot{\chi}\left(
\tau \right) . \label{eq:b=}
\end{equation}%
Here we have used that
\begin{eqnarray*}
D_{u}e(0,0,0,\mu ) &=&\left. \frac{d}{ds}\right\vert _{s=0}\left[ \frac{%
d\chi _{s}}{d\tau ^{\prime }}\left( \tau ^{\prime }\right) -{\Greekmath 0272} f\left(
\chi _{s}\left( \tau ^{\prime }\right) \right) \right] \\
&=&\left. \frac{d}{ds}\right\vert _{s=0}\left[ \frac{d\chi }{d\tau ^{\prime }%
}\left( \tau \right) -{\Greekmath 0272} f\left( \chi \left( \tau \right) \right) \right]
\\
&=&\left. \frac{d}{ds}\right\vert _{s=0}\left[ \frac{l}{l-s\mu }\frac{d\chi
}{d\tau }\left( \tau \right) \right] =\frac{\mu }{l}\dot{\chi}\left( \tau
\right) .
\end{eqnarray*}%
For the simplicity of notation, we denote
\begin{eqnarray*}
W_{u}^{1,p}(z_{-},z_{+};dfd):= &&T_{u}{\mathcal{B}}_{l}^{dfd}(z_{-},z_{+}) \\
&\subset &W^{1,p}(u_{-}^{\ast }TM)\times W^{1,p}(\chi ^{\ast }TM)\times
W^{1,p}(u_{+}^{\ast }TM).
\end{eqnarray*}%
Now we show $E(u)$ is Fredholm and compute its index.
\begin{prop}
\label{prop:dfdindex} The operator $E(u)$ is a Fredholm operator and we have
\begin{equation}
\operatorname{Index}E(u)=\mu _{H_{-}}([z_{-},w_{-}])-\mu
_{H_{+}}([z_{+},w_{+}])+2c_{1}(A_{-})+2c_{1}(A_{+}).
\end{equation}%
for any
\begin{equation*}
u=(u_{-},\chi ,u_{+})\in {\mathcal{M}}^{l}([z_{-},w_{-}];f;[z_{+},w_{+}];A_{%
\pm }).
\end{equation*}
\end{prop}
\begin{proof}
We compute the kernel and the cokernel of
\begin{equation*}
E(u):W_{u}^{1,p}(z_{-},z_{+};dfd)\rightarrow L_{u}^{p}(z_{-},z_{+}).
\end{equation*}%
By the matching condition \eqref{p-match} it is clear that
\begin{multline}
\operatorname{ker}E(u)=\Big\{(\xi _{-},\xi _{+},a)\mid \xi _{\pm }\in \operatorname{ker}%
D_{u_{\pm }}{\overline{\partial }}_{(K^{\pm },J^{\pm })}, \\
\frac{Da}{\partial \tau }-{\Greekmath 0272} _{a}\operatorname{grad}f(\chi )=0,\xi
_{-}(o_{-})=a(-l),\,\xi _{+}(o_{+})=a(l)\Big\},
\end{multline}%
It is easy to see
\begin{equation*}
\xi _{+}(o_{+})=d\phi _{f}^{2l}\cdot \xi _{-}(o_{-})
\end{equation*}%
for any $(\xi _{-},\xi _{+},a)\in \ker E(u)$ noticing that $a$ is determined
by its initial value $a(-l)$ and by the equation
\begin{equation}
\frac{Da}{\partial \tau }-{\Greekmath 0272} _{a}\operatorname{grad}f(\chi )=0. \label{eq:a}
\end{equation}%
Therefore the $(\xi _{-},\xi _{+},a)\in \ker E(u)$ have 1-1 correspondence
to the points of $\left[ d\left( \phi _{f}^{2l}ev_{-}\times ev_{+}\right) %
\right] ^{-1}\left( \Delta _{u_{+}\left( o_{+}\right) }^{tn}\right) ,$ where
\begin{equation*}
d\left( \phi _{f}^{2l}ev_{-}\times ev_{+}\right) :\ker D_{u_{-}}{\overline{%
\partial }}_{(K^{-},J^{-})}\times \ker D_{u_{+}}{\overline{\partial }}%
_{(K_{+},J_{+})}\rightarrow T_{u_{+}\left( o_{+}\right) }M\times
T_{u_{+}\left( o_{+}\right) }M
\end{equation*}%
and $\Delta _{u_{+}\left( o_{+}\right) }^{tn}=\left\{ \left( v,v\right)
|v\in T_{u_{+}\left( o_{+}\right) }M\right\} \subset $ $T_{u_{+}\left(
o_{+}\right) }M\times T_{u_{+}\left( o_{+}\right) }M$, which has codimension
$2n$. Let the subspaces
\begin{equation*}
V_{\pm }:=ev_{o_{\pm }}\left( \ker D_{u_{\pm }}{\overline{\partial }}%
_{(K^{\pm },J^{\pm })}\right) \subset T_{u_{\pm }\left( o_{\pm }\right) }M.
\end{equation*}%
We have
\begin{eqnarray}
\mathop{\kern0pt{\rm dim}}\nolimits\ker E(u) &=&\mathop{\kern0pt{\rm dim}}%
\nolimits\ker D_{u_{-}}{\overline{\partial }}_{(K^{-},J^{-})}+%
\mathop{\kern0pt{\rm dim}}\nolimits\ker D_{u_{+}}{\overline{\partial }}%
_{(K_{+},J_{+})}-2n \notag \\
&{}&\quad +4n-\mathop{\kern0pt{\rm dim}}\nolimits\left( \left( d\phi
_{f}^{2l}\cdot V_{-}\right) \times V_{+}+\Delta _{u_{+}\left( o_{+}\right)
}^{tn}\right) \notag \\
&=&\mathop{\kern0pt{\rm dim}}\nolimits\ker D_{u_{-}}{\overline{\partial }}%
_{(K^{-},J^{-})}+\mathop{\kern0pt{\rm dim}}\nolimits\ker D_{u_{+}}{\overline{%
\partial }}_{(K_{+},J_{+})} \notag \\
&&-\mathop{\kern0pt{\rm dim}}\nolimits\left( d\phi _{f}^{2l}\cdot
V_{-}+V_{+}\right) \label{kerdim:dfd}
\end{eqnarray}%
where in the last identity we have used the linear algebra fact that
\begin{equation*}
\mathop{\kern0pt{\rm dim}}\nolimits\left( A\times B+\Delta \right) =%
\mathop{\kern0pt{\rm dim}}\nolimits\Delta +\mathop{\kern0pt{\rm dim}}%
\nolimits\left( A-B\right)
\end{equation*}%
for linear subspaces $A,B$ in $V$ and diagonal $\Delta $ in $V\times V$.
Next we compute the cokernel of $E(u)$. Let $E(u)^{\ast }$ be the adjoint
operator of $E(u)$, such that
\begin{equation*}
E(u)^{\ast }:L_{u}^{p}(z_{-},z_{+})^{\ast }\rightarrow
W^{1,p}(z_{-},z_{+};dfd)^{\ast }.
\end{equation*}%
Using the nondegenerate $L^{2}$ pairing
\begin{equation*}
L^{p}(\Lambda ^{(0,1)}u^{\ast }TM)\times L^{q}(\Lambda ^{(1,0)}u^{\ast
}TM)\rightarrow {\mathbb{R}}
\end{equation*}%
we identify $L_{u}^{p}(z_{-},z_{+})^{\ast }$ with $L^{q}(\Lambda
^{(1,0)}u^{\ast }TM)$. On the other hand, we can identify $%
W^{1,p}(z_{-},z_{+};dfd)^{\ast }$ with
\begin{eqnarray*}
&{}& W^{-1,q}(z_{-},z_{+};dfd) \\
&:= &\left\{ (\xi _{-},a,\xi _{+})\in W^{1,p}(z_{-},z_{+})\mid \xi
_{-}(o_{-})=a(-l),\,\xi _{+}(o_{+})=a(l)\right\} ^{\perp }
\end{eqnarray*}%
in the direct product
\begin{equation*}
W^{-1,q}(z_{-},z_{+})=W^{-1,q}(u_{-}^{\ast }TM)\times W^{-1,q}(\chi ^{\ast
}TM)\times W^{-1,q}(u_{+}^{\ast }TM)
\end{equation*}%
where $(\cdot )^{\perp }$ denotes the $L^{2}$-orthogonal complement. Here we
have $1<q<2$ since $2<p<\infty $.
We denote by
\begin{equation*}
E(u)^\dagger: L^q(\Lambda^{(1,0)}u^*TM) \to W^{-1,q}(z_-,z_+;dfd)
\end{equation*}
the corresponding $L^2$-adjoint with respect to these identifications.
Now we derive the formula for $E(u)^{\dagger }$. Recall by definition, we
have
\begin{equation*}
\langle E(u)\xi ,\eta \rangle =\langle \xi ,E(u)^{\dagger }\eta \rangle .
\end{equation*}%
Then for any given $\eta :=(\eta _{-},b,\eta _{+})\in \operatorname{ker}E^{\dagger
}(u)\subset L^{q}(z_{-},z_{+})$ it satisfies
\begin{eqnarray*}
0 &=&\int_{-l}^{l}\left\langle \frac{Da}{\partial \tau }-{\Greekmath 0272} \operatorname{grad}%
f(\chi )a,b\right\rangle \\
&{}&\quad +\int_{\dot{\Sigma}_{-}}\left\langle D_{u_{-}}{\overline{\partial }%
}_{(K^{-},J^{-})}\xi _{-},\eta _{-}\right\rangle +\int_{\dot{\Sigma}%
_{+}}\left\langle D_{u_{+}}{\overline{\partial }}_{(K_{+},J_{+})}\xi
_{+},\eta _{+}\right\rangle
\end{eqnarray*}%
for all $(\xi _{+},a,\xi _{-})\in W_{u}^{1,p}(z_{-},z_{+};dfd)$, i.e., for
all the triples satisfying the matching condition
\begin{equation}
\xi _{-}(o_{-})=a(-l),\quad \xi _{+}(o_{+})=a(l). \label{eq:matching}
\end{equation}%
Integrating by parts, we have
\begin{eqnarray}
0 &=&\langle a(l),b(l)\rangle -\langle a(-l),b(-l)\rangle
+\int_{-l}^{l}\left\langle -\frac{Db}{\partial \tau }-{\Greekmath 0272} \operatorname{grad}%
f(\chi )b,a\right\rangle \notag \label{eq:0=aebe} \\
&{}&\quad -\int_{\dot{\Sigma}_{-}}\langle (D_{u_{-}}{\overline{\partial }}%
_{(K^{-},J^{-})})^{\dagger }\eta _{-},\xi _{-}\rangle -\int_{\dot{\Sigma}%
_{+}}\langle (D_{u_{+}}{\overline{\partial }}_{(K^{+},J^{+})})^{\dagger
}\eta _{+},\xi \rangle .
\end{eqnarray}%
Here $D_{u_{\pm }}\partial _{(K^{\pm },J^{\pm })})$ is the formal adjoint of
$D_{u_{\pm }}{\overline{\partial }}_{(K^{\pm },J^{\pm })})$ which has its
symbol of that of the Dolbeault operator $\partial $ (near $z=0$ in ${%
\mathbb{C}}$) and so elliptic. \emph{Here we note that we are using a metric
on $\Sigma _{\pm }\cong {\mathbb{C}}$ that is standard near the origin }$%
o_{\pm }$ \emph{and cylindrical near the end.}
Substituting \eqref{eq:matching} into this we can rewrite \eqref{eq:0=aebe}
into
\begin{eqnarray*}
0 &=&\int_{-l}^{l}\left\langle -\frac{Db}{\partial \tau }-{\Greekmath 0272} \operatorname{grad}%
f(\chi )b,a\right\rangle +\langle \xi _{+}(o_{+}),b(\varepsilon )\rangle
-\langle \xi _{-}(o_{-}),b(0)\rangle \\
&{}&\quad -\int_{\dot{\Sigma}_{-}}\langle (D_{u_{-}}{\overline{\partial }}%
_{(K^{-},J^{-})})^{\dagger }\eta _{-},\xi _{-}\rangle -\int_{\dot{\Sigma}%
_{+}}\langle (D_{u_{+}}{\overline{\partial }}_{(K^{+},J^{+})})^{\dagger
}\eta _{+},\xi \rangle .
\end{eqnarray*}%
Note $a$ can be varied arbitrarily on the interior $(-l,l)$ and can be
matched to any given $\xi _{\pm }(o_{\pm })$ at $-l,\,l$. Therefore
considering the variation $\xi _{-}=\xi _{+}=0$, we derive that $b$ must
satisfy
\begin{equation*}
\left\langle -\frac{Db}{\partial \tau }-{\Greekmath 0272} \operatorname{grad}f(\chi
)b,a\right\rangle =0
\end{equation*}%
for all $a$ with $a(-l)=0=a(l)$. Therefore $b$ satisfies
\begin{equation}
-\frac{Db}{\partial \tau }-{\Greekmath 0272} \operatorname{grad}f(\chi )b=0 \label{eq:b}
\end{equation}%
on $[-l,l]$ first in the distribution sense and then in the classical sense
by the bootstrap regularity of the ODE \eqref{eq:b} and so it is smooth. \
Let
\begin{equation*}
P^{\dag }:T_{u_{-}\left( o_{-}\right) }M\rightarrow T_{u_{+}\left(
o_{+}\right) }M,\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ }b\left( -l\right) \rightarrow b\left( l\right)
\end{equation*}%
be the linear map for solutions $b$ of ODE $\left( \ref{eq:b}\right) $, then
\
\begin{equation*}
b\left( l\right) =P^{\dag }b\left( -l\right) .
\end{equation*}
Substituting \eqref{eq:b} into the above we obtain
\begin{eqnarray*}
0 &=&-\langle \xi _{-}(o_{-}),b(-l)\rangle +\int_{\dot{\Sigma}_{-}}\langle
(D_{u_{-}}{\overline{\partial }}_{(K^{-},J^{-})})^{\dagger }\eta _{-},\xi
_{-}\rangle \\
&{}&\quad +\langle \xi _{+}(o_{+}),b(l))\rangle +\int_{\dot{\Sigma}%
_{+}}\langle (D_{u_{+}}{\overline{\partial }}_{(K^{+},J^{+})})^{\dagger
}\eta _{+},\xi _{+}\rangle .
\end{eqnarray*}%
Now we can vary $\xi _{\pm }$ independently and hence we have
\begin{eqnarray*}
0 &=&-\langle \xi _{-}(o_{-}),b(-l)\rangle -\int_{\dot{\Sigma}_{-}}\langle
(D_{u_{-}}{\overline{\partial }}_{(K^{-},J^{-})})^{\dagger }\eta _{-},\xi
_{-}\rangle \\
0 &=&\langle \xi _{+}(o_{+}),b(l)\rangle +\int_{\dot{\Sigma}_{+}}\langle
(D_{u_{+}}{\overline{\partial }}_{(K^{+},J^{+})})^{\dagger }\eta _{+},\xi
_{+}\rangle
\end{eqnarray*}%
and hence
\begin{eqnarray}
(D_{u_{-}}{\overline{\partial }}_{(K^{-},J^{-})})^{\dagger }\eta
_{-}-b(-l)\delta _{o_{-}} &=&0 \notag \\
(D_{u_{+}}{\overline{\partial }}_{(K^{+},J^{+})})^{\dagger }\eta
_{+}+b(l)\delta _{o_{+}} &=&0. \label{cdbar}
\end{eqnarray}%
Here $\delta _{o}$ denotes the Dirac-delta measure supported at the point $%
\{o\}\subset \Sigma $. Due to the choice of our metric on the domain ${%
\mathbb{C}}$ of $u_{\pm }$, $\eta _{\pm }$ must have the singularity of the
type $\frac{1}{\bar{z}}$ which is the fundamental solution to $\partial \eta
=\vec{b}\delta _{o}$ which lies in $L^{q}$ for any $1<q<2$. Therefore one
can solve the distributional equation
\begin{equation*}
(D_{u_{\pm }}{\overline{\partial }}_{(K^{\pm },J^{\pm })})^{\dagger }\eta =%
\vec{b}\cdot \delta _{o_{\pm }}
\end{equation*}%
provided that $\vec{b}$ satisfies the Fredholm alternative:
\begin{equation*}
\langle \vec{b},\xi _{\pm }\left( o_{\pm }\right) \rangle =\int_{\dot{\Sigma}%
_{\pm }}\langle \vec{b}\cdot \delta _{o_{\pm }},\xi _{\pm }\rangle =0\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ }
\end{equation*}%
for all $\xi _{\pm }\in \ker \left( (D_{u_{\pm }}{\overline{\partial }}%
_{(K^{\pm },J^{\pm })})^{\dagger }\right) ^{\dagger }=\ker D_{u_{\pm }}{%
\overline{\partial }}_{(K^{\pm },J^{\pm })}$. Namely $\vec{b}\in \left(
V_{\pm }\right) ^{\perp }$, where
\begin{equation*}
V_{\pm }:=ev_{o_{\pm }}\left( \ker D_{u_{\pm }}{\overline{\partial }}%
_{(K^{\pm },J^{\pm })}\right) .
\end{equation*}%
We fix such a solution denoted by $\eta _{\vec{b}}\in L^{q}$.
Then \eqref{cdbar} can be written as
\begin{equation*}
(D_{u_{-}}{\overline{\partial }}_{(K^{-},J^{-})})^{\dagger }(\eta _{-}-\eta
_{b(-l)})=0,\quad (D_{u_{+}}{\overline{\partial }}_{(K^{+},J^{+})})^{\dagger
}(\eta _{+}+\eta _{b(l)})=0
\end{equation*}%
i.e.,
\begin{eqnarray*}
\eta _{-}+\eta _{b(-l)} &\in &\operatorname{ker}(D_{u_{-}}{\overline{\partial }}%
_{(K_{-},J_{-})})^{\dagger }, \\
\eta _{+}-\eta _{b(l)} &\in &\operatorname{ker}(D_{u_{+}}{\overline{\partial }}%
_{(K_{+},J_{+})})^{\dagger }.
\end{eqnarray*}%
Therefore we have the exact sequence
\begin{eqnarray*}
0& \rightarrow & \operatorname{Graph}P^{\dag }\cap \left( V_{-}^{\perp }\times
V_{+}^{\perp }\right) \overset{i}{\rightarrow }\operatorname{ker}E^{\dagger }(u) \\
&\overset{j}{\rightarrow }& \operatorname{ker}(D_{u_{+}}{\overline{\partial }}%
_{(K_{+},J_{+})})^{\dagger }\oplus \operatorname{ker}(D_{u_{-}}{\overline{\partial }}%
_{(K_{-},J_{-})})^{\dagger }\rightarrow 0:
\end{eqnarray*}%
Here the first homomorphism is the map
\begin{equation*}
i(b_{1},b_{2})=(\eta _{b_{1}},b_{b_{1}},-\eta _{b_{2}})
\end{equation*}%
where $b_{b_{1}}$ is a solution of \eqref{eq:b} satisfying $%
b_{b_{1}}(-l)=b_{1}$. Note that we have $b_{b_{1}}(l)=b_{2}$ if and only if $%
(b_{1},b_{2})\in \operatorname{Graph}P^{\dag }$. And the second map $j$ is given by
\begin{equation*}
j(\eta _{-},b,\eta _{+})=\left( \eta _{-}+\eta _{b(-l)},\eta _{+}-\eta
_{b(l)}\right) .
\end{equation*}%
and so $\operatorname{ker}E^{\dagger }(u)$ has its dimension given by
\begin{eqnarray}
&&\mathop{\kern0pt{\rm dim}}\nolimits\operatorname{ker}\left( D_{u_{+}}{\overline{%
\partial }}_{(K_{+},J_{+})}\right) ^{\dagger }+\mathop{\kern0pt{\rm dim}}%
\nolimits\operatorname{ker}\left( D_{u_{-}}{\overline{\partial }}_{(K_{-},J_{-})}%
\right) ^{\dagger }+\mathop{\kern0pt{\rm dim}}\nolimits\operatorname{Graph}%
P^{\dagger }\cap \left( V_{+}^{\perp }\times V_{-}^{\perp }\right) \notag \\
&=&\mathop{\kern0pt{\rm dim}}\nolimits\operatorname{ker}\left( D_{u_{+}}({\overline{%
\partial }}_{(K_{+},J_{+})}\right) ^{\dagger }+\mathop{\kern0pt{\rm dim}}%
\nolimits\operatorname{ker}\left( D_{u_{-}}{\overline{\partial }}_{(K_{-},J_{-})}%
\right) ^{\dagger }+\mathop{\kern0pt{\rm dim}}\nolimits\left( P^{\dagger
}\cdot V_{-}^{\perp }\cap V_{+}^{\perp }\right) \label{dim:dfd_coker}
\end{eqnarray}%
Equivalently $E(u)$ has a closed range and its $\operatorname{coker}E(u)$ has
dimension the same as this. Combining this dimension counting of $\operatorname{coker%
}E(u)$ with that of $\operatorname{ker}E(u)$ in $\left( \ref{kerdim:dfd}\right) $,
we conclude that $E(u)$ is Fredholm and has index given by
\begin{eqnarray*}
\operatorname{Index}E(u) &=&\left( \Big(\mathop{\kern0pt{\rm dim}}\nolimits\ker
D_{u_{+}}{\overline{\partial }}_{(K^{+},J^{+})}+\mathop{\kern0pt{\rm dim}}%
\nolimits\ker D_{u_{-}}{\overline{\partial }}_{(K^{-},J^{-})}-%
\mathop{\kern0pt{\rm dim}}\nolimits(P\cdot V_{-}+V_{+})\right) \\
&{}&-\left( \Big(\mathop{\kern0pt{\rm dim}}\nolimits\ker (D_{u_{+}}{%
\overline{\partial }}_{(K^{+},J^{+})})^{\dagger }{}+\mathop{\kern0pt{\rm
dim}}\nolimits\ker (D_{u_{-}}{\overline{\partial }}_{(K^{-},J^{-})})^{%
\dagger }+\mathop{\kern0pt{\rm dim}}\nolimits\left( P^{\dagger }\cdot
V_{-}^{\perp }\cap V_{+}^{\perp }\right) )\right) \\
&=&\operatorname{Index}D_{u_{+}}{\overline{\partial }}_{(K^{+},J^{+})}+\operatorname{Index}%
D_{u_{-}}{\overline{\partial }}_{(K^{-},J^{-})}-2n \\
&=&(n+\mu _{H_{-}}([z_{-},w_{-}])+2c_{1}(A_{-}))+(n-\mu
_{H_{+}}([z_{+},w_{+}])+2c_{1}(A_{+})) \\
&=&\mu _{H_{-}}([z_{-},w_{-}])-\mu
_{H_{+}}([z_{+},w_{+}])+c_{1}(A_{-})+c_{1}(A_{+}).
\end{eqnarray*}%
Here we have used
\begin{eqnarray*}
&&\mathop{\kern0pt{\rm dim}}\nolimits(P\cdot V_{-}+V_{+})+%
\mathop{\kern0pt{\rm dim}}\nolimits\left( P^{\dagger }\cdot V_{-}^{\perp
}\cap V_{+}^{\perp }\right) \\
&=&\mathop{\kern0pt{\rm dim}}\nolimits(P\cdot V_{-}+V_{+})+%
\mathop{\kern0pt{\rm dim}}\nolimits\left( \left( P\cdot V_{-}\right) ^{\perp
}\cap V_{+}^{\perp }\right) \\
&=&\mathop{\kern0pt{\rm dim}}\nolimits(P\cdot V_{-}+V_{+})+%
\mathop{\kern0pt{\rm dim}}\nolimits(P\cdot V_{-}+V_{+})^{\perp }= 2n,
\end{eqnarray*}%
for the second identity and
\begin{eqnarray*}
\operatorname{Index}D_{u_{-}}{\overline{\partial }}_{(K_{-},J_{-})} &=&(n+\mu
_{H_{-}}([z_{-},w_{-}])+2c_{1}(A_{-})) \\
\operatorname{Index}D_{u_{+}}{\overline{\partial }}_{(K_{+},J_{+})} &=&(n-\mu
_{H_{+}}([z_{+},w_{+}])+2c_{1}(A_{+}))
\end{eqnarray*}%
for the third identity. \
To justify that $P^{\dagger }\cdot V_{-}^{\perp }=\left( P\cdot V_{-}\right)
^{\perp }$: For convenience we let
\begin{eqnarray*}
P &:&T_{u_{-}\left( o_{-}\right) }M\rightarrow T_{u_{+}\left( o_{+}\right)
}M,\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ }a\left( -l\right) \rightarrow a\left( l\right) \\
P^{\dag } &:&T_{u_{-}\left( o_{-}\right) }M\rightarrow T_{u_{+}\left(
o_{+}\right) }M,\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ }b\left( -l\right) \rightarrow b\left( l\right)
\end{eqnarray*}%
be the linear maps for solutions $a$ and $b$ of ODE $\left( \ref{eq:a}%
\right) $ and $\left( \ref{eq:b}\right) $ respectively. Then for solutions $%
a,b$ we have \
\begin{equation*}
\frac{d}{d\tau }\left\langle a\left( \tau \right) ,b\left( \tau \right)
\right\rangle =\left\langle {\Greekmath 0272} \operatorname{grad}f\left( \chi \right) a\left(
\tau \right) ,b\left( \tau \right) \right\rangle +\left\langle a\left( \tau
\right) ,-{\Greekmath 0272} \operatorname{grad}f\left( \chi \right) b\left( \tau \right)
\right\rangle =0,
\end{equation*}%
This in particular implies
\begin{equation*}
\left\langle Pa,P^{\dag }b\right\rangle =\left\langle a\left( l\right)
,b\left( l\right) \right\rangle =\left\langle a\left( -l\right) ,b\left(
-l\right) \right\rangle =\left\langle a,b\right\rangle .
\end{equation*}%
and hence%
\begin{equation*}
P^{\dag }V_{-}^{\bot }=\left( PV_{-}\right) ^{\bot }.
\end{equation*}
\end{proof}
\begin{cor}
\label{achieve-dfd-trans}Suppose $u_{\pm }\in {\mathcal{M}}([z_{\pm },w_{\pm
}];A_{\pm })$ are Fredholm regular, then $u\in {\mathcal{M}}%
^{l}([z_{-},w_{-}];f;[z_{+},w_{+}];A_{\pm })$ is Fredholm regular (in the
sense that $E(u)$ is surjective) if and only if the configuration $u=\left(
u_{-},\chi ,u_{+}\right) $ satisfies the \textquotedblleft
disk-flow-disk\textquotedblright\ transversality in definition \ref%
{dfd-trans}.
\end{cor}
\begin{proof}
In $\left( \ref{dim:dfd_coker}\right) $ of the above proposition we have
obtained
\begin{eqnarray*}
\mathop{\kern0pt{\rm dim}}\nolimits \operatorname{ker}E^{\dagger }(u) &=&%
\mathop{\kern0pt{\rm dim}}\nolimits\operatorname{ker}\left( D_{u_{+}}({\overline{%
\partial }}_{(K_{+},J_{+})}\right) ^{\dagger }+\mathop{\kern0pt{\rm dim}}%
\nolimits\operatorname{ker}\left( D_{u_{-}}{\overline{\partial }}_{(K_{-},J_{-})}%
\right) ^{\dagger } \\
&{}& \quad +\mathop{\kern0pt{\rm dim}}\nolimits\left( P^{\dagger }\cdot
V_{-}^{\perp }\cap V_{+}^{\perp }\right) \\
&=&\mathop{\kern0pt{\rm dim}}\nolimits\left( P^{\dagger }\cdot V_{-}^{\perp
}\cap V_{+}^{\perp }\right) \\
&=&\mathop{\kern0pt{\rm dim}}\nolimits\left( PV_{-}+V_{+}\right) ^{\perp }.
\end{eqnarray*}
Hence $E(u)$ is surjective if and only if $\left( PV_{-}+V_{+}\right)
^{\perp }=\left\{ 0\right\} ,$ i.e.
\begin{equation*}
PV_{-}+V_{+}=T_{u_{+}\left( o_{+}\right) }M.
\end{equation*}
But this is equivalent to%
\begin{equation*}
d\phi _{f}^{2l}ev_{-}\times ev_{+}\pitchfork \Delta _{u_{+}\left(
o_{+}\right) }^{tn},
\end{equation*}%
as derived in $\left( \ref{kerdim:dfd}\right) $.
\end{proof}
We then show $E(u)$ is surjective for generic $f$, for any $u=\left(
u_{-},\chi ,u_{+}\right) \in {\mathcal{M}}%
^{l}([z_{-},w_{-}];f;[z_{+},w_{+}];A_{\pm })$ of \emph{any} $l>0$. We need
the variation of $l$ to get this surjectivity. More precisely we have the
following
\begin{prop}
\label{family-dfd-trans}Suppose that $u_{\pm }\in {\mathcal{M}}(K_{\pm
},J_{\pm };[z_{\pm },w_{+}];A_{\pm })$ is Fredholm regular, i.e., its
linearization is surjective. Then for each given $l_{0}>0$, there exists a
dense subset of $f\in C^{\infty }\left( M\right) $ such that any element $%
u=\left( u_{-},\chi ,u_{+},l\right) $ in
\begin{equation*}
{\mathcal{M}}^{para}([z_{-},w_{-}];f;[z_{+},w_{+}];A_{\pm })=\bigcup_{l\geq
l_{0}}{\mathcal{M}}^{l}([z_{-},w_{-}];f;[z_{+},w_{+}];A_{\pm })
\end{equation*}%
is Fredholm regular, in the sense that $E(u)$ is surjective. ${\mathcal{M}}%
^{para}([z_{-},w_{-}];f;[z_{+},w_{+}];A_{\pm })$ is a smooth manifold with
dimension equal to the index of $E(u)$:
\begin{eqnarray*}
&{}& \operatorname{dim}{\mathcal{M}}^{para}([z_{-},w_{-}];f;[z_{+},w_{+}];A_{\pm })
\\
& = &\mu_{H_{-}}([z_{-},w_{-}])-\mu
_{H_{+}}([z_{+},w_{+}])+c_{1}(A_{-})+c_{1}(A_{+})+1.
\end{eqnarray*}
In particular when $A_- + A_+ = 0$ in $\pi_2(M)$, the index becomes
\begin{equation} \label{eq:A-A+=0}
\mu_{H_{-}}([z_{-},w_{-}])-\mu_{H_{+}}([z_{+},w_{+}])+1.
\end{equation}
\end{prop}
\begin{proof}
Consider the section $e$ of the following Banach bundle
\begin{eqnarray*}
e &:&{\mathcal{B}}^{dfd}(z_{-},z_{+})\times C^{\infty }\left( M\right)
\rightarrow {\mathcal{L}}^{dfd}(z_{-},z_{+}) \\
e &:&\left( u,f\right) \rightarrow ({\overline{\partial }}%
_{(K_{-},J_{-})}u_{-},\dot{\chi}-{\Greekmath 0272} f(\chi ),{\overline{\partial }}%
_{(K_{+},J_{+})}u_{+}).
\end{eqnarray*}%
where $u=\left( u_{-},\chi ,u_{+},l\right) \in $ ${\mathcal{B}}%
_{l}^{dfd}(z_{-},z_{+})\subset {\mathcal{B}}^{dfd}(z_{-},z_{+})$. Denote the
linearization of $e$ at $\left( u,f\right) $ by
\begin{equation*}
E(u,f):=D_{(u,f)}e:T_{u}{\mathcal{B}}^{dfd}(z_{-},z_{+})\times
T_{f}C^{\infty }\left( M\right) \rightarrow L_{u}^{p}(z_{-},z_{+}).
\end{equation*}%
For $\xi _{\pm }\in W^{1,p}(u_{\pm }^{\ast }TM)$, $a\in W^{1,p}(\chi ^{\ast
}TM)$, $\mu \in T_{l}\mathbb{R}$ and $h\in T_{f}C^{\infty }\left( M\right) $%
,
\begin{equation*}
E(u,f):\left( \xi _{-},a,\xi _{+},\mu ,h\right) \rightarrow \eta :=(\eta
_{-},b,\eta _{+})
\end{equation*}%
where
\begin{eqnarray*}
\eta _{-} &=&D_{u_{-}}{\overline{\partial }}_{(K_{-},J_{-})}\xi _{-}, \\
b &=&\frac{D}{d\tau }a-{\Greekmath 0272} _{a}\operatorname{grad}(f)+\frac{\mu }{l}\dot{\chi}%
\left( \tau \right) -{\Greekmath 0272} h(\chi ), \\
\eta _{+} &=&D_{u_{+}}{\overline{\partial }}_{(K_{+},J_{+})}\xi _{+}\,.
\end{eqnarray*}
Next we show the cokernel of $E(u,f)$ vanishes. Let $E(u,f)^{\ast }$ be the
adjoint operator of $E(u,f)$, such that
\begin{equation*}
E(u,f)^{\ast }:L_{u}^{p}(z_{-},z_{+})^{\ast }\rightarrow \left(
W^{1,p}(z_{-},z_{+};dfd)\times T_{f}C^{\infty }\left( M\right) \right)
^{\ast }.
\end{equation*}%
Using the nondegenerate $L^{2}$ pairing
\begin{equation*}
L^{p}(\Lambda ^{(0,1)}u^{\ast }TM)\times L^{q}(\Lambda ^{(1,0)}u^{\ast
}TM)\rightarrow {\mathbb{R}}
\end{equation*}%
we identify $L_{u}^{p}(z_{-},z_{+})^{\ast }$ with $L^{q}(\Lambda
^{(1,0)}u^{\ast }TM)$. On the other hand, we can identify $\left(
W^{1,p}(z_{-},z_{+};dfd)\times T_{f}C^{\infty }\left( M\right) \right)
^{\ast }$ with $W^{-1,q}(z_{-},z_{+};dfd)\times \left( T_{f}C^{\infty
}\left( M\right) \right) ^{\ast }$ where $W^{-1,q}(z_{-},z_{+};dfd)$ is
defined to be
\begin{equation*}
\left\{ (\xi _{-},a,\xi _{+},\mu )\in W^{1,p}(z_{-},z_{+})\mid \xi
_{-}(o_{-})=a(-l)-\mu \dot{\chi}(-l),\,\xi _{+}(o_{+})=a(l)+\mu \dot{\chi}%
(l)\right\} ^{\perp }
\end{equation*}%
in the direct product
\begin{equation*}
W^{-1,q}(z_{-},z_{+})=W^{-1,q}(u_{-}^{\ast }TM)\times W^{-1,q}(\chi ^{\ast
}TM)\times W^{-1,q}(u_{+}^{\ast }TM)
\end{equation*}%
where $(\cdot )^{\perp }$ denotes the $L^{2}$-orthogonal complement. Here we
have $1<q<2$ since $2<p<\infty $.
We denote by
\begin{equation*}
E(u,f)^{\dagger }:L^{q}(\Lambda ^{(1,0)}u^{\ast }TM)\rightarrow
W^{-1,q}(z_{-},z_{+};dfd)\times \left( T_{f}C^{\infty }\left( M\right)
\right) ^{\ast }
\end{equation*}%
the corresponding $L^{2}$-adjoint with respect to these identifications.
Recall by definition, we have
\begin{equation*}
\langle E(u,f)\xi ,\eta \rangle =\langle \xi ,E(u,f)^{\dagger }\eta \rangle .
\end{equation*}%
For any given $\eta :=(\eta _{-},b,\eta _{+})\in \operatorname{ker}E^{\dagger
}(u,f)\subset L^{q}(\Lambda ^{(1,0)}u^{\ast }TM),$ it satisfies
\begin{equation*}
\langle E(u,f)\xi ,\eta \rangle =0
\end{equation*}
for all $(\xi _{+},a,\xi _{-},\mu )\in T_{u}{\mathcal{B}}^{dfd}(z_{-},z_{+})$%
, especially for $(\xi _{+},a,\xi _{-},0)\in T_{u}{\mathcal{B}}%
_{l}^{dfd}(z_{-},z_{+}),$ namely
\begin{eqnarray}
0 &=&\int_{-l}^{l}\left\langle \frac{Da}{\partial \tau }-{\Greekmath 0272} \operatorname{grad}%
f(\chi )a,b\right\rangle -\int_{-l}^{l}\left\langle {\Greekmath 0272} h(\chi
),b\right\rangle \notag \\
&&{}+\int_{\dot{\Sigma}_{-}}\left\langle D_{u_{-}}{\overline{\partial }}%
_{(K^{-},J^{-})}\xi _{-},\eta _{-}\right\rangle +\int_{\dot{\Sigma}%
_{+}}\left\langle D_{u_{+}}{\overline{\partial }}_{(K_{+},J_{+})}\xi
_{+},\eta _{+}\right\rangle \label{dfd-coker-h}
\end{eqnarray}%
for all the triples satisfying the matching condition
\begin{equation}
\xi _{-}(o_{-})=a(-l),\quad \xi _{+}(o_{+})=a(l). \label{match2}
\end{equation}%
Letting $a=\xi _{-}=\xi _{+}=0$, then $\left( \ref{dfd-coker-h}\right) $
becomes
\begin{equation*}
\int_{-l}^{l}\left\langle {\Greekmath 0272} h(\chi ),b\right\rangle =0
\end{equation*}%
for all $h\in C^{\infty }\left( M\right) $, so $b=0$. Now $\left( \ref%
{dfd-coker-h}\right) $ becomes%
\begin{equation*}
\int_{\dot{\Sigma}_{-}}\left\langle D_{u_{-}}{\overline{\partial }}%
_{(K^{-},J^{-})}\xi _{-},\eta _{-}\right\rangle +\int_{\dot{\Sigma}%
_{+}}\left\langle D_{u_{+}}{\overline{\partial }}_{(K_{+},J_{+})}\xi
_{+},\eta _{+}\right\rangle =0.
\end{equation*}%
Notice that $\xi _{-}$ and $\xi _{+}$ can vary independently since $a\left(
-l\right) $ and $a\left( l\right) $ can be any vector without restriction
and hence the matching condition $\left( \ref{match2}\right) $ does not put
any restriction on $\xi _{\pm }\left( o_{\pm }\right) $. Therefore we have%
\begin{eqnarray*}
\int_{\dot{\Sigma}_{-}}\left\langle D_{u_{-}}{\overline{\partial }}%
_{(K^{-},J^{-})}\xi _{-},\eta _{-}\right\rangle &=&0, \\
\int_{\dot{\Sigma}_{+}}\left\langle D_{u_{+}}{\overline{\partial }}%
_{(K_{+},J_{+})}\xi _{+},\eta _{+}\right\rangle &=&0
\end{eqnarray*}%
for \emph{all} $\xi _{\pm }$. In other words, $\eta _{\pm }$ lie in coker$%
D_{u_{\pm }}{\overline{\partial }}_{(K_{\pm },J_{\pm })}$. By the
hypothesis, we conclude $\eta _{\pm }=0$. Therefore the cokernel of $E(u,f)$
vanishes in $E(u,f):W^{1,p}\rightarrow L^{p}$ setting.
Now we raise the regularity to $W^{k,p}$ setting. For any $k\geq 1$ and $%
\zeta \in W^{k-1,p}\subset L^{p}$, from vanishing of the above cokernal we
can always solve $E(u,f)\xi =\zeta $ with $\xi \in W^{1,p}$. By ellipticity
of $E(u,f)$ we have $\xi \in W^{k,p}$. Therefore, the cokernal of $E(u,f)$
vanishes if we treat $E(u,f)$ as a map from $W^{k,p}$ to $W^{k-1,p}$ for any
$k$. We fix a large enough $k$ (which depends on $n,c_{1}(A_{\pm }),\mu
_{H_{\pm }}([z_{\pm },w_{\pm }]$) to define the $W^{k,p}$ Banach norm on ${%
\mathcal{B}}^{dfd}(z_{-},z_{+})$. Therefore, the universal \ moduli space
\begin{equation*}
{\mathcal{M}}_{univ}^{para}([z_{-},w_{-}];f;[z_{+},w_{+}];A_{\pm
})=e^{-1}\left( 0\right)
\end{equation*}%
is a $W^{k,p}$ Banach manifold in ${\mathcal{B}}^{dfd}(z_{-},z_{+})\times
C^{\infty }\left( M\right) $. For the natural projection
\begin{equation*}
\pi :{\mathcal{M}}_{univ}^{para}([z_{-},w_{-}];f;[z_{+},w_{+}];A_{\pm
})\rightarrow C^{\infty }\left( M\right) ,
\end{equation*}%
since our $k$ is large enough, we can apply Sard-Smale theorem and conclude
its regular values $f$ form a set of second category in $C^{\infty }\left(
M\right) $. For any regular value $f$,
\begin{equation*}
{\mathcal{M}}^{para}([z_{-},w_{-}];f;[z_{+},w_{+}];A_{\pm })=\pi ^{-1}\left(
f\right) \cap {\mathcal{M}}_{univ}^{para}([z_{-},w_{-}];f;[z_{+},w_{+}];A_{%
\pm })
\end{equation*}%
is a finite dimensional smooth submanifold with
\begin{eqnarray*}
&&\mathop{\kern0pt{\rm dim}}\nolimits{\mathcal{M}}%
^{para}([z_{-},w_{-}];f;[z_{+},w_{+}];A_{\pm }) \\
&=&\RIfM@\expandafter\text@\else\expandafter\mbox\fi{index}(\pi ) = \RIfM@\expandafter\text@\else\expandafter\mbox\fi{index}(E\left( u\right) ) \\
&=&\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ }\mu _{H_{-}}([z_{-},w_{-}])-\mu
_{H_{+}}([z_{+},w_{+}])+c_{1}(A_{-})+c_{1}(A_{+})+1.
\end{eqnarray*}%
The index$E\left( u\right) $ here differs from previous Proposition by $1$
because the linearization $E\left( u\right) $ here is in $T_{u}{\mathcal{B}}%
^{dfd}(z_{-},z_{+})$ instead of $T_{u}{\mathcal{B}}_{l}^{dfd}(z_{-},z_{+})$.
Since $f$ is a regular value of $\pi $, $E\left( u\right) :T_{u}{\mathcal{B}}%
^{dfd}(z_{-},z_{+})\rightarrow \mathcal{L}^{dfd}(z_{-},z_{+})$ is surjective
and $u$ is Fredholm regular.
\end{proof}
Next we establish the condition to ensure the joint points $u_{\pm }\left(
o_{\pm }\right) $ are immersed. This condition satisfies for a generic
choice of almost complex structures $J$ and will be needed in the proof of
the surjectivity of our gluing: The Hausdorff convergence imposed in
Definition \ref{defn:adiabatic} (2) is not strong enough to detect multiple
covers of the thin cylinder although it captures all simple thin cylinders.
Immersion condition at the joint points then make the thin part of the
adiabatic limit with immersed joint points be automatically simple.
\begin{prop}
\label{immersed-joint}For generic $J_{0}\in \mathcal{J}_{\omega }$ and $f\in
C^{\infty }\left( M\right) $, any element $u=\left( u_{-},\chi
,u_{+},l\right) \in {\mathcal{M}}^{para}([z_{-},w_{-}];f;[z_{+},w_{+}];A_{%
\pm })$ whose $u_{\pm }$ are somewhere injective must be Fredholm regular,
and have both $u_{\pm }\left( o_{\pm }\right) $ immersed, provided that
\begin{equation*}
\mu _{H_{-}}([z_{-},w_{-}])-\mu
_{H_{+}}([z_{+},w_{+}])+2c_{1}(A_{-})+2c_{1}(A_{+})<2n-1,
\end{equation*}%
or equivalently
\begin{equation*}
\mathop{\kern0pt{\rm dim}}\nolimits{\mathcal{M}}%
^{para}([z_{-},w_{-}];f;[z_{+},w_{+}];A_{\pm })\leq 2n-1.
\end{equation*}%
Especially, for any ${\mathcal{M}}^{para}([z_{-},w_{-}];f;[z_{+},w_{+}];A_{%
\pm })$ of virtual dimension $0,1$, both $u_{\pm }\left( o_{\pm }\right) $
are immersed for generic $\left( J_{0},f\right) $.
\end{prop}
\begin{proof}
The proof is a small variant of Theorem 1.2 of \cite{oh-zhu2}. Consider the
section $\Upsilon $ of the following Banach bundle%
\begin{eqnarray*}
\Upsilon &:&{\mathcal{B}}^{dfd}(z_{-},z_{+})\times C^{\infty }\left(
M\right) \times \mathcal{J}_{\omega }\rightarrow {\mathcal{L}}%
^{dfd}(z_{-},z_{+})\times H_{-}^{1,0}\times H_{+}^{1,0}, \\
\Upsilon &:&\left( u,f,J_{0}\right) \rightarrow \left( {\overline{\partial }}%
_{(K_{-},J_{-})}u_{-},\dot{\chi}-{\Greekmath 0272} f(\chi ),{\overline{\partial }}%
_{(K_{+},J_{+})}u_{+},{\partial }_{J_{0}}u_{-}\left( o_{-}\right) ,{\partial
}_{J_{0}}u_{+}\left( o_{+}\right) \right)
\end{eqnarray*}%
where
\begin{equation*}
H_{\pm }^{0,1}\left( u_{\pm }\right) =Hom_{i,J_{0}}\left( T_{o_{\pm }}\Sigma
_{\pm },T_{u_{\pm }\left( o_{\pm }\right) }M\right)
\end{equation*}%
is a rank $2n$ vector bundle over ${\mathcal{B}}^{dfd}(z_{-},z_{+})\times
C^{\infty }\left( M\right) \times \mathcal{J}_{\omega }$ whose fiber
consists of $\left( j,J_{0}\right) $-linear maps from $T_{o_{\pm }}\Sigma
_{\pm }$ to $T_{u_{\pm }\left( o_{\pm }\right) }M$. Here the $j$ for $\Sigma
_{\pm }\simeq \left( D^{2},o_{\pm }\right) $ is fixed. Any cokernal element
\begin{equation*}
\left( \eta _{-},b,\eta _{+},\alpha _{-},\alpha _{+}\right) \in {\mathcal{L}}%
^{dfd}(z_{-},z_{+})\times H_{-}^{1,0}\times H_{+}^{1,0}
\end{equation*}%
of $D_{\left( u,f,J_{0}\right) }\Upsilon $ must satisfy%
\begin{eqnarray}
0 &=&\int_{-l}^{l}\left\langle \frac{Da}{\partial \tau }-{\Greekmath 0272} \operatorname{grad}%
f(\chi )a,b\right\rangle -\int_{-l}^{l}\left\langle {\Greekmath 0272} h(\chi
),b\right\rangle \notag \\
&&{}+\int_{\dot{\Sigma}_{-}}\left\langle D_{u_{-}}{\overline{\partial }}%
_{(K^{-},J^{-})}\xi _{-}+\frac{1}{2}du_{-}\circ B\circ j,\eta
_{-}\right\rangle \notag \\
&&+\int_{\dot{\Sigma}_{+}}\left\langle D_{u_{+}}{\overline{\partial }}%
_{(K_{+},J_{+})}\xi _{+}\frac{1}{2}du_{+}\circ B\circ j,\eta
_{+}\right\rangle \notag \\
&&+\left\langle D_{u_{-}}{\partial }_{(K^{-},J^{-})}\xi _{-},\alpha
_{-}\delta _{o_{-}}\right\rangle _{\dot{\Sigma}_{-}}+\left\langle D_{u_{+}}{%
\partial }_{(K_{+},J_{+})}\xi _{+},\alpha _{+}\delta _{o_{+}}\right\rangle _{%
\dot{\Sigma}_{+}} \label{coker-jt}
\end{eqnarray}%
for all $(\xi _{+},a,\xi _{-},\mu )\in T_{u}{\mathcal{B}}^{dfd}(z_{-},z_{+})$%
, $h\in T_{f}C^{\infty }\left( M\right) $ and $B\in T_{J_{0}}\mathcal{J}%
_{\omega }$, where $\delta _{o_{\pm }}$ are delta functions at $o_{\pm }$ on
$\dot{\Sigma}_{\pm }$.
Letting $a=\xi _{-}=\xi _{+}=B=0$, then $\left( \ref{coker-jt}\right) $
becomes
\begin{equation*}
\int_{-l}^{l}\left\langle {\Greekmath 0272} h(\chi ),b\right\rangle =0
\end{equation*}%
for all $h\in C^{\infty }\left( M\right) $, so $b=0$. Now $\left( \ref%
{coker-jt}\right) $ becomes%
\begin{eqnarray*}
0 &=&\int_{\dot{\Sigma}_{-}}\left\langle D_{u_{-}}{\overline{\partial }}%
_{(K^{-},J^{-})}\xi _{-}+\frac{1}{2}du_{-}\circ B\circ j,\eta
_{-}\right\rangle +\left\langle D_{u_{-}}{\partial }_{(K^{-},J^{-})}\xi
_{-},\alpha _{-}\delta _{o_{-}}\right\rangle _{\dot{\Sigma}_{-}} \\
&&+\int_{\dot{\Sigma}_{+}}\left\langle D_{u_{+}}{\overline{\partial }}%
_{(K_{+},J_{+})}\xi _{+}+\frac{1}{2}du_{+}\circ B\circ j,\eta
_{+}\right\rangle +\left\langle D_{u_{+}}{\partial }_{(K_{+},J_{+})}\xi
_{+},\alpha _{+}\delta _{o_{+}}\right\rangle _{\dot{\Sigma}_{+}}.
\end{eqnarray*}%
Notice that $\xi _{-}$ and $\xi _{+}$ can vary independently since $a\left(
-l\right) $ and $a\left( l\right) $ can be any vector without restriction
and hence the matching condition $\left( \ref{match2}\right) $ does not put
any restriction on $\xi _{\pm }\left( o_{\pm }\right) $. Therefore we have%
\begin{eqnarray*}
\int_{\dot{\Sigma}_{-}}\left\langle D_{u_{-}}{\overline{\partial }}%
_{(K^{-},J^{-})}\xi _{-}+\frac{1}{2}du_{-}\circ B\circ j,\eta
_{-}\right\rangle +\left\langle D_{u_{-}}{\partial }_{(K^{-},J^{-})}\xi
_{-},\alpha _{-}\delta _{o_{-}}\right\rangle _{\dot{\Sigma}_{-}} &=&0, \\
\int_{\dot{\Sigma}_{+}}\left\langle D_{u_{+}}{\overline{\partial }}%
_{(K_{+},J_{+})}\xi _{+}+\frac{1}{2}du_{+}\circ B\circ j,\eta
_{+}\right\rangle +\left\langle D_{u_{+}}{\partial }_{(K_{+},J_{+})}\xi
_{+},\alpha _{+}\delta _{o_{+}}\right\rangle _{\dot{\Sigma}_{+}} &=&0,
\end{eqnarray*}%
which are identical to $\left( 2.12\right) $ in \cite{oh-zhu2}. Then the
remaining steps are the same as in \cite{oh-zhu2} to show $\eta _{\pm }=0$
and $\alpha _{\pm }=0$, hence $\Upsilon \,$\ is transversal to the section
zero sections
\begin{eqnarray*}
&&o_{{\mathcal{L}}^{dfd}(z_{-},z_{+})}\times o_{H_{-}^{1,0}}\times
H_{+}^{1,0}\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ and} \\
&&o_{{\mathcal{L}}^{dfd}(z_{-},z_{+})}\times H_{-}^{1,0}\times
o_{H_{-}^{1,0}},
\end{eqnarray*}
which both have codimension $2n$ in $o_{{\mathcal{L}}^{dfd}(z_{-},z_{+})}%
\times H_{-}^{1,0}\times H_{+}^{1,0}$. Hence for generic $\left(
J_{0},f\right) $, by Sard Theorem as in \cite{oh-zhu2},
\begin{equation*}
\Upsilon \,^{-1}\left( o_{{\mathcal{L}}^{dfd}(z_{-},z_{+})}\times
o_{H_{-}^{1,0}}\times H_{+}^{1,0}\cup o_{{\mathcal{L}}^{dfd}(z_{-},z_{+})}%
\times H_{-}^{1,0}\times o_{H_{-}^{1,0}}\right)
\end{equation*}%
has dimension $2n$ less than
\begin{equation*}
\Upsilon \,^{-1}\left( o_{{\mathcal{L}}^{dfd}(z_{-},z_{+})}\times
H_{-}^{1,0}\times H_{+}^{1,0}\right) ={\mathcal{M}}%
^{para}([z_{-},w_{-}];f;[z_{+},w_{+}];A_{\pm }),
\end{equation*}
so is of dimension
\begin{equation*}
\mu _{H_{-}}([z_{-},w_{-}])-\mu
_{H_{+}}([z_{+},w_{+}])+2c_{1}(A_{-})+2c_{1}(A_{+})+1-2n,
\end{equation*}%
which is negative if the index condition of the proposition holds. But this
means the set of the $\left( u_{-},\chi ,u_{+}\right) $ who fails the
immersion condition at $o_{-}$ or $o_{+}$ is empty. The proposition is
proved.
\end{proof}
\section{$J$-holomorphic curves from cylindrical domain}
The (perturbed) $J$-holomorphic curves from punctured Riemannian surface $%
\dot{\Sigma}$ have been extensively studied in the literature. Since we will
use cylindrical coordinates of punctured Riemannian surface in our gluing
analysis , we briefly review the Banach norms in this setting. For our $%
u_{\pm }$, the domain $\left( D^{2},o_{\pm }\right) $ can be thought as
\begin{equation*}
\Sigma _{\pm }=\left( D^{2},o_{\pm }\right) \simeq \lbrack -\infty ,\infty
]\times S^{1},
\end{equation*}%
where $o_{\pm }\simeq \left\{ -\pm \infty \right\} \times S^{1}$ and $%
\partial D^{2}\simeq \left\{ \pm \infty \right\} \times S^{1}$. Let $O_{\pm
\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ }}\simeq -[\pm \infty ,0)\times S^{1}$ be the cylindrical
neighborhood near the puncture $o_{\pm }=\left\{ -\pm \infty \right\} \times
S^{1}$. Without loss of generality let's assume the image of $O_{\pm }$
under $u_{\pm }\ $has diameter less than the injective radius of $\left(
M,g\right) $ (otherwise we can shift the ${\mathbb{R}}$ component to
reparametrize $u_{\pm }$)$.$ One can use the following norm to define the
Banach manifold hosting such Floer trajectories. Let%
\begin{equation*}
\mathcal{B}_{+}\left( z_{\pm }\right) =\left\{ u_{\pm }:\Sigma _{\pm
}\rightarrow M\in W_{loc}^{1,p},\left\vert
\begin{array}{c}
u_{+}\left( -\infty ,t\right) =p_{\pm }\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ for some }p_{+}\in M\RIfM@\expandafter\text@\else\expandafter\mbox\fi{, }
\\
u_{+}\left( \infty ,t\right) =z_{+}\left( t\right) , \\
u_{+}\left( \tau ,t\right) =\exp _{p_{+}}\xi \left( \tau ,t\right) \RIfM@\expandafter\text@\else\expandafter\mbox\fi{
near }p_{+}\RIfM@\expandafter\text@\else\expandafter\mbox\fi{, } \\
u_{+}\left( \tau ,t\right) =\exp _{z_{+}\left( t\right) }\xi \left( \tau
,t\right) \RIfM@\expandafter\text@\else\expandafter\mbox\fi{ near }z_{+}\RIfM@\expandafter\text@\else\expandafter\mbox\fi{,} \\
\RIfM@\expandafter\text@\else\expandafter\mbox\fi{and }e^{\frac{2\pi \delta \left\vert \tau \right\vert }{p}}\xi \in
W^{1,p}\left( {\mathbb{R}}\times S^{1}\right) .%
\end{array}%
\right. \right\} .
\end{equation*}%
We define $\mathcal{B}_{-}$ similarly with $\infty $ and $-\infty $ switched
around. For any section $\xi _{\pm }\in W_{loc}^{1,p}\left( u_{\pm }^{\ast
}TM\right) $, we define its Banach norm to be%
\begin{equation*}
\left\Vert \xi _{\pm }\right\Vert _{W_{\alpha }^{1,p}\left( \Sigma _{\pm
}\right) }:=\left\Vert \widetilde{\xi _{\pm }}\right\Vert _{W_{\delta
}^{1,p}\left( \Sigma _{\pm }\right) }+\left\vert \xi _{\pm }\left( o_{\pm
}\right) \right\vert ,
\end{equation*}%
where $\widetilde{\xi _{\pm }}$ is defined by
\begin{equation*}
\widetilde{\xi _{\pm }}=\xi _{\pm }-\varphi _{\pm }\left( \tau ,t\right)
Pal_{\pm }\left( \tau ,t\right) \left( \xi _{\pm }\left( o_{\pm }\right)
\right)
\end{equation*}%
and
\begin{equation*}
\left\Vert \widetilde{\xi _{\pm }}\right\Vert _{W_{\delta }^{1,p}\left(
\Sigma _{\pm }\right) }:=\left\Vert e^{\frac{2\pi \delta \left\vert \tau
\right\vert }{p}}\widetilde{\xi _{\pm }}\right\Vert _{W^{1,p}\left( {\mathbb{%
R}}\times S^{1}\right) }.
\end{equation*}%
Here $\varphi _{\pm }\left( \tau ,t\right) $ is a smooth cut-off function
such that
\begin{equation*}
\varphi _{\pm }\left( \tau ,t\right) =%
\begin{cases}
1\quad & \mbox{ on }\,\pm \lbrack -\infty ,-1)\times S^{1} \\
0\quad & \mbox{ outside the neighborhood of }O_{\pm }%
\end{cases}%
\end{equation*}%
and $\left\vert d\varphi _{\pm }\right\vert \leq 2$, and $Pal_{\pm }\left(
\tau ,t\right) $ is the parallel transport along the shortest geodesics from
$u_{\pm }\left( o_{\pm }\right) $ to $u_{\pm }\left( \tau ,t\right) $. This
defines the Banach manifold structure on $\mathcal{B}_{\pm }\left( z_{\pm
}\right) $. For any section $\eta _{\pm }\in \Gamma \left( u_{\pm }^{\ast
}TM\right) \otimes \Lambda ^{0,1}\left( \Sigma _{\pm }\right) $, we let
\begin{equation*}
\left\Vert \eta \right\Vert _{L_{\delta }^{p}\left( \Sigma _{\pm }\right)
}:=\left\Vert e^{\frac{2\pi \delta \left\vert \tau \right\vert }{p}}\eta
\right\Vert _{L^{p}\left( {\mathbb{R}}\times S^{1}\right) }
\end{equation*}%
and denote the set of all such $\eta $ with $\left\Vert \eta \right\Vert
_{L_{\delta }^{p}\left( \Sigma _{\pm }\right) }<\infty $ by $L_{\delta
}^{p}\left( \Gamma \left( u_{\pm }^{\ast }TM\right) \otimes \Lambda
^{0,1}\left( \Sigma _{\pm }\right) \right) $. Then we let
\begin{equation*}
\mathcal{L}_{\pm }\left( z_{\pm }\right) =\bigcup\limits_{u_{\pm }\in
\mathcal{B}_{\pm }\left( z_{\pm }\right) }L_{\delta }^{p}\left( \Gamma
\left( u_{\pm }^{\ast }TM\right) \otimes \Lambda ^{0,1}\left( \Sigma _{\pm
}\right) \right)
\end{equation*}%
be the Banach bundle over $\mathcal{B}_{\pm }\left( z_{\pm }\right) $. It is
well-known that
\begin{equation*}
\overline{\partial }_{\left( K_{\pm },J_{\pm }\right) }:\mathcal{B}_{\pm
}\left( z_{\pm }\right) \rightarrow \mathcal{L}_{\pm }\left( z_{\pm }\right)
\end{equation*}%
is a smooth Fredholm section, and for generic $\left( K_{\pm },J_{\pm
}\right) $ the Floer trajectory $u_{\pm }\in \left( \overline{\partial }%
_{\left( K_{\pm },J_{\pm }\right) }\right) ^{-1}\left( 0\right) $ are
Fredholm regular and hence
\begin{equation*}
D_{u_{\pm }}\overline{\partial }_{\left( K_{\pm },J_{\pm }\right) }:T_{u\pm }%
\mathcal{B}_{\pm }\left( z_{\pm }\right) \simeq W_{\delta }^{1,p}\left(
\Gamma \left( u_{\pm }^{\ast }TM\right) \right) \oplus T_{p_{\pm
}}M\rightarrow L_{\delta }^{p}\left( \Gamma \left( u_{\pm }^{\ast }TM\right)
\otimes \Lambda ^{0,1}\left( \Sigma _{\pm }\right) \right)
\end{equation*}%
is surjective and has bounded right inverse, denoted by $Q_{\pm }$. Here we
recall
\begin{equation*}
T_{u_{\pm }}\mathcal{B}_{\pm }\left( z_{\pm }\right) \simeq W_{\delta
}^{1,p}\left( \Gamma \left( u_{\pm }^{\ast }TM\right) \right) \oplus
T_{p_{\pm }}M.
\end{equation*}%
For the simplicity of notations in later calculation, we denote by
\begin{equation}
v_{\pm }:=\varphi _{\pm }\left( \tau ,t\right) Pal_{\pm }\left( \tau
,t\right) \left( \xi _{\pm }\left( o_{\pm }\right) \right) \label{Vpm}
\end{equation}%
the cut-off \emph{constant vector field} extending the vector $\xi _{\pm
}\left( o_{\pm }\right) $.
\begin{rem}
With obvious modifications, we can use $\mathcal{B}_{\pm }\left( z_{\pm
}\right) $ (using cylindrical measure) instead of $W^{1,p}\left( \dot{\Sigma}%
_{\pm },M,z_{\pm }\right) $ (using compact measure ) in last section to
define the Banach manifold ${\mathcal{B}}^{dfd}(z_{-},z_{+})$. The index of $%
E(u)$ is the same, and $E(u)$ is surjective for generic $\left( J,f\right) $%
. Proofs are similar so omitted.
\end{rem}
\section{Approximate solutions}
In this section, we construct an approximate solution $u_{app}^{\varepsilon
} $ of the Floer equation. For the notational simplicity, we denote
\begin{equation} \label{eq:tau(e)}
\tau(\varepsilon) = \frac{l}{\varepsilon} + \frac{p-1}{\delta}S(\varepsilon)
\end{equation}
which will appear very often in the discussion henceforth. We also denote
the translation map
\begin{equation} \label{eq:Itaue}
I_{\tau_0}: {\mathbb{R}} \times S^1 \to {\mathbb{R}} \times S^1, \quad
I_{\tau_0}(\tau)(\tau,t) = (\tau-\tau_0,t)
\end{equation}
for $\tau_0 \in {\mathbb{R}}$. To simplify the notations, we also denote
\begin{equation} \label{eq:E(x,v)}
E(x,y) = (\exp_x)^{-1}(y)
\end{equation}
whenever $d(x,y)< \iota_M$ where $\iota_M$ is the injectivity radius of $%
(M,g)$.
We recall the decomposition of ${\mathbb{R}}$ into
\begin{equation*}
-\infty < -\tau(\varepsilon) - 1 < -\tau(\varepsilon) < -R(\varepsilon) <
R(\varepsilon) < \tau(\varepsilon) < \tau(\varepsilon) + 1 < \infty
\end{equation*}
we made from the beginning of section 3 where we made a choice
\begin{equation*}
R(\varepsilon) = \frac{l}{\varepsilon}, \quad \tau(\varepsilon) =
R(\varepsilon) + \frac{1}{2\pi}\ln\left(1+\frac{l}{\varepsilon}\right).
\end{equation*}
And we denoted $K_\pm(\tau,t,x) = \kappa^\pm(\tau)H_t(x)$. Note that this
latter representation of $K_\pm$ depend on the choice of analytic
coordinates $(\tau,t)$ compatible to the parameter $t$ parameterizing $H_t$
near the punctures respectively. The coordinates are unique modulo
translations by $\tau$.
Now let $u_\pm$ be solutions of the equations $(du +
P_{K_\pm})_{J_\pm}^{(0,1)}=0$ respectively and fix the coordinate
representations of $u_\pm = u_\pm(\tau,t)$ so that they are compatible with
the choice of analytic coordinates given at the punctures. This can be
always done by adjusting the choice of analytic coordinates near the two
punctures $(e_+,o_+)$ and $(e_-,o_-)$ respectively.
With this preparation, we define our approximate solution by
\begin{equation}
u_{app}^{\varepsilon }\left(\tau ,t\right) =%
\begin{cases}
u_{-}(\tau +\tau (\varepsilon ),t) & -\infty <\tau \leq -l/\varepsilon -1 \\
\exp _{\chi (\varepsilon \tau )}\left[ \left( 1-\kappa _{\varepsilon
}^{0}\left( \tau \right) \right) E(\chi (\varepsilon \tau ),
u_{-}^{\varepsilon }\left( \tau ,t\right) ) \right] & -l/\varepsilon -1\leq
\tau \leq -l/\varepsilon \\
\chi \left( \varepsilon \tau \right) & -l/\varepsilon \leq \tau \leq
l/\varepsilon , \\
\exp _{\chi (\varepsilon \tau )}\left[ \left( 1-\kappa _{\varepsilon
}^{0}\left( \tau \right) \right) E(\chi (\varepsilon \tau ),
u_{+}^{\varepsilon }\left( \tau ,t\right) ) \right] & l/\varepsilon \leq
\tau \leq l/\varepsilon +1 \\
u_{+}(\tau -\tau (\varepsilon ),t) & l/\varepsilon +1\leq \tau <\infty%
\end{cases}
\label{eq:uappe}
\end{equation}%
where
\begin{eqnarray*}
u_{-}^{\varepsilon }\left( \tau ,t\right) &=&u_{-}(\tau +\tau (\varepsilon
),t), \\
u_{+}^{\varepsilon }\left( \tau ,t\right) &=&u_{+}(\tau -\tau (\varepsilon
),t), \\
S\left( \varepsilon \right) &=&\frac{1}{2\pi }\ln \left( 1+l/\varepsilon
\right) ,~
\end{eqnarray*}%
and $\kappa _{\varepsilon }^{0}\left( \tau \right) $ is a smooth cut-off
functions defined in $\left( \ref{eq:beta0}\right) $, $\kappa _{\varepsilon
}^{0}\left( \tau \right) =1$ when $\left\vert \tau \right\vert \leq
l/\varepsilon $ and $\kappa _{\varepsilon }^{0}\left( \tau \right) =0$ when $%
\left\vert \tau \right\vert \geq l/\varepsilon +1$.
The assignment of the approximate solution to each $(u_{-},\chi ,u_{+})$ and
$0<\varepsilon <\varepsilon _{0}$ defines a smooth map
\begin{eqnarray}
&{}&\RIfM@\expandafter\text@\else\expandafter\mbox\fi{preG}:{\mathcal{M}}_{1}(K_{-},z_{-};A_{-}){}_{ev_{+}}\times
_{ev_{0}}{\mathcal{M}}(f;[0,\ell ]){}_{ev_{\ell }}\times _{ev_{-}}{\mathcal{M%
}}_{1}(K_{+},z_{+};A_{+})\times (0,\varepsilon _{0}] \notag
\label{eq:preglue} \\
&{}&\hskip1.5in\rightarrow \bigcup_{0<\varepsilon \leq \varepsilon _{0}}{%
\mathcal{M}}^{\varepsilon }(K_{\varepsilon };z_{-},z_{+};B).
\end{eqnarray}
Later in carrying out the $\overline{\partial }_{\left( K_{\varepsilon
},J_{\varepsilon };\varepsilon f\right) }$-error estimate, we make some
simplification of the expression of $u_{app}^{\varepsilon }$. Notice that
the interpolation in $u_{app}^{\varepsilon }$ takes place \emph{in a ball of
radius }$C\varepsilon $ around $p_{\pm }$, since
\begin{eqnarray}
\sup_{l/\varepsilon \leq \left\vert \tau \right\vert \leq l/\varepsilon +1}%
\RIfM@\expandafter\text@\else\expandafter\mbox\fi{dist}\left( \chi _{\varepsilon }\left( \tau \right) ,p_{\pm }\right)
&=&\sup_{l/\varepsilon \leq \left\vert \tau \right\vert \leq l/\varepsilon
+1}\RIfM@\expandafter\text@\else\expandafter\mbox\fi{dist}\left( \chi _{\varepsilon }\left( \tau \right) ,\chi
_{\varepsilon }\left( \pm l/\varepsilon \right) \right) \notag \\
&\leq &\varepsilon \sup \left\vert {\Greekmath 0272} f\right\vert \leq C\varepsilon
\label{dist-xp}
\end{eqnarray}%
and
\begin{eqnarray}
\sup_{l/\varepsilon \leq \left\vert \tau \right\vert \leq l/\varepsilon +1}%
\RIfM@\expandafter\text@\else\expandafter\mbox\fi{dist}\left( u_{\pm }^{\varepsilon }\left( \tau \right) ,p_{\pm
}\right) &=&\sup_{l/\varepsilon \leq \left\vert \tau \right\vert \leq
l/\varepsilon +1}\RIfM@\expandafter\text@\else\expandafter\mbox\fi{dist}\left( u_{\pm }^{\varepsilon }\left( \tau
\right) ,u_{\pm }^{\varepsilon }\left( \pm \infty \right) \right) \notag \\
&\leq &Ce^{-2\pi \frac{p-1}{\delta }S\left( \varepsilon \right) }=C\left(
1+l/\varepsilon \right) ^{-\frac{p-1}{\delta }} \label{dist-up} \\
&\leq &\widetilde{C}\left( l\right) \varepsilon ,
\end{eqnarray}%
where the first inequality is because we have
\begin{equation*}
\left\vert u_{\pm }^{\varepsilon }\left( \tau ,t\right) -u_{\pm
}^{\varepsilon }\left( \pm \infty ,t\right) \right\vert \leq Ce^{-2\pi
\left\vert \tau -\pm (\tau (\varepsilon ))\right\vert }
\end{equation*}%
by the $J$-holomorphic property of \ $u_{\pm }$ and the second inequality is
because we have chosen $\delta <p-1$ and set $\widetilde{C}\left( l\right)
=Cl^{-\frac{p-1}{\delta }}$.
At the interpolation $\left\vert \tau \right\vert \in \lbrack l/\varepsilon
,l/\varepsilon +1]$, both $u_{\pm }^{\varepsilon }\left( \tau ,t\right) $
and $\chi \left( \varepsilon \tau \right) $ are in the ball of radius $%
C\varepsilon $ around $p_{\pm }$. The expression
\begin{equation*}
\exp _{\chi \left( \varepsilon \tau \right) }\left[ \left( 1-\kappa
_{\varepsilon }^{0}\left( \tau \right) \right) E(\chi (\varepsilon \tau ),
u_{\pm}^{\varepsilon }\left( \tau ,t\right)) \right] =:
v_\varepsilon^\pm(\tau,t)
\end{equation*}%
in \eqref{eq:uappe} is $C^{1}$ close to $\chi (\varepsilon \tau )+\left(
1-\kappa _{\varepsilon }^{0}\left( \tau \right) \right) \left( u_{\pm
}^{\varepsilon }\left( \tau ,t\right) -\chi \left( \varepsilon \tau \right)
\right) $, i.e.
\begin{equation*}
\kappa _{\varepsilon }^{0}\left( \tau \right) \chi (\varepsilon \tau
)+\left( 1-\kappa _{\varepsilon }^{0}\left( \tau \right) \right) u_{\pm
}^{\varepsilon }\left( \tau ,t\right) =: \widetilde v_\varepsilon^\pm(\tau,t)
\end{equation*}%
in coordinates, where the $+$ is from the vector space structure from $%
T_{p_{\pm }}M$, and the $C^{1}$ difference is of order $C\varepsilon $. To
see this, we identify the geodesic ball of radius $C\varepsilon $ around $%
p_{\pm }$ to the ball in $T_{p_{\pm }}M$, and equip it with the Euclidean
metric $g_{p_{\pm }}$. If we deform the metric involved in the exponential
map in the expression
\begin{equation*}
\exp _{\chi \left( \varepsilon \tau \right) }\left[ \left( 1-\kappa
_{\varepsilon }^{0}\left( \tau \right) \right) E(\chi (\varepsilon \tau ),
u_{\pm }^{\varepsilon }\left( \tau ,t\right)) \right]
\end{equation*}
to the Euclidean metric $g_{p_{\pm }}$, the result becomes $\kappa
_{\varepsilon }^{0}\left( \tau \right) \chi (\varepsilon \tau )+\left(
1-\kappa _{\varepsilon }^{0}\left( \tau \right) \right) u_{\pm
}^{\varepsilon }\left( \tau ,t\right) $. \ Since the geodesic equation is a
second order differential equation whose coefficients are polynomials on
metric $g$ and its first order derivatives, by the differentiability of
solutions of ODE on finite interval $\left[ 0,1\right] $ with respect to its
initial condition and parameter, we see the $C^{1}$ norm of $\exp _{\chi
\left( \varepsilon \tau \right) }\bigskip \left[ \left( 1-\kappa
_{\varepsilon }^{0}\left( \tau \right) \right) \exp _{\chi (\varepsilon \tau
)}^{-1}\left( u_{\pm }^{\varepsilon }\left( \tau ,t\right) \right) \right] $
depends on the $C^{1}$ norm of the section $g\in \Gamma \left( Sym_{+}\left(
M\right) \right) $ with bounded Lipshitz constant, where $Sym_{+}\left(
M\right) $ is the space of positive definite symmetric tensors on $M$. Since
the $C^{1}$ difference of the metrics $g_{p_{\pm }}$ and $g$ is of order $%
C\varepsilon $ inside such ball, we have
\begin{eqnarray*}
\operatorname{dist}\left( v_\varepsilon^\pm(\tau,t),\widetilde
v_\varepsilon^\pm(\tau,t) \right) &\leq & C\varepsilon , \\
\Vert {\Greekmath 0272} v_\varepsilon^\pm(\tau,t) -Pal{\Greekmath 0272} \widetilde
v_\varepsilon^\pm(\tau,t)\Vert &\leq & C\varepsilon ,
\end{eqnarray*}%
where $Pal$ is the parallel transport along short geodesic from $\widetilde
v_\varepsilon^\pm(\tau,t)$ to $v_\varepsilon^\pm(\tau,t)$.
After the above simplification, we use the more explicit approximate solution%
\begin{equation}
u_{app}^{\varepsilon }\left( \tau ,t\right) =%
\begin{cases}
u_{-}(\tau +\tau (\varepsilon ),t) & -\infty <\tau \leq -l/\varepsilon \\
\kappa _{\varepsilon }^{0}\left( \tau \right) \chi (\varepsilon \tau
)+\left( 1-\kappa _{\varepsilon }^{0}\left( \tau \right) \right)
u_{-}^{\varepsilon }\left( \tau ,t\right) & -l/\varepsilon -1\leq \tau \leq
-l/\varepsilon \\
\chi (\varepsilon \tau ) & -l/\varepsilon \leq \tau \leq l/\varepsilon \\
\kappa _{\varepsilon }^{0}\left( \tau \right) \chi (\varepsilon \tau
)+\left( 1-\kappa _{\varepsilon }^{0}\left( \tau \right) \right)
u_{+}^{\varepsilon }\left( \tau ,t\right) & l/\varepsilon \leq \tau \leq
l/\varepsilon +1 \\
u_{+}(\tau -\tau (\varepsilon ),t) & l/\varepsilon +1\leq \tau <\infty%
\end{cases}
\label{eq:uappe-easy}
\end{equation}%
where the vector sum $+$ is from the linear space structure of $T_{p_{\pm
}}M $.
\begin{rem}
Apparently there enters no local model inserted at the joint points $p_{\pm
} $ to smooth out the join points in the construction of the above
approximate solution. Implicitly there is, though. The local model at $%
p_{\pm } $ is $u_{\pm }^{lmd}:{\mathbb{R}}\times S^{1}\rightarrow \mathbb{C}%
^{n}\simeq \left( T_{p_{\pm }}M,J_{p_{\pm }}\right) $,
\begin{equation*}
u_{\pm }^{lmd}\left( \tau ,t\right) =A_{\pm }z+a_{\pm }\tau ,
\end{equation*}%
where $z=e^{2\pi \left( \tau +it\right) }$, $A_{\pm }=u_{\pm }^{\prime
}\left( o_{\pm }\right) $, and $a_{\pm }={\Greekmath 0272} f\left( p_{\pm }\right) $.
Then one can see the local model is the linearized version of the above
intropolation of $u_{\pm }$ and $\chi $ in $u_{app}^{\varepsilon }$. Because
we can identify the portions of the approximate solution to $u_{\pm }$ and $%
\chi $ respectively and borrow Fredholm theories there, we do not need to
develop a Fredholm theory of the local model. However, when the gradient
flow length is $0$, the Fredholm theory of local model is needed for gluing,
because during compactification to nodal Floer trajectories the information
of ${\Greekmath 0272} f$ is lost (See \cite{oh-zhu}).
\end{rem}
\section{Off-shell formulation of resolved Floer trajectories for $\protect%
\varepsilon> 0$}
\label{sec:off-shell}
We define the Banach manifold to host resolved Floer trajectories near the
\textquotedblleft disk-flow-disk\textquotedblright\ Floer trajectories $%
(u_{-},(\chi ,l),u_{+})$. The precise description is in order.
First we define the Banach manifold ${\mathcal{B}}_{res}^{\varepsilon }={%
\mathcal{B}}_{res}^{\varepsilon }(z_{-},z_{+};l)$ for any $\varepsilon \in
(0,\varepsilon _{0})$ and $\ell \in (0,\infty )$, where $\varepsilon _{0}>0$
is a small constant to be determine later. To define $\mathcal{B}%
_{\varepsilon }$ that hosts resolved Floer trajectories out of $(u_{-},\chi
,l,u_{+})$, we define the weighting function $\beta _{\delta ,\varepsilon }$
as the gluing of the power weight and the exponential weight:
\begin{equation*}
\beta _{\delta ,\varepsilon }\left( \tau \right) =
\begin{cases}
1 & \tau <-\tau (\varepsilon ) \\
e^{2\pi \delta (\tau +\tau (\varepsilon ))} & -\tau (\varepsilon )\leq \tau
\leq -l/\varepsilon \\
\kappa _{\varepsilon }^{0}\left( \tau \right) \varepsilon ^{1-p+\delta
}\left( 1+\left\vert \tau \right\vert \right) ^{\delta }+\left( 1-\kappa
_{\varepsilon }^{0}\left( \tau \right) \right) e^{2\pi \delta (\tau +\tau
(\varepsilon ))} & -l/\varepsilon -1\leq \tau \leq -l/\varepsilon \\
\varepsilon ^{1-p+\delta }\left( 1+\left\vert \tau \right\vert \right)
^{\delta } & -l/\varepsilon \leq \tau \leq l/\varepsilon \\
\kappa _{\varepsilon }^{0}\left( \tau \right) \varepsilon ^{1-p+\delta
}\left( 1+\left\vert \tau \right\vert \right) ^{\delta }+\left( 1-\kappa
_{\varepsilon }^{0}\left( \tau \right) \right) e^{2\pi \delta (-\tau +\tau
(\varepsilon ))} & l/\varepsilon \leq \tau \leq l/\varepsilon +1 \\
e^{2\pi \delta (-\tau +\tau (\varepsilon ))} & l/\varepsilon +1\leq \tau
\leq \tau (\varepsilon ) \\
1 & \tau >\tau (\varepsilon )%
\end{cases}%
\end{equation*}
where $\kappa _{\varepsilon }^{0}\left( \tau \right) $ is the smooth cut-off
function defined in \ $\left( \ref{eq:beta0}\right) $ such that $\kappa
_{\varepsilon }^{0}\left( \tau \right) =1$ for $\left\vert \tau \right\vert
\leq l/\varepsilon $ and $\kappa _{\varepsilon }^{0}\left( \tau \right) =0$
for $\left\vert \tau \right\vert \geq l/\varepsilon +1$. Note that $\beta
_{\delta ,\varepsilon }\left( \tau \right) $ is no less than $1$ everywhere,
which is important for the uniform Sobolev constant we will discuss later in
Section \ref{sec:quadratic}.
In the above figure we put the graphs of various weighting functions
together, where the higher constant weight $\varepsilon^{1-p}$ in the
adiabatic weight $\|\cdot\|_{W^{1,p}_\varepsilon}$ is in red horizontal
line, power order weight $\rho_\varepsilon(\tau)$ is in green, and the
exponential weight is in blue. The weight $\beta_{\delta,\varepsilon}(\tau)$
is the glue of the power weight and the exponential weight with smoothing at
the corners, but to avoid too many graphs in the picture we did not draw the
smoothing. The two intervals $[l/\varepsilon-T(\varepsilon),
l/\varepsilon+T(\varepsilon)]$ and $[-l-T(\varepsilon),-l/\varepsilon+T(%
\varepsilon)]$ cut by four brown vertical lines are the places where
weighting function comparison occurs in right inverse estimates. For
convenience of readers, we also include the schematic picture of the
preglued solution $u^{\varepsilon}_{app}$ and Floer datum $%
(K_\varepsilon,J_\varepsilon)$ of the perturbed Cauchy-Riemann equation.
Note that in interval $\pm [l/\varepsilon,\tau(\varepsilon)]$, the
Hamiltonian $K_\varepsilon=0$ and $J_\varepsilon=J_0$ because of the cut-off
function $\kappa^{\pm}_\varepsilon(\tau)$.
\begin{figure}[tbp]
\centering
\includegraphics{figures2.eps}
\caption{Weighting functions}
\end{figure}
This Banach manifold can be thought as the gluing of the Banach manifolds $%
\mathcal{B}_{\pm }$ and $\mathcal{B}_{\chi _{\varepsilon }}$. More precisely
${\mathcal{B}}_{res}^{\varepsilon }(z_{-},z_{+};l/\varepsilon )$ $\left(
l\geq l_{0}>0\right) $ consists of maps $u:\Sigma _{\varepsilon }\rightarrow
M$ satisfying:
\begin{enumerate}
\item $\Sigma_\varepsilon$ is diffeomorphic ${\mathbb{R}} \times S^1$ but
equipped with the conformal structure induced by the following metric
arising from the decomposition into the standard cylinder and hemispheres $%
S_\pm$,
\begin{equation} \label{eq:Sigmae}
\Sigma _{\varepsilon } = S_{-}\cup ([-\tau(\varepsilon) ,\tau(\varepsilon)]
\times S^{1}) \cup S_{+}.
\end{equation}
\item $u\in W^{1,p}_{loc}(\Sigma_\varepsilon,M)$
\item $\lim_{\tau\to +\infty} u(\tau,t)=z_{+}(t)$ and $\lim_{\tau\to
-\infty}u(\tau,t)=z_{-}(t)$ for all $t\in S^1$.
\item For $u\in \mathcal{B}_{\varepsilon }$, and any variation vector field $%
\xi \in \Gamma \left( W^{1,p}\left( u^{\ast }TM\right) \right) $, we define
the Banach norm to be
\begin{equation}
\left\Vert \xi \right\Vert _{\varepsilon }=\left\Vert \widetilde{\xi }%
\right\Vert _{W_{\beta _{\delta ,\varepsilon }}^{1,p}\left( \Sigma
_{\varepsilon }\right) }+\left\Vert \xi _{0}\right\Vert _{W_{\varepsilon
}^{1,p}\left( -l/\varepsilon ,l/\varepsilon \right) }+\left\vert \xi
_{0}\left( \pm l/\varepsilon \right) \right\vert , \label{eq:xienorm}
\end{equation}%
where
\begin{eqnarray*}
\xi _{0}\left( \tau \right) &=&\left\{
\begin{tabular}{ll}
$\int_{S^{1}}\xi \left( \tau ,t\right) dt,$ & $\left\vert \tau \right\vert
\leq l/\varepsilon $ \\
$\kappa _{\varepsilon }^{0}\left( \tau \right) \int_{S^{1}}\xi \left( \pm
l/\varepsilon ,t\right) dt,$ & $\left\vert \tau \right\vert \geq
l/\varepsilon $%
\end{tabular}%
\right. , \\
\widetilde{\xi }\left( \tau ,t\right) &=&\xi \left( \tau ,t\right) -\xi
_{0}\left( \tau \right) , \\
\left\Vert \widetilde{\xi }\right\Vert _{W_{\beta _{\delta ,\varepsilon
}}^{1,p}\left( \Sigma _{\varepsilon }\right) }^{p} &=&\int \int_{\Sigma
_{\varepsilon }}\left( |\widetilde{\xi }|^{p}+|{\Greekmath 0272} \widetilde{\xi }%
|^{p}\right) \beta _{\delta ,\varepsilon }\left( \tau \right) dvol_{\Sigma
_{\varepsilon }}.
\end{eqnarray*}
\end{enumerate}
Therefore, we have an $\varepsilon $-family of Banach manifolds ${\mathcal{B}%
}_{res}^{\varepsilon }(z_{-},z_{+};l/\varepsilon )$, and an $\varepsilon $%
-family of equations ${\overline{\partial }}_{(J_{\varepsilon
},K_{\varepsilon })}u^{\varepsilon }=0$ defined on each Banach bundle
\begin{equation*}
\pi :{\mathcal{L}}_{res}^{\varepsilon }(z_{-},z_{+};l/\varepsilon
)\rightarrow {\mathcal{B}}_{res}^{\varepsilon }(z_{-},z_{+};l/\varepsilon ),
\end{equation*}%
where
\begin{equation*}
{\mathcal{L}}_{res}^{\varepsilon }(z_{-},z_{+};l/\varepsilon )=\bigcup_{u\in
{\mathcal{B}}_{res}^{\varepsilon }(z_{-},z_{+};l/\varepsilon )}L_{\beta
_{\delta ,\varepsilon }}^{p}(\Lambda ^{0,1}(u^{\ast }TM)).
\end{equation*}%
Here each fiber $L_{\beta _{\delta ,\varepsilon }}^{p}(\Lambda
^{0,1}(u^{\ast }TM))$ consists of sections $\eta \in L^{p}\left( \Lambda
^{0,1}(u^{\ast }TM)\right) $ with $\Vert \eta \Vert _{\varepsilon }<\infty $
and the norm $\Vert \eta \Vert _{\varepsilon }$ is given by
\begin{equation*}
\left\Vert \eta \right\Vert _{\varepsilon }=\left\Vert \widetilde{\eta }%
\right\Vert _{L_{\beta _{\delta ,\varepsilon }}^{p}\left( \Sigma
_{\varepsilon }\right) }+\left\Vert \kappa _{\varepsilon }^{0}\left( \tau
\right) \eta _{0}\left( \tau \right) \right\Vert _{L_{\varepsilon }^{p}\left[
-l/\varepsilon ,l/\varepsilon \right] },
\end{equation*}%
where $\overline{\eta },\eta _{0}$ and $\widetilde{\eta }$ are defined
similarly as those for $\xi ,$ namely
\begin{eqnarray*}
\eta _{0}\left( \tau \right) &=&\left\{
\begin{tabular}{ll}
$\int_{S^{1}}\eta \left( \tau ,t\right) dt,$ & $\left\vert \tau \right\vert
<l/\varepsilon $ \\
$0$ & $\left\vert \tau \right\vert \geq l/\varepsilon $%
\end{tabular}%
\right. , \\
\widetilde{\eta }\left( \tau ,t\right) &=&\eta \left( \tau ,t\right) -\eta
_{0}\left( \tau \right) , \\
\left\Vert \widetilde{\eta }\right\Vert _{L_{\beta _{\delta ,\varepsilon
}}^{p}\left( \Sigma _{\varepsilon }\right) }^{p} &=&\int \int_{\Sigma
_{\varepsilon }}\left\vert \widetilde{\eta }\right\vert ^{p}\beta _{\delta
,\varepsilon }\left( \tau \right) dvol_{\Sigma _{\varepsilon }}.
\end{eqnarray*}
We define
\begin{eqnarray}
{\mathcal{B}}_{res}^{\varepsilon }(z_{-},z_{+}) &=&\bigcup_{l\geq l_{0}}{%
\mathcal{B}}_{res}^{\varepsilon }(z_{-},z_{+};l/\varepsilon ) \label{eq:CBe}
\\
{\mathcal{L}}_{res}^{\varepsilon }(z_{-},z_{+}) &=&\bigcup_{l\geq l_{0}}{%
\mathcal{L}}_{res}^{\varepsilon }(z_{-},z_{+};l/\varepsilon ).
\end{eqnarray}%
For $(u,l/\varepsilon )\in {\mathcal{B}}_{res}^{\varepsilon }(z_{-},z_{+})$,
its tangent space consists of elements $(\xi ,\mu )$ where $\xi \in T_{u}{%
\mathcal{B}}_{res}^{\varepsilon }(z_{-},z_{+};l/\varepsilon )$ and $\mu \in
T_{l/\varepsilon }{\mathbb{R}}_{+}\cong {\mathbb{R}}$ with the norm
\begin{equation*}
\Vert (\xi ,\mu )\Vert _{\varepsilon }=\Vert \xi \Vert _{\varepsilon }+|\mu
|.
\end{equation*}%
Geometrically $\mu $ corresponds to the variation of conformal structure of
the neck cylinder $\left[ -l/\varepsilon ,l/\varepsilon \right] \times S^{1}$
by varying the length of the cylinder but keeping the radius of $S^{1}$
fixed. For $\mu \in T_{l/\varepsilon }\mathbb{R}$, the induced path in ${%
\mathcal{B}}_{res}^{\varepsilon }\left( z_{-},z_{+}\right) $, starting from $%
u\in {\mathcal{B}}_{res}^{\varepsilon }\left( z_{-},z_{+};l/\varepsilon
\right) $, is $u_{s}\in {\mathcal{B}}_{res}^{\varepsilon }\left(
z_{-},z_{+};l/\varepsilon -s\mu \right) $, where $u_{s}(\tau ^{\prime },t)$
is the reparameterization of $u(\tau ,t)$ on the neck part,%
\begin{equation}
u_{s}(\tau ^{\prime },t):=u\left( \frac{\tau ^{\prime }}{l-s\varepsilon \mu }%
,t\right) ,\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ \ \ \ \ \ }\left( \tau ^{\prime },t\right) \in \left[
-\left( l/\varepsilon -s\mu \right) ,l/\varepsilon -s\mu \right] \times
S^{1}, \label{reparameterize-tau}
\end{equation}%
for $s$ nearby $0$. There is a canonical way to associate points on $u$ to
points on $u_{s}:$%
\begin{equation*}
u\left( \tau ,t\right) \longleftrightarrow u_{s}\left( \tau ^{\prime
},t\right) ,\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ \ \ \ \ \ where }\tau ^{\prime }=\frac{(l-s\varepsilon
\mu )\tau }{l}.
\end{equation*}%
Using this identification we can realize the variation of conformal
structure as a vector field on $\left[ -l/\varepsilon ,l/\varepsilon \right]
\times S^{1}$.
Here we define the Banach norm on $T_{u}{\mathcal{B}}_{res}^{\varepsilon
}(z_{-},z_{+};l/\varepsilon )$ by
\begin{equation}
\left\Vert \xi \right\Vert _{\varepsilon }=\left\Vert \widetilde{\xi }%
\right\Vert _{W_{\beta _{\delta ,\varepsilon }}^{1,p}\left( \mathbb{R}\times
S^{1}\right) }+\left\Vert \xi _{0}\right\Vert _{W_{\varepsilon }^{1,p}\left( %
\left[ -l/\varepsilon ,l/\varepsilon \right] \right) }+\left\vert \xi
_{0}\left( \pm l/\varepsilon \right) \right\vert , \label{eq:normxie}
\end{equation}%
where
\begin{eqnarray*}
\xi _{0}\left( \tau \right) &=&%
\begin{cases}
\int_{S^{1}}\xi \left( \tau ,t\right) dt,\quad & \left\vert \tau \right\vert
\leq l/\varepsilon \\
\kappa _{\varepsilon }^{0}\left( \tau \right) \int_{S^{1}}\xi \left( \pm
l/\varepsilon ,t\right) dt, & \left\vert \tau \right\vert \geq l/\varepsilon%
\end{cases}
\\
\widetilde{\xi }\left( \tau ,t\right) &=&\xi \left( \tau ,t\right) -\xi
_{0}\left( \tau \right) , \\
\left\Vert \widetilde{\xi }\right\Vert _{W_{\beta _{\delta ,\varepsilon
}}^{1,p}\left( \mathbb{R}\times S^{1}\right) }^{p} &=&\int \int_{\mathbb{R}%
\times S^{1}}\left( \left\vert \widetilde{\xi }\right\vert ^{p}+\left\vert
{\Greekmath 0272} \widetilde{\xi }\right\vert ^{p}\right) \beta _{\delta ,\varepsilon
}\left( \tau \right) d\tau dt
\end{eqnarray*}%
This Banach manifold can be thought as the gluing of the Banach manifolds $%
\mathcal{B}_{\pm }$ and $\mathcal{B}_{\chi _{\varepsilon }}$. Similarly for $%
\eta \in \Gamma \left( L^{p}\left( u^{\ast }TM\otimes \Lambda ^{0,1}\left(
\Sigma _{\varepsilon }\right) \right) \right) $, the Banach norm is
\begin{equation*}
\left\Vert \eta \right\Vert _{\varepsilon }=\left\Vert \widetilde{\eta }%
\right\Vert _{L_{\beta _{\delta ,\varepsilon }}^{p}\left( \mathbb{R}\times
S^{1}\right) }+\left\Vert \kappa _{\varepsilon }^{0}\left( \tau \right) \eta
_{0}\left( \tau \right) \right\Vert _{L_{\varepsilon }^{p}\left[
-l/\varepsilon ,l/\varepsilon \right] },
\end{equation*}%
where $\overline{\eta },\eta _{0}$ and $\widetilde{\eta }$ are defined
similarly as those for $\xi ,$ namely
\begin{eqnarray*}
\eta _{0}\left( \tau \right) &=&%
\begin{cases}
\int_{S^{1}}\eta \left( \tau ,t\right) dt,\quad & \left\vert \tau
\right\vert <l/\varepsilon \\
0\quad & \left\vert \tau \right\vert \geq l/\varepsilon%
\end{cases}
\\
\widetilde{\eta }\left( \tau ,t\right) &=&\eta \left( \tau ,t\right) -\eta
_{0}\left( \tau \right) , \\
\left\Vert \widetilde{\eta }\right\Vert _{L_{\beta _{\delta ,\varepsilon
}}^{p}\left( \mathbb{R}\times S^{1}\right) }^{p} &=&\int \int_{\mathbb{R}%
\times S^{1}}\left\vert \widetilde{\eta }\right\vert ^{p}\beta _{\delta
,\varepsilon }\left( \tau \right) d\tau dt.
\end{eqnarray*}
\begin{rem}
In the norm $\left\Vert \xi \right\Vert _{\varepsilon }$ we could have
dropped the term $\left\vert \xi _{0}\left( \pm l/\varepsilon \right)
\right\vert $ but still get an equivalent norm, because from Sobolev
embedding we have
\begin{equation*}
\left\vert \xi _{0}\left( \pm l/\varepsilon \right) \right\vert \leq
C\left\Vert \xi _{0}\right\Vert _{W_{\varepsilon }^{1,p}\left(
[-l/\varepsilon ,l/\varepsilon ]\right) }
\end{equation*}%
where $C$ is uniform for all $\varepsilon $ and $l\geq l_{0}>0$. However we
still keep the $\left\vert \xi _{0}\left( \pm l/\varepsilon \right)
\right\vert $ term because $\left\Vert \widetilde{\xi }\right\Vert
_{W_{\beta _{\delta ,\varepsilon }}^{1,p}\left( \left\{ \left\vert \tau
\right\vert >l/\varepsilon \right\} \times S^{1}\right) }+\left\vert \xi
_{0}\left( \pm l/\varepsilon \right) \right\vert $ mimics the Banach norm of
$\xi _{\pm }$ and $\left\Vert \widetilde{\xi }\right\Vert _{W_{\beta
_{\delta ,\varepsilon }}^{1,p}\left( \left\{ \left\vert \tau \right\vert
\leq l/\varepsilon \right\} \times S^{1}\right) }+\left\Vert \xi _{0}\left(
\tau \right) \right\Vert _{W_{\varepsilon }^{1,p}\left( \left[
-l/\varepsilon ,l/\varepsilon \right] \right) }$ mimics the Banach norm of $%
\xi _{\chi _{\varepsilon }}$ and in this sense the norm $\left\Vert \xi
\right\Vert _{\varepsilon }$ is roughly the sum of them.
\end{rem}
Finally we consider the Fredholm section
\begin{equation*}
{\mathcal{B}}_{res}^{\varepsilon }(z_{-},z_{+})\rightarrow {\mathcal{L}}%
_{res}^{\varepsilon }(z_{-},z_{+})
\end{equation*}%
given by
\begin{equation*}
(u,l/\varepsilon )\mapsto {\overline{\partial }}_{K_{\varepsilon
},J_{\varepsilon };l/\varepsilon }(u):{\mathcal{B}}_{res}^{\varepsilon
}(z_{-},z_{+})\rightarrow {\mathcal{L}}_{res,u}^{\varepsilon
}(z_{-},z_{+};l/\varepsilon )
\end{equation*}%
defined fiberwise over $l/\varepsilon \in {\mathbb{R}}$. We denote the
linearization of this map by $D_{para}^{\varepsilon }$ which has expression
\begin{equation}
D_{para}^{\varepsilon }(\xi ,\mu )=D^{\varepsilon }(\xi )+\frac{\mu }{l}%
{\Greekmath 0272} \varepsilon f(\chi _{\varepsilon }). \label{eq:tildeDe}
\end{equation}
\section{$\overline{\partial }_{\left( K_{\protect\varepsilon },J_{\protect%
\varepsilon }\right) }$-error estimate\label{section:d-bar-error}}
In this section, we estimate the norm $\Vert \overline{\partial }%
_{(K_{\varepsilon },J_{\varepsilon })}u_{app}^{\varepsilon }\Vert
_{\varepsilon }$. Let $\eta =\overline{\partial }_{(K_{\varepsilon
},J_{\varepsilon })}u_{app}^{\varepsilon }$. We do this estimation
separately in several regions: \medskip
(1). When $\left\vert \tau \right\vert \leq l/\varepsilon $, $%
u_{app}^{\varepsilon }\left( \tau ,t\right) =\chi \left( \varepsilon \tau
\right) $, so%
\begin{equation*}
\overline{\partial }_{\left( K_{\varepsilon },J_{\varepsilon }\right)
}u_{app}^{\varepsilon }\left( \tau ,t\right) =\overline{\partial }_{\left(
J_{0},\varepsilon f\right) }\chi \left( \varepsilon \tau \right) =\frac{%
\partial }{\partial \tau }\chi \left( \varepsilon \tau \right) -\varepsilon
{\Greekmath 0272} f\left( \chi \left( \varepsilon \tau \right) \right) =0;
\end{equation*}
(2). When $l/\varepsilon \leq \left\vert \tau \right\vert \leq l/\varepsilon
+1$, say $\tau >0$ case we have $u_{app}^{\varepsilon }\left( \tau ,t\right)
=\kappa _{\varepsilon }^{0}\left( \tau \right) \chi (\varepsilon \tau
)+\left( 1-\kappa _{\varepsilon }^{0}\left( \tau \right) \right) u_{\pm
}^{\varepsilon }\left( \tau ,t\right) $, so%
\begin{eqnarray}
\left\vert \eta \left( \tau ,t\right) \right\vert &=&\left\vert \overline{%
\partial }_{\left( J_{0},\varepsilon f\right) }u_{app}^{\varepsilon
}\right\vert \notag \\
&=&\left\vert \left( \kappa _{\varepsilon }^{0}\left( \tau \right) \right)
^{\prime }\left( \chi \left( \varepsilon \tau \right) -u_{\pm }^{\varepsilon
}\right) +\left( 1-\kappa _{\varepsilon }^{0}\left( \tau \right) \right)
\overline{\partial }_{\left( J_{0},\varepsilon f\right) }u_{\pm
}^{\varepsilon }\right\vert \notag \\
&\leq &C\left( \left\vert \chi \left( \varepsilon \tau \right) -u_{\pm
}^{\varepsilon }\left( \tau ,t\right) \right\vert +\left\vert \varepsilon
{\Greekmath 0272} f\left( u_{\pm }^{\varepsilon }\left( \tau ,t\right) \right)
\right\vert \right) \notag \\
&\leq &\widetilde{C}\left( l\right) \varepsilon +C\varepsilon :=C\left(
l\right) \varepsilon , \label{d-bar-error}
\end{eqnarray}%
where in the first inequality we have used $\overline{\partial }_{\left(
J_{0},\varepsilon f\right) }u_{\pm }^{\varepsilon }=\varepsilon {\Greekmath 0272}
f\left( u_{\pm }^{\varepsilon }\right) $ and in the second inequality used $%
\left( \ref{dist-xp}\right) $ and $\left( \ref{dist-up}\right) $, and $%
\widetilde{C}\left( l\right) =Cl^{-\frac{p-1}{\delta }}$. The weight $\beta
_{\delta ,\varepsilon }$ in this interval satisfies%
\begin{equation*}
\beta _{\delta ,\varepsilon }\left( \tau \right) =e^{2\pi \delta \left(
-\tau +\tau \left( \varepsilon \right) \right) }\leq e^{2\pi \delta \left(
-l/\varepsilon +\tau \left( \varepsilon \right) \right) }=e^{\left(
p-1\right) \ln \left( 1+l/\varepsilon \right) }\leq D\left( l\right)
\varepsilon ^{1-p},
\end{equation*}%
where the constant $D\left( l\right) \approx l^{p-1}$. Since $\left\vert
\tau \right\vert \geq l/\varepsilon $, we do not need to distinguish the $0$%
-mode and higher mode of $\eta $. So we have
\begin{equation*}
\left\Vert \eta \left( \tau \right) \right\Vert _{L_{\beta _{\delta
,\varepsilon }}^{p}\left( \left[ l/\varepsilon 1,l/\varepsilon +1\right]
\times S^{1}\right) }\leq \left( \int_{l/\varepsilon }^{l/\varepsilon
+1}\left( C\left( l\right) \varepsilon \right) ^{p}\cdot D\left( l\right)
\varepsilon ^{1-p}d\tau \right) ^{\frac{1}{p}}:=E\left( l\right) \varepsilon
^{\frac{1}{p}},
\end{equation*}%
Similarly result holds for $-l/\varepsilon -1\leq \tau \leq -l/\varepsilon $
case.
(3). When $\left\vert \tau \right\vert >l/\varepsilon $, say $\tau
>l/\varepsilon $, we recall
\begin{equation*}
u_{app}^{\varepsilon }\left( \tau ,t\right) =u_{+}\circ I_{\tau (\varepsilon
)}(\tau ,t)=u_{+}(\tau -\tau (\varepsilon ),t)
\end{equation*}%
and so satisfies
\begin{equation*}
\overline{\partial }_{\left( K_{\varepsilon },J_{\varepsilon }\right)
}u_{app}^{\varepsilon }\left( \tau ,t\right) =\overline{\partial }_{\left(
K_{\varepsilon }^{+},J_{\varepsilon }^{+}\right) }u_{+}\left( \tau -\tau
(\varepsilon ),t\right) =0.
\end{equation*}%
Similarly $\overline{\partial }_{\left( K_{\varepsilon },J_{\varepsilon
}\right) }u_{app}^{\varepsilon }\left( \tau ,t\right) =0$ for $\tau
<-l/\varepsilon .$
Combining the 3 pieces, we have%
\begin{equation}
\left\Vert \overline{\partial }_{\left( J,\varepsilon f\right)
}u_{app}^{\varepsilon }\right\Vert _{_{L_{\beta _{\delta ,\varepsilon
}}^{p}\left( \mathbb{R}\times S^{1}\right) }}\leq E\left( l\right)
\varepsilon ^{\frac{1}{p}}, \label{eq:error}
\end{equation}%
where the constant $E\left( l\right) \approx \left( C+l^{-\frac{p-1}{\delta }%
}\right) l^{\frac{p-1}{p}}.$
\begin{rem}
If we use $u_{app}^{\varepsilon }=\exp _{\chi \left( \varepsilon \tau
\right) }\bigskip \left[ \left( 1-\kappa _{\varepsilon }^{0}\left( \tau
\right) \right) \exp _{\chi (\varepsilon \tau )}^{-1}\left( u_{\pm
}^{\varepsilon }\left( \tau ,t\right) \right) \right] $ for $\tau \in
\lbrack l/\varepsilon ,l/\varepsilon +1]$, then $\left\vert \eta \right\vert
=$ $\left\vert \overline{\partial }_{(J_{0},\varepsilon
f)}u_{app}^{\varepsilon }\right\vert $ is controlled pointwise by $%
C\varepsilon $ as above case $\left( 2\right) $, plus the $C^{1}$ difference
between $\exp _{\chi \left( \varepsilon \tau \right) }\bigskip \left[ \left(
1-\kappa _{\varepsilon }^{0}\left( \tau \right) \right) \exp _{\chi
(\varepsilon \tau )}^{-1}\left( u_{\pm }^{\varepsilon }\left( \tau ,t\right)
\right) \right] $ and $\kappa _{\varepsilon }^{0}\left( \tau \right) \chi
(\varepsilon \tau )+\left( 1-\kappa _{\varepsilon }^{0}\left( \tau \right)
\right) u_{-}^{\varepsilon }\left( \tau ,t\right) $, which is also of order $%
C\varepsilon $. Therefore we get the same pointwise estimate
\begin{equation*}
\left\vert \eta \right\vert =\left\vert \overline{\partial }%
_{(J_{0},\varepsilon f)}u_{app}^{\varepsilon }\right\vert \leq C\varepsilon .
\end{equation*}%
Continuing the remaining steps in case $\left( 2\right) $ we get the same $%
L_{\beta _{\delta ,\varepsilon }}^{p}$ estimate
\begin{equation*}
\left\Vert \eta \right\Vert _{L_{\beta _{\delta ,\varepsilon }}^{p}\left(
\pm \left[ l/\varepsilon -1,l/\varepsilon \right] \times S^{1}\right) }\leq
E\left( l\right) \varepsilon ^{\frac{1}{p}}.
\end{equation*}%
The case $\left( 1\right) $ and $\left( 3\right) $ are the same as above.
Therefore we still get
\begin{equation*}
\left\Vert \overline{\partial }_{\left( J_{0},\varepsilon f\right)
}u_{app}^{\varepsilon }\right\Vert _{_{L_{\beta _{\delta ,\varepsilon
}}^{p}\left( \mathbb{R}\times S^{1}\right) }}\leq E\left( l\right)
\varepsilon ^{\frac{1}{p}}.
\end{equation*}
\end{rem}
\begin{rem}
\label{exp-weight-dbar-error}The constant $E\left( l\right) $ is bounded if $%
l$ is in a bounded interval, because we see from the above estimates%
\begin{equation*}
E\left( l\right) \approx \left( C+l^{-\frac{p-1}{\delta }}\right) l^{\frac{%
p-1}{p}}.
\end{equation*}%
So if we assume that $l_{0}\leq l\leq L$, then we get uniform $\overline{%
\partial }_{\left( K_{\varepsilon },J_{\varepsilon }\right) }$ error
estimate. When $l\rightarrow $ $\infty $, $E\left( l\right) \rightarrow
\infty $, so the above estimate in $\left( \ref{d-bar-error}\right) $ is too
coarse. But notice that when $l\rightarrow \infty $, ${\Greekmath 0272} f\left( \chi
\left( \pm l\right) \right) $ has exponential decay $e^{-cl}$, so using this
in $\left( \ref{d-bar-error}\right) $ one can get even better $\overline{%
\partial }_{\left( K_{\varepsilon },J_{\varepsilon }\right) }$ error
estimate with $E\left( l\right) \approx \left( Ce^{-cl}+l^{-\frac{p-1}{%
\delta }}\right) l^{\frac{p-1}{p}}\rightarrow 0$ as $l\rightarrow \infty $.
Thus the error estimate is uniform for all $l\geq l_{0}$.
\end{rem}
\section{The combined right inverse}
\label{sec:comb-rightinverse}
To keep notation simple, in the following we denote $D^{\varepsilon
}:=D_{u_{app}^{\varepsilon }}\overline{\partial }_{\left( K_{\varepsilon
},J_{\varepsilon };\right) }$ for the linearization of
\begin{equation*}
{\overline{\partial }}_{(K_{\varepsilon },J_{\varepsilon
})}:T_{(u,l/\varepsilon )}{\mathcal{B}}_{res}^{\varepsilon
}(z_{-},z_{+};l/\varepsilon )\rightarrow {\mathcal{L}}_{res;(u,l/\varepsilon
)}^{\varepsilon }(z_{-},z_{+};l).
\end{equation*}%
In this section, we first construct an approximate right inverse, denoted by
$Q_{para}^{app;\varepsilon }$, of the parameterized ${\overline{\partial }}%
_{(K_{\varepsilon },J_{\varepsilon })}$ which allows $l$ to vary, which we
denote by
\begin{equation*}
D_{para}^{\varepsilon }:T_{(u,l/\varepsilon )}{\mathcal{B}}%
_{res}^{\varepsilon }(z_{-},z_{+})\rightarrow {\mathcal{L}}%
_{res;(u,l/\varepsilon )}^{\varepsilon }(z_{-},z_{+})
\end{equation*}%
(see \eqref{eq:tildeDe}) by gluing the right inverses $Q_{\pm }$ of $D{%
\overline{\partial }}_{(K_{\pm },J_{\pm })}(u_{\pm })$ and another operator
that takes care of the part of the gradient segment in the middle. We recall
\begin{equation*}
T_{(u,l/\varepsilon )}{\mathcal{B}}_{res}^{\varepsilon }(z_{-},z_{+})=T_{u}{%
\mathcal{B}}_{res}^{\varepsilon }(z_{-},z_{+};l/\varepsilon )\oplus
T_{l/\varepsilon }{\mathbb{R}}=W_{\rho _{\varepsilon }}^{(1,p)}(u^{\ast
}TM)\oplus {\mathbb{R}}
\end{equation*}%
and
\begin{equation*}
{\mathcal{L}}_{res;(u,l/\varepsilon )}^{\varepsilon }(z_{-},z_{+})={\mathcal{%
L}}_{res;u}^{\varepsilon }(z_{-},z_{+};l/\varepsilon )=L_{\rho _{\varepsilon
}}^{p}(u^{\ast }TM).
\end{equation*}%
Therefore the image of $Q_{para}^{app;\varepsilon }$ at $(u,l/\varepsilon )$
is decomposed into
\begin{equation*}
Q_{para}^{app;\varepsilon }(\eta )=(\xi _{\varepsilon },\mu ),\,\xi
_{\varepsilon }\in W_{\rho _{\varepsilon }}^{(1,p)}(u^{\ast }TM),\,\mu \in {%
\mathbb{R}}.
\end{equation*}%
We will define each of $\xi _{\varepsilon }$ and $\mu $ in describing the
image of the operator $Q_{para}^{app;\varepsilon }(\eta )$.
We introduce several cut-off functions:
\begin{enumerate}
\item $\kappa _{0}^{\varepsilon }=\kappa _{0}^{\varepsilon }\left( \tau
\right) $ is the characteristic function of the interval $\left[
-l/\varepsilon ,l/\varepsilon \right] \subset \mathbb{R}$,
\item $\varphi _{0}^{K}$ is the cut-off function defined by
\begin{equation*}
\varphi_{0}^{K}\left( \tau \right) =
\begin{cases}
1 \quad & \mbox{ for }\, \left\vert \tau \right\vert \leq K \\
0 \quad & \mbox{ for }\, \left\vert \tau \right\vert >K+1,%
\end{cases}%
\end{equation*}
\item $\varphi_{+}^{K}$ is the cut-off function
\begin{equation*}
\varphi_{+}^{K}\left( \tau \right) =
\begin{cases}
1\quad & \mbox{ for }\, \tau \leq K \\
0 \quad & \mbox{ for }\, \tau >K+1,%
\end{cases}%
\end{equation*}
and $\varphi_{-}^{K}$ is the function defined by $\varphi_-^K\left( \tau
\right) =1-\varphi _{+}^{K}\left( \tau \right) $.
\end{enumerate}
Now let $\eta \in {\mathcal{L}}_{res}^{\varepsilon }(z_{-},z_{+})$. We split
$\eta $ into 3 pieces of ${\mathbb{R}}$ according to the division of the
expression of the approximate solution \eqref{eq:uappe}:
\begin{equation}
\eta |_{(-\infty ,l/\varepsilon ]},\,\eta _{\lbrack l/\varepsilon ,\infty
)},\,\eta _{\lbrack -l/\varepsilon ,l/\varepsilon ]}. \label{eq:divided-eta}
\end{equation}%
Multiplying the characteristic functions
\begin{equation*}
\kappa _{-}^{\varepsilon },\,\kappa _{+}^{\varepsilon },\,\kappa
_{0}^{\varepsilon }
\end{equation*}%
of the corresponding intervals, we regard each of them defined on the whole
cylinder ${\mathbb{R}}\times S^{1}$. (Remark: the smooth cut-off functions
we used before are $\kappa _{\varepsilon }^{-},\,\kappa _{\varepsilon
}^{+},\kappa _{\varepsilon }^{0}$, \ whose notations are similar to
characteristic functions but lower and upper indices are switched.)
Consider the translations of the first and the third pieces
\begin{equation*}
\eta_\pm^\varepsilon(\tau,t) := (\kappa_{\pm}^{\varepsilon }\eta) \circ
I_{\pm \tau(\varepsilon)}(\tau,t) = (\kappa_{\pm}^{\varepsilon }\eta)(\tau -
\pm \tau(\varepsilon),t)
\end{equation*}
whose supports become
\begin{equation*}
\left(-\infty,\frac{p-1}{\delta}S(\varepsilon)\right) \times S^1, \, \left(-%
\frac{p-1}{\delta}S(\varepsilon),\infty\right) \times S^1
\end{equation*}
respectively. Then we define
\begin{equation} \label{eq:xiepm}
\xi^\varepsilon_\pm = Q_\pm(\eta_\pm^\varepsilon)\circ
I_{\pm(-\tau(\varepsilon))}.
\end{equation}
Now we consider the middle piece $\kappa _{0}^{\varepsilon }\eta $ which is
supported on $[-l/\varepsilon ,l/\varepsilon ]\times S^{1}$. To describe
this piece precisely, we need some careful examination how the Banach
manifold \eqref{eq:CBe} and the tangent vectors thereof at the approximate
solution $u_{\varepsilon }$ near $(u_{-},\chi ,u_{+},l)$ are made of, and
how the operator ${\overline{\partial }}_{(K_{\varepsilon },J_{\varepsilon
})}$ acts on $T_{u_{\varepsilon }}{\mathcal{B}}_{res}^{\varepsilon
}(z_{-},z_{+})$.
Recall $\chi$ satisfies $\dot \chi + \operatorname{grad} f(\chi) = 0$. Since we
assume that the pair $(f,g)$ is Morse-Smale, the linearization
\begin{equation*}
D_\chi = {\Greekmath 0272}_\tau + {\Greekmath 0272} \operatorname{grad}f(\chi)
\end{equation*}
is invertible as mentioned before. We denote by $Q_\chi$ its right inverse.
We denote the renormalized $\chi :[-l,l]\rightarrow M$ by
\begin{equation*}
\chi _{\varepsilon }:[-l/\varepsilon ,l/\varepsilon ]\rightarrow M,\quad
\chi _{\varepsilon }(\tau ):=\chi (\varepsilon \tau )
\end{equation*}%
which satisfies $\dot{\chi}_{\varepsilon }+\varepsilon \operatorname{grad}f(\chi )=0$%
. We apply the right inverse $Q_{\chi _{\varepsilon }}^{para}$ (allowing $l$
to vary) of $D_{\chi _{\varepsilon }}^{para}$ to $\kappa _{0}^{\varepsilon
}\eta $ and write it as
\begin{equation*}
Q_{\chi _{\varepsilon }}^{para}(\kappa _{0}^{\varepsilon }\eta )=(\xi _{\chi
_{\varepsilon }},\mu ).
\end{equation*}%
We recall the operator
\begin{equation*}
D_{para}^{\varepsilon }:W^{1,p}(\chi _{\varepsilon }^{\ast }TM\times
S^{1})\oplus {\mathbb{R}}\rightarrow L^{p}(\chi _{\varepsilon }^{\ast
}TM\times S^{1})
\end{equation*}%
has the form
\begin{equation*}
D_{para}^{\varepsilon }(\xi ,\mu )=D_{\chi _{\varepsilon }}(\xi )+\frac{\mu
}{l}{\Greekmath 0272} \varepsilon f(\chi _{\varepsilon })
\end{equation*}%
with $D_{\chi _{\varepsilon }}=\frac{\partial }{\partial \tau }+J_{0}\frac{%
\partial }{\partial t}+{\Greekmath 0272} \,$grad$\left( \varepsilon f\right) $. First
we define
\begin{equation*}
\xi _{\chi _{\varepsilon }}=Q_{\chi _{\varepsilon }}(\eta _{\chi
_{\varepsilon }})
\end{equation*}%
where $Q_{\chi _{\varepsilon }}$ is the right inverse of $D_{\chi
_{\varepsilon }}$ constructed in section \ref{sec:Qinmiddle} Then we
determine $\mu $ by solving the $0$\emph{-mode matching condition}
\begin{equation}
\left( \xi _{\chi _{\varepsilon }}\right) _{0}\left( \pm l/\varepsilon
\right) +\frac{\varepsilon \mu }{l}{\Greekmath 0272} f\left( p_{\pm }\right) =\xi _{\pm
}\left( o_{\pm }\right) . \label{eq:defmu}
\end{equation}%
Now we are ready to write down the formula for $(\xi _{\varepsilon },\mu
)=Q_{para}^{app;\varepsilon }(\eta )$. Here we define
\begin{eqnarray*}
\xi _{\varepsilon } =
\begin{cases}
\xi _{+}^{\varepsilon }, \hfill & l/\varepsilon +T\left( \varepsilon \right)
\leq \tau , \\
Pal_{\chi ,\varepsilon }\left[ \left( \xi _{\chi _{\varepsilon }}\right)
_{0}+\phi _{0}^{l/\varepsilon +T\left( \varepsilon \right) }\left( \tau
\right) \left( \xi _{\chi _{\varepsilon }}-\left( \xi _{\chi _{\varepsilon
}}\right) _{0}\right) \right] & \\
\qquad +Pal_{+,\varepsilon }\left[ \phi _{+}^{\left( l/\varepsilon -T\left(
\varepsilon \right) \right) }\left( \tau \right) \left( \xi
_{+}^{\varepsilon }-v_{+}\right) \right], \hfill & l/\varepsilon \leq \tau
\leq l/\varepsilon +T\left( \varepsilon \right) , \\
\xi _{\chi _{\varepsilon }}, \hfill & -l/\varepsilon \leq \tau \leq
l/\varepsilon , \\
Pal_{\chi ,\varepsilon }\left[ \left( \xi _{\chi _{\varepsilon }}\right)
_{0}+\phi _{0}^{l/\varepsilon +T\left( \varepsilon \right) }\left( \tau
\right) \left( \xi _{\chi _{\varepsilon }}-\left( \xi _{\chi _{\varepsilon
}}\right) _{0}\right) \right] & \\
\qquad +Pal_{-,\varepsilon }\left[ \phi _{-}^{\left( l/\varepsilon -T\left(
\varepsilon \right) \right) }\left( \tau \right) \left( \xi
_{-}^{\varepsilon }-v_{-}\right) \right] , \hfill & \ -l/\varepsilon
-T\left( \varepsilon \right) \leq \tau \leq -l/\varepsilon , \\
\xi _{-}^{\varepsilon }, \hfill & \tau \leq -l/\varepsilon -T\left(
\varepsilon \right) ,%
\end{cases}%
\end{eqnarray*}%
where
\begin{eqnarray*}
\xi _{\chi _{\varepsilon }} &=&Q_{\chi _{\varepsilon }}\circ \left( \kappa
_{0}^{\varepsilon }\eta \right) , \\
~\left( \xi _{\chi _{\varepsilon }}\right) _{0} &=&\int_{S^{1}}\xi _{\chi
_{\varepsilon }}dt, \\
\xi _{\pm }^{\varepsilon } &=&Q_{\pm }(\eta _{\pm }^{\varepsilon })\circ
I_{\pm (-\tau (\varepsilon ))}
\end{eqnarray*}%
and $\mu $ is determined by \eqref{eq:defmu}.
Here we recall from \eqref{Vpm} that $v_{\pm }$ is defined as the \emph{%
cut-off vector field} of the constant vector field extending the vector $\xi
_{\pm }\left( o_{\pm }\right) $,%
\begin{equation*}
v_{\pm }:=\varphi _{\pm }\left( \tau ,t\right) Pal_{\pm }\left( \tau
,t\right) \left( \xi _{\pm }\left( o_{\pm }\right) \right) ,
\end{equation*}%
$\varphi _{\pm }\left( \tau ,t\right) $ is a cut-off function, $\varphi
_{\pm }\left( \tau ,t\right) =1$ near $o_{\pm }$ and $\varphi _{\pm }\left(
\tau ,t\right) =0$ outside the cylindrical neighborhood of $o_{\pm }$, $%
Pal_{\pm }\left( \tau ,t\right) $ is the parallel transport along the
shortest geodesics from $u_{\pm }\left( o_{\pm }\right) $ to $u_{\pm }\left(
\tau ,t\right) $, and the $Pal_{\chi ,\varepsilon }$ and $Pal_{\pm
,\varepsilon }$ are parallel transports from $\chi $ and $u_{\pm }$ to the
corresponding points on $u_{app}^{\varepsilon }$ along the shortest
geodesics respectively.
The interpolation happens on the regions
\begin{equation*}
\pm \lbrack l/\varepsilon -T\left( \varepsilon \right) ,l/\varepsilon
-T\left( \varepsilon \right) +1]\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ and }\pm \left[ l/\varepsilon
+T\left( \varepsilon \right) ,l/\varepsilon +T\left( \varepsilon \right) +1%
\right] ,
\end{equation*}%
which avoids the peak $\tau =\pm l/\varepsilon $ of the weighting function $%
\beta _{\delta ,\varepsilon }$. Here we choose $T\left( \varepsilon \right)
>0$ so that as $\varepsilon \rightarrow 0$, it behaves as
\begin{equation*}
T\left( \varepsilon \right) \rightarrow \infty,\, \, \varepsilon T\left(
\varepsilon \right) \rightarrow 0,\, T\left( \varepsilon \right) <\frac{p-1}{%
\delta }S\left( \varepsilon \right)
\end{equation*}%
For example, we can take and fix
\begin{equation} \label{eq:Te}
T\left( \varepsilon \right) =\frac{1}{3}\frac{p-1}{\delta }S(\varepsilon )
\end{equation}
henceforth.
We note that in this construction we have $\xi _{\chi _{\varepsilon }}$
defined for all $\tau \in \mathbb{R}$, solving $D_{\varepsilon }\xi _{\chi
_{\varepsilon }}=\kappa _{0}^{\varepsilon }\eta $ which is equivalent to the
first order linear ODE
\begin{equation*}
\frac{\partial \xi }{\partial \tau }+J_{0}\frac{\partial \xi }{\partial t}%
+\varepsilon {\Greekmath 0272} _{\xi }\,\RIfM@\expandafter\text@\else\expandafter\mbox\fi{grad}\!f=\kappa _{0}^{\varepsilon }\eta
\end{equation*}%
on $\mathbb{R\times }S^{1}$, not just on $\left[ -l/\varepsilon
,l/\varepsilon \right] $. Since $\kappa _{0}^{\varepsilon }\eta \in L^{p}$, $%
\xi _{\chi _{\varepsilon }}$ lies in $W^{1,p}$ and in particular is
continuous on $\mathbb{R\times }S^{1}$. Therefore, considering the
evaluation of $\left( \xi _{\chi _{\varepsilon }}\right) _{0}$ at $\left(
\pm \left( l/\varepsilon +T\left( \varepsilon \right) \right) \right) $
makes sense. Thinking $\xi _{\chi _{\varepsilon }}$this way is to avoid its
cut-off too close to $\tau =\pm l/\varepsilon $, the peak of the weighting
function. Plus, later we will show the $W^{1,p}$ norm of $\xi _{\chi
_{\varepsilon }}$on $\left[ -l/\varepsilon ,l/\varepsilon \right] \mathbb{%
\times }S^{1}$ controls its $W^{1,p}$ norm on the whole $\mathbb{R\times }%
S^{1}$.
\begin{rem}
In the construction of $\left( Q_{para}^{\varepsilon }\right) _{actual}$,
since the interpolation happens in a $C\varepsilon $ radius ball around $%
p_{\pm }$,we can use the connection from the constant Euclidean metric $%
g_{p_{\pm }}$ on the Weinstein neighborhood around $p_{\pm }$ to identify
different tangent spaces such that the vector sum $\pm $ makes sense without
parallel transport. We call this approximate right inverse as $%
Q_{para}^{app;\varepsilon }$, which is simpler in exposition and estimates:%
\begin{eqnarray*}
\xi _{\varepsilon } =
\begin{cases}
\xi _{+}^{\varepsilon }, \hfill & l/\varepsilon +T\left( \varepsilon \right)
\leq \tau , \\
\left( \xi _{\chi _{\varepsilon }}\right) _{0}+\phi _{0}^{l/\varepsilon
+T\left( \varepsilon \right) }\left( \tau \right) \left( \xi _{\chi
_{\varepsilon }}-\left( \xi _{\chi _{\varepsilon }}\right) _{0}\right) & \\
\qquad +\phi _{+}^{\left( l/\varepsilon -T\left( \varepsilon \right) \right)
}\left( \tau \right) \left( \xi _{+}^{\varepsilon }-v_{+}\right), \hfill &
l/\varepsilon \leq \tau \leq l/\varepsilon +T\left( \varepsilon \right) , \\
\xi _{\chi _{\varepsilon }}, \hfill & -l/\varepsilon \leq \tau \leq
l/\varepsilon , \\
\left( \xi _{\chi _{\varepsilon }}\right) _{0}+\phi _{0}^{l/\varepsilon
+T\left( \varepsilon \right) }\left( \tau \right) \left( \xi _{\chi
_{\varepsilon }}-\left( \xi _{\chi _{\varepsilon }}\right) _{0}\right) & \\
\qquad + \phi _{-}^{\left( l/\varepsilon -T\left( \varepsilon \right)
\right) }\left( \tau \right) \left( \xi _{-}^{\varepsilon }-v_{-}\right) ,
\hfill & \ -l/\varepsilon -T\left( \varepsilon \right) \leq \tau \leq
-l/\varepsilon , \\
\xi _{-}^{\varepsilon }, \hfill & \tau \leq -l/\varepsilon -T\left(
\varepsilon \right) ,%
\end{cases}%
\end{eqnarray*}
The interpolation in $\left( Q_{para}^{app;\varepsilon }\right) _{actual}$
uses the parallel transport from the nonconstant metric $g$ inside the $%
C\varepsilon $ radius ball, but the difference is a smooth tensor of order $%
C\varepsilon $, first in $C^{1}$ pointwise then in operator norm, namely%
\begin{equation*}
\left( Q_{para}^{app;\varepsilon }\right)
_{actual}=Q_{para}^{app;\varepsilon }+H_{\varepsilon }
\end{equation*}%
with the operator norm $\left\Vert H_{\varepsilon }\right\Vert \leq
C\varepsilon $. This will not effect the approximate inverse. Therefore it
is suffice to estimate the above (simpler) $Q_{para}^{app;\varepsilon }$ in
the next section.
\end{rem}
\section{Estimate of the combined right inverse}
\begin{prop}
\label{prop:uniform-inverse-bound} $Q_{para}^{app;\varepsilon }$ has a
uniform bound independent on $\varepsilon $. More precisely, there exists a
uniform constant $C$ such that for all $\eta $,
\begin{equation*}
\left\Vert Q_{para}^{app;\varepsilon }\eta \right\Vert _{\varepsilon }\leq
C\left( \left\Vert \xi _{\chi _{\varepsilon }}\right\Vert _{W_{\rho
_{\varepsilon }}^{1,p}\left( \left[ -l/\varepsilon ,l/\varepsilon \right]
\times S^{1}\right) }+\left\Vert \xi _{\pm }\right\Vert _{W_{\alpha
}^{1,p}\left( \Sigma _{\pm }\right) }\right) \leq C\left\Vert \eta
\right\Vert _{\varepsilon }.
\end{equation*}
\end{prop}
\begin{proof}
The proof is by splitting the Banach norm on $u_{app}^{\varepsilon }$ to
those associated to $u_{\chi _{\varepsilon }}$ and $u_{\pm }$. We also need
to estimates on
\begin{enumerate}
\item uniform convergence of $\xi _{\pm }\left( \tau \right) $ and $\xi
_{\chi _{\varepsilon }}\left( \tau \right) $ into $v_{\pm }$, because our
Banach norms involve first taking out the Morse-Bott variation,
\item and the matching constant $\mu$
\end{enumerate}
in terms of the norm $\Vert \eta \Vert _{\varepsilon }$. In the
interpolation region $Q_{para}^{app;\varepsilon }\eta =\left( \xi
_{\varepsilon },\mu \right) \,\ $ is given by
\begin{equation*}
\xi _{\varepsilon }=\left( \xi _{\chi _{\varepsilon }}\right) _{0}+\phi
_{0}^{l/\varepsilon +T\left( \varepsilon \right) }\left( \tau \right) \left(
\xi _{\chi _{\varepsilon }}-\left( \xi _{\chi _{\varepsilon }}\right)
_{0}\right) +\phi _{+}^{\left( l/\varepsilon -T\left( \varepsilon \right)
\right) }\left( \tau \right) \left( \xi _{+}^{\varepsilon }-v_{+}\right) ,
\end{equation*}%
when $\tau \geq 0$ and $\mu $ by solving the matching condition
\begin{equation*}
\left( \xi _{\chi _{\varepsilon }}\right) _{0}\left( l/\varepsilon \right)
=v_{+}-\varepsilon \frac{\mu }{l}{\Greekmath 0272} f\left( p_{+}\right) .
\end{equation*}
We also recall the definition of the Banach norm
\begin{equation*}
\left\Vert \left( \xi ,\mu \right) \right\Vert _{\varepsilon }=\left\Vert
\xi \right\Vert _{\varepsilon }+\left\vert \mu \right\vert =\left\Vert
\widetilde{\xi }\right\Vert _{W_{\beta _{\delta ,\varepsilon }}^{1,p}\left(
\mathbb{R\times }S^{1}\right) }+\left\Vert \xi _{0}\left( \tau \right)
\right\Vert _{W_{\varepsilon }^{1,p}\left( \left[ -l/\varepsilon
,l/\varepsilon \right] \right) }+\left\vert \xi _{0}\left( \pm l/\varepsilon
\right) \right\vert +\left\vert \mu \right\vert.
\end{equation*}
With these preparation, we are now ready to carry out the estimate of the
norm $\Vert Q_{para}^{app;\varepsilon }(\eta )\Vert _{\varepsilon }$. For
the $\mu $ component, since $\left( \left( \xi _{\chi _{\varepsilon
}}\right) _{0},\mu \right) =\left( Q_{\chi _{\varepsilon }}^{para}\right)
_{0}\circ \kappa _{0}^{\varepsilon }\eta $ and $\left( Q_{\chi _{\varepsilon
}}^{para}\right) _{0}$ has uniform bound (one can use $\left( \ref%
{e-eta-relation}\right) $ to see the $\left( Q_{\chi _{\varepsilon
}}^{para}\right) _{0}$ has the same operator bound as $\left( Q_{\chi
}^{para}\right) _{0}$ for all $\varepsilon $), we have%
\begin{equation*}
\left\vert \mu \right\vert \leq C\left\Vert \left( \kappa _{0}^{\varepsilon
}\eta \right) _{0}\right\Vert _{L_{\varepsilon }^{p}\left( \left[
-l/\varepsilon ,l/\varepsilon \right] \right) }\leq C\left\Vert \eta
\right\Vert _{\varepsilon }.
\end{equation*}
For the $\left\Vert \xi _{\varepsilon }\right\Vert _{\varepsilon }$
component, we consider two regions separately, one on $[0,l/\varepsilon
]\times S^{1}$ and the other on $[l/\varepsilon ,\tau (\varepsilon )]\times
S^{1}=[l/\varepsilon ,l/\varepsilon +\frac{p-1}{\delta }S(\varepsilon
)]\times S^{1}$.
On $[0,l/\varepsilon ]\times S^{1}$. Since $\phi _{0}^{l/\varepsilon
+T\left( \varepsilon \right) }\left( \tau \right) \equiv 1$, we have
\begin{eqnarray*}
\xi _{\varepsilon } &=&\left( \xi _{\chi _{\varepsilon }}\right) _{0}+\left(
\xi _{\chi _{\varepsilon }}-\left( \xi _{\chi _{\varepsilon }}\right)
_{0}\right) +\phi _{+}^{l/\varepsilon -T\left( \varepsilon \right) }\left(
\tau \right) \left( \xi _{+}^{\varepsilon }-v_{+}\right) \\
&=&\xi _{\chi _{\varepsilon }}+\phi _{+}^{\left( l/\varepsilon -T\left(
\varepsilon \right) \right) }\left( \tau \right) \left( \xi
_{+}^{\varepsilon }-v_{+}\right) .
\end{eqnarray*}%
Therefore for the $0$-mode $\left( \xi _{\varepsilon }\right) _{0}$ we have
\begin{eqnarray*}
\left\Vert \left( \xi _{\varepsilon }\right) _{0}\right\Vert
_{W_{\varepsilon }^{1,p}\left[ 0,l/\varepsilon \right] } &=&\left\Vert
\left( \xi _{\chi _{\varepsilon }}\right) _{0}+\phi _{+}^{l/\varepsilon
-T\left( \varepsilon \right) }\left( \tau \right) \left( \xi
_{+}^{\varepsilon }-v_{+}\right) _{0}\right\Vert _{W_{\varepsilon
}^{1,p}\left( \left[ 0,l/\varepsilon \right] \right) } \\
&\leq &\left\Vert \left( \xi _{\chi _{\varepsilon }}\right) _{0}\right\Vert
_{W_{\varepsilon }^{1,p}\left( \left[ 0,l/\varepsilon \right] \right)
}+\left\Vert \phi _{+}^{l/\varepsilon -T\left( \varepsilon \right) }\left(
\tau \right) \left( \xi _{+}^{\varepsilon }-v_{+}\right) _{0}\right\Vert
_{W_{\varepsilon }^{1,p}\left( \left[ 0,l/\varepsilon \right] \right) } \\
&\leq &\left\Vert \xi _{\chi _{\varepsilon }}\right\Vert _{W_{\rho
_{\varepsilon }}^{1,p}\left( \mathbb{R\times }S^{1}\right) }+C\left\Vert
\left( \xi _{+}^{\varepsilon }-v_{+}\right) _{0}\cdot \varepsilon ^{\frac{1-p%
}{p}}\right\Vert _{W^{1,p}\left( \left[ 0,l/\varepsilon \right] \times
S^{1}\right) } \\
&\leq &\left\Vert \xi _{\chi _{\varepsilon }}\right\Vert _{W_{\rho
_{\varepsilon }}^{1,p}\left( \mathbb{R\times }S^{1}\right) }+C\left\Vert \xi
_{+}\right\Vert _{W_{\alpha }^{1,p}\left( \left[ 0,l/\varepsilon \right]
\times S^{1}\right) },
\end{eqnarray*}%
where the third inequality is because $\left\Vert \cdot \right\Vert
_{W_{\varepsilon }^{1,p}}$ is a component of $\left\Vert \cdot \right\Vert
_{W_{\rho _{\varepsilon }}^{1,p}}$, and the last inequality holds because
for $\tau \in $ $\left[ 0,l/\varepsilon \right] \,$, the exponential weight
\begin{equation*}
e^{2\pi \delta (-\tau +\tau (\varepsilon ))}=e^{2\pi \delta \left( -\tau
+\tau (\varepsilon )\right) }\geq e^{2\pi \delta \left( \frac{p-1}{\delta }%
S\left( \varepsilon \right) \right) }=\left( 1+l/\varepsilon \right)
^{p-1}\geq C\varepsilon ^{1-p},
\end{equation*}%
then
\begin{eqnarray*}
&{}&\left\Vert \left( \xi _{+}^{\varepsilon }-v_{+}\right) _{0}\cdot
\varepsilon ^{\frac{1-p}{p}}\right\Vert _{W^{1,p}\left( \left[
0,l/\varepsilon \right] \times S^{1}\right) } \\
&\leq &C\left\Vert \left( \xi _{+}^{\varepsilon }-v_{+}\right) _{0}\cdot e^{%
\frac{2\pi \delta }{p}\left( -\tau +\tau (\varepsilon )\right) }\right\Vert
_{W^{1,p}\left( \left[ 0,l/\varepsilon \right] \times S^{1}\right) } \\
&=&C\left\Vert \left( \xi _{+}-v_{+}\right) _{0}\cdot e^{\frac{2\pi \delta }{%
p}\left\vert \tau \right\vert }\right\Vert _{W^{1,p}\left( \left[ -\tau
(\varepsilon ),-\frac{p-1}{\delta }S(\varepsilon )\right] \times
S^{1}\right) } \\
&\leq &C\left\Vert \xi _{+}\right\Vert _{W_{\alpha }^{1,p}\left( \Sigma
_{+}\right) }.
\end{eqnarray*}%
For the higher mode $\widetilde{\xi _{\varepsilon }}$, we have
\begin{eqnarray*}
\Vert \widetilde{\xi _{\varepsilon }}\Vert _{W_{\rho _{\varepsilon
}}^{1,p}([0,l/\varepsilon ]\times S^{1})} &=&\left\Vert \widetilde{\xi
_{\chi _{\varepsilon }}}+\phi _{+}^{l/\varepsilon -T\left( \varepsilon
\right) }\left( \tau \right) \left( \widetilde{\xi _{+}^{\varepsilon }-v_{+}}%
\right) \right\Vert _{W_{\rho _{\varepsilon }}^{1,p}([0,l/\varepsilon
]\times S^{1})} \\
&\leq &\left\Vert \widetilde{\xi _{\chi _{\varepsilon }}}\right\Vert
_{W_{\rho _{\varepsilon }}^{1,p}([0,l/\varepsilon ]\times
S^{1})}+C\left\Vert \widetilde{\left( \xi _{+}^{\varepsilon }-v_{+}\right) }%
\right\Vert _{W_{\rho _{\varepsilon }}^{1,p}([0,l/\varepsilon ]\times S^{1})}
\\
&\leq &C\left( \left\Vert \xi _{\chi _{\varepsilon }}\right\Vert _{W_{\rho
_{\varepsilon }}^{1,p}\left( \mathbb{R\times }S^{1}\right) }+\left\Vert \xi
_{+}\right\Vert _{W_{\alpha }^{1,p}\left( \Sigma _{+}\right) }\right) .
\end{eqnarray*}%
The last inequality holds because the exponential weight $e^{2\pi \delta
\left( -\tau +\tau (\varepsilon )\right) }$ of $W_{\alpha }^{1,p}$ is bigger
than the power weight $\varepsilon ^{1-p+\delta }\left( 1+\left\vert \tau
\right\vert \right) ^{\delta }$ of $W_{\rho _{\varepsilon }}^{1,p}$ on $%
[0,l/\varepsilon ]\times S^{1}$, by
\begin{equation*}
e^{2\pi \delta \left( -\tau +\tau (\varepsilon )\right) }\geq \left(
1+l/\varepsilon \right) ^{p-1}=C\varepsilon ^{1-p+\delta }\left(
1+l/\varepsilon \right) ^{\delta }\geq C\varepsilon ^{1-p+\delta }\left(
1+\left\vert \tau \right\vert \right) ^{\delta }.
\end{equation*}%
\ In the last inequality we also used that the projection operators
\begin{equation*}
P_{0}:\xi \rightarrow \xi _{0}\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ and }\widetilde{P}:\xi \rightarrow
\widetilde{\xi }
\end{equation*}%
from $W_{\delta }^{1,p}$ to $W_{\delta }^{1,p}$ are uniformly bounded
operators by H\"{o}lder inequality on $S^{1}$
\begin{eqnarray*}
\int_{I}\left( \left\vert \xi _{0}\right\vert ^{p}+\left\vert {\Greekmath 0272} _{\tau
}\xi _{0}\right\vert ^{p}\right) e^{2\pi \delta \left\vert \tau \right\vert
}d\tau &=&\int_{I}\left( \left\vert \int_{S^{1}}\xi dt\right\vert
^{p}+\left\vert \int_{S^{1}}{\Greekmath 0272} _{\tau }\xi dt\right\vert ^{p}\right)
e^{2\pi \delta \left\vert \tau \right\vert }d\tau \\
&\leq &\int_{I}\left( \int_{S^{1}}\left\vert \xi \right\vert
^{p}dt+\int_{S^{1}}\left\vert {\Greekmath 0272} \xi \right\vert ^{p}dt\right) e^{2\pi
\delta \left\vert \tau \right\vert }d\tau
\end{eqnarray*}%
for any interval $I$.
On $[l/\varepsilon ,\tau (\varepsilon )]\times S^{1}.$ On this region the
contribution to $\left\Vert \xi _{\varepsilon }\right\Vert _{\varepsilon }$
is
\begin{equation*}
\left\vert \left( \xi _{\varepsilon }\right) _{0}\left( l/\varepsilon
\right) \right\vert +\left\Vert \xi _{\varepsilon }\left( \tau ,t\right)
-\left( \xi _{\varepsilon }\right) _{0}\left( l/\varepsilon \right)
\right\Vert _{W_{\beta _{\delta ,\varepsilon }}^{1,p}\left( [l/\varepsilon
,\tau (\varepsilon )]\times S^{1}\right) .}
\end{equation*}%
>From the matching condition \eqref{eq:matching}, we have
\begin{equation*}
\left( \xi _{\varepsilon }\right) _{0}\left( l/\varepsilon \right) =\left(
\xi _{\chi _{\varepsilon }}\right) _{0}\left( l/\varepsilon \right) +\left(
\xi _{+}\right) _{0}\left( l/\varepsilon \right) -v_{+}=\left( \xi
_{+}\right) _{0}\left( l/\varepsilon \right) -\varepsilon \frac{\mu }{l}%
{\Greekmath 0272} f\left( p_{+}\right) .
\end{equation*}%
Therefore for the term $\left( \xi _{\varepsilon }\right) _{0}\left(
l/\varepsilon \right) $,
\begin{eqnarray*}
\left\vert \left( \xi _{\varepsilon }\right) _{0}\left( l/\varepsilon
\right) \right\vert &\leq &\left\vert \left( \xi _{+}\right) _{0}\left(
l/\varepsilon \right) \right\vert +\left\vert \frac{\varepsilon }{l}{\Greekmath 0272}
f\left( p_{+}\right) \right\vert \left\vert \mu \right\vert \\
&\leq &C\left\Vert \xi _{+}\right\Vert _{W_{\alpha }^{1,p}\left( \Sigma
_{+}\right) }+C\varepsilon \left\vert \mu \right\vert \\
&\leq &C\left\Vert \eta _{+}\right\Vert _{L_{\alpha }^{p}\left( \Sigma
_{+}\right) }+C\varepsilon \left\Vert \eta _{\varepsilon }\right\Vert
_{\varepsilon }\leq C\left\Vert \eta \right\Vert _{\varepsilon }.
\end{eqnarray*}%
For $\widetilde{\xi _{\varepsilon }}=\xi _{\varepsilon }\left( \tau
,t\right) -\left( \xi _{\varepsilon }\right) _{0}\left( l/\varepsilon
\right) $,
\begin{eqnarray}
\xi _{\varepsilon }\left( \tau ,t\right) -\left( \xi _{\varepsilon }\right)
_{0}\left( l/\varepsilon \right) &=&\left( \xi _{\chi _{\varepsilon
}}\right) _{0}\left( \tau \right) +\phi _{0}^{l/\varepsilon +T\left(
\varepsilon \right) }\left( \tau \right) \left( \xi _{\chi _{\varepsilon
}}-\left( \xi _{\chi _{\varepsilon }}\right) _{0}\right) \notag \\
&{}&+\left( \xi _{+}^{\varepsilon }\left( \tau ,t\right) -v_{+}\right)
-\left( \xi _{+}^{\varepsilon }\right) _{0}\left( l/\varepsilon \right)
+\varepsilon \frac{\mu }{l}{\Greekmath 0272} f\left( p_{+}\right) \notag \\
&=&\left[ \xi _{+}^{\varepsilon }\left( \tau ,t\right) -\left( \xi
_{+}^{\varepsilon }\right) _{0}\left( l/\varepsilon \right) \right] +\phi
_{0}^{l/\varepsilon +T\left( \varepsilon \right) }\left( \tau \right) \left(
\xi _{\chi _{\varepsilon }}-\left( \xi _{\chi _{\varepsilon }}\right)
_{0}\right) \notag \\
&&+\left[ \left( \xi _{\chi _{\varepsilon }}\right) _{0}\left( \tau \right)
-\left( \xi _{\chi _{\varepsilon }}\right) _{0}\left( l/\varepsilon \right) %
\right] , \label{difference-all}
\end{eqnarray}%
where in the last row we have used \eqref{eq:matching}. For the first term,
same as the computation $\left( 10.55\right) \sim \left( 10.60\right) $ in
[OZ1] Lemma 10.10, we have
\begin{eqnarray}
&{}&\int_{S^{1}}\int_{l/\varepsilon }^{\tau (\varepsilon )}\left\vert \xi
_{+}^{\varepsilon }\left( \tau ,t\right) -\left( \xi _{+}^{\varepsilon
}\right) _{0}\left( l/\varepsilon \right) \right\vert ^{p} \notag \\
&{}&+\left\vert {\Greekmath 0272} \left( \xi _{+}^{\varepsilon }\left( \tau ,t\right)
-\left( \xi _{+}^{\varepsilon }\right) _{0}\left( l/\varepsilon \right)
\right) \right\vert ^{p}e^{2\pi \delta \left( -\tau +\tau (\varepsilon
)\right) }d\tau dt\leq C\left\Vert \xi _{+}\right\Vert _{W_{\alpha
}^{1,p}\left( \Sigma _{+}\right) }^{p}. \label{difference-1}
\end{eqnarray}%
For the second term, by comparing the exponential weight and power weight on
$\left[ l/\varepsilon ,l/\varepsilon +\frac{p-1}{\delta }S\left( \varepsilon
\right) \right] $
\begin{equation}
e^{2\pi \delta \left( -\tau +\tau (\varepsilon )\right) }=\left(
1+l/\varepsilon \right) ^{p-1}e^{-2\pi \left( \tau -l/\varepsilon \right)
}\leq C\varepsilon ^{1-p}\left( 1+\left\vert \tau \right\vert \right)
^{\delta }, \label{eq:edtaue}
\end{equation}%
we have
\begin{eqnarray}
&&\left\Vert \phi _{0}^{l/\varepsilon +T\left( \varepsilon \right) }\left(
\tau \right) \left( \xi _{\chi _{\varepsilon }}-\left( \xi _{\chi
_{\varepsilon }}\right) _{0}\right) e^{\frac{2\pi \delta }{p}\left( -\tau
+\tau (\varepsilon )\right) }\right\Vert _{W^{1,p}\left( \left[
l/\varepsilon ,\tau (\varepsilon )\right] \times S^{1}\right) } \notag \\
&\leq &C\left\Vert \phi _{0}^{l/\varepsilon +T\left( \varepsilon \right)
}\left( \tau \right) \left( \xi _{\chi _{\varepsilon }}-\left( \xi _{\chi
_{\varepsilon }}\right) _{0}\right) \left( \varepsilon ^{1-p}\left(
1+\left\vert \tau \right\vert \right) ^{\delta }\right) ^{\frac{1}{p}%
}\right\Vert _{W^{1,p}\left( \left[ l/\varepsilon ,\tau (\varepsilon )\right]
\times S^{1}\right) } \notag \\
&=&C\left\Vert \xi _{\chi _{\varepsilon }}-\left( \xi _{\chi _{\varepsilon
}}\right) _{0}\right\Vert _{W_{\rho _{\varepsilon }}^{1,p}\left( \mathbb{%
R\times }S^{1}\right) } \notag \\
&\leq &C\left\Vert \xi _{\chi _{\varepsilon }}\right\Vert _{W_{\rho
_{\varepsilon }}^{1,p}\left( \mathbb{R\times }S^{1}\right) }.
\label{difference-2}
\end{eqnarray}%
For the third term, note that from $\left( \ref{Schauder-Convergence}\right)
$
\begin{equation*}
\left\vert \left( \xi _{\chi _{\varepsilon }}\right) _{0}\left( \tau \right)
-\left( \xi _{\chi _{\varepsilon }}\right) _{0}\left( l/\varepsilon \right)
\right\vert \leq \left\vert \varepsilon \tau \pm l\right\vert ^{\gamma
}\left\Vert \left( \xi _{\chi _{\varepsilon }}\right) _{0}\right\Vert
_{W_{\varepsilon }^{1,p}\left( [l/\varepsilon ,\tau (\varepsilon )]\right) }
\end{equation*}%
where $\gamma =1-\frac{1}{p}$, we have
\begin{eqnarray*}
&&\int_{S^{1}}\int_{l/\varepsilon }^{\tau (\varepsilon )}\left\vert \left(
\xi _{\chi _{\varepsilon }}\right) _{0}\left( \tau \right) -\left( \xi
_{\chi _{\varepsilon }}\right) _{0}\left( l/\varepsilon \right) \right\vert
^{p}e^{2\pi \delta \left( -\tau +\tau (\varepsilon )\right) }d\tau dt \\
&\leq &C\varepsilon ^{p\gamma }e^{2\pi \left( p-1\right) S(\varepsilon
)}\int_{S^{1}}\int_{0}^{\frac{p-1}{\delta }S(\varepsilon )}s^{p\gamma
}e^{-2\pi \delta s}dsdt\cdot \left\Vert \left( \xi _{\chi _{\varepsilon
}}\right) _{0}\right\Vert _{W_{\varepsilon }^{1,p}\left( [l/\varepsilon
,\tau (\varepsilon )]\right) }^{p} \\
&\leq &C\left( p,l\right) \left\Vert \left( \xi _{\chi _{\varepsilon
}}\right) _{0}\right\Vert _{W_{\varepsilon }^{1,p}\left( [l/\varepsilon
,\tau (\varepsilon )]\right) }^{p},
\end{eqnarray*}%
where $s=\tau -\frac{l}{\varepsilon }$ and in the last inequality we have
used that the integral $\int_{0}^{\infty }s^{p\gamma }e^{-2\pi \delta s}ds$
converges and $\varepsilon ^{p\gamma }e^{2\pi \left( p-1\right) S\left(
\varepsilon \right) }\approx l^{p-1}$. From linearized gradient operator $%
\frac{\partial }{\partial \tau }+A_{\varepsilon }(\tau )$ we also have
\begin{equation*}
{\Greekmath 0272} _{\tau }\left( \left( \xi _{\chi _{\varepsilon }}\right) _{0}\left(
\tau \right) -v_{+}\right) ={\Greekmath 0272} _{\tau }\left( \xi _{\chi _{\varepsilon
}}\right) _{0}\left( \tau \right) =\left( \eta _{\chi _{\varepsilon
}}\right) _{0}-A_{\varepsilon }(\tau )\left( \xi _{\chi _{\varepsilon
}}\right) _{0}.
\end{equation*}%
Since $\left( \xi _{\chi _{\varepsilon }}\right) _{0}\left( \tau \right)
-v_{+}$ is $t$-independent, and from \eqref{eq:edtaue} for $\tau \in \lbrack
l/\varepsilon ,\tau (\varepsilon )]$ we have
\begin{eqnarray*}
&&\left\Vert {\Greekmath 0272} \left( \left( \xi _{\chi _{\varepsilon }}\right)
_{0}\left( \tau \right) -v_{+}\right) e^{-\frac{2\pi \delta }{p}\left( \tau
-\tau (\varepsilon )\right) }\right\Vert _{L^{p}\left( [l/\varepsilon ,\tau
(\varepsilon )]\right) } \\
&\leq &C\left\Vert {\Greekmath 0272} _{\tau }\left( \left( \xi _{\chi _{\varepsilon
}}\right) _{0}\left( \tau \right) -v_{+}\right) \varepsilon ^{\frac{1-p}{p}%
}\right\Vert _{L^{p}\left( [l/\varepsilon ,\tau (\varepsilon )]\right) } \\
&=&C\left\Vert {\Greekmath 0272} _{\tau }\left( \xi _{\chi _{\varepsilon }}\right)
_{0}\left( \tau \right) \right\Vert _{L_{\varepsilon }^{p}\left(
[l/\varepsilon ,\tau (\varepsilon )]\right) } \\
&\leq &C\left( \left\Vert \left( \eta _{\chi _{\varepsilon }}\right)
_{0}\right\Vert _{L_{\varepsilon }^{p}\left( [l/\varepsilon ,\tau
(\varepsilon )]\right) }+\left\Vert A_{\varepsilon }(\tau )\right\Vert
_{C^{o}}\left\Vert \left( \xi _{\chi _{\varepsilon }}\right) _{0}\right\Vert
_{L_{\varepsilon }^{p}\left( [l/\varepsilon ,\tau (\varepsilon )]\right)
}\right) \\
&\leq &C\left( \left\Vert \left( \xi _{\chi _{\varepsilon }}\right)
_{0}\right\Vert _{W_{\varepsilon }^{1,p}\left( \mathbb{R}\right)
}+\left\Vert \left( \xi _{\chi _{\varepsilon }}\right) _{0}\right\Vert
_{W_{\varepsilon }^{1,p}\left( [l/\varepsilon ,\tau (\varepsilon )]\right)
}\right) .\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ \ }
\end{eqnarray*}%
\ \ \ Note in the last inequality we must use $\left\Vert \left( \xi _{\chi
_{\varepsilon }}\right) _{0}\right\Vert _{W_{\varepsilon }^{1,p}\left(
\mathbb{R}\right) }$ instead of
\begin{equation*}
\left\Vert \left( \xi _{\chi _{\varepsilon }}\right) _{0}\right\Vert
_{W_{\varepsilon }^{1,p}\left( [l/\varepsilon ,\tau (\varepsilon )]\right) }
\end{equation*}%
since we have used the right inverse defined on whole $\mathbb{R}$.
Combining these we get%
\begin{equation}
\left\Vert \left( \left( \xi _{\chi _{\varepsilon }}\right) _{0}\left( \tau
\right) -v_{+}\right) \right\Vert _{W_{\varepsilon }^{1,p}\left(
[l/\varepsilon ,\tau (\varepsilon )]\right) }\leq C\left\Vert \left( \xi
_{\chi _{\varepsilon }}\right) _{0}\right\Vert _{W_{\varepsilon
}^{1,p}\left( \mathbb{R}\right) }. \label{difference-3}
\end{equation}%
In $\left( \ref{difference-2}\right) $ and $\left( \ref{difference-3}\right)
$, the interval $\mathbb{R}$ is bigger than $[-l/\varepsilon ,l/\varepsilon
] $ where $\xi _{\chi _{\varepsilon }}$ was originally defined. However, we
will prove the following inequality
\begin{eqnarray}
\left\Vert \left( \xi _{\chi _{\varepsilon }}\right) _{0}\right\Vert
_{W_{_{\varepsilon }}^{1,p}\left( \mathbb{R}\right) } &\leq &C\left\Vert
\left( \xi _{\chi _{\varepsilon }}\right) _{0}\right\Vert _{W_{_{\rho
_{\varepsilon }}}^{1,p}\left( [-l/\varepsilon ,l/\varepsilon ]\right) },
\notag \label{eq:xichie} \\
\left\Vert \widetilde{\xi _{\chi _{\varepsilon }}}\right\Vert _{W_{\rho
_{\varepsilon }}^{1,p}\left( \mathbb{R\times }S^{1}\right) }^{p} &\leq
&C\left\Vert \widetilde{\xi _{\chi _{\varepsilon }}}\right\Vert _{W_{\rho
_{\varepsilon }}^{1,p}\left( \left[ -l/\varepsilon ,l/\varepsilon \right]
\mathbb{\times }S^{1}\right) }^{p},
\end{eqnarray}%
later in Lemma \ref{lem:xie0} and \ref{lem:tildexie} where $C$ is
independent on $\varepsilon $. Therefore combining $\left( \ref%
{difference-all}\right) ,\left( \ref{difference-1}\right) ,\left( \ref%
{difference-2}\right) $ and $\left( \ref{difference-3}\right) ,$ we have
\begin{equation*}
\left\Vert \xi _{\varepsilon }-\left( \xi _{\varepsilon }\right) _{0}\left(
l/\varepsilon \right) \right\Vert _{W_{\beta _{\delta ,\varepsilon
}}^{1,p}\left( [l/\varepsilon ,\tau (\varepsilon )]\times S^{1}\right) }\leq
C\left( \left\Vert \xi _{+}\right\Vert _{W_{\alpha }^{1,p}\left( \Sigma
_{+}\right) }+\left\Vert \xi _{\chi _{\varepsilon }}\right\Vert _{W_{_{\rho
_{\varepsilon }}}^{1,p}\left( [-l/\varepsilon ,l/\varepsilon ]\times
S^{1}\right) }\right) .
\end{equation*}%
The estimate for $\tau \in \left[ -\tau (\varepsilon ),0\right] $ is similar.
For $\left\vert \tau \right\vert >\tau (\varepsilon )$, $\xi _{\varepsilon
}=\xi _{\pm }^{\varepsilon }$ is a shift of $\xi _{\pm }$, hence%
\begin{equation*}
\left\Vert \widetilde{\xi _{\varepsilon }}\right\Vert _{W_{\beta _{\delta
,\varepsilon }}^{1,p}\left( \pm \left[ \tau (\varepsilon ),\infty \right]
\times S^{1}\right) }=\left\Vert \widetilde{\xi _{\pm }}\right\Vert
_{W_{\beta _{\delta ,\varepsilon }}^{1,p}\left( \pm \left[ 0,\infty \right]
\times S^{1}\right) }\leq \left\Vert \xi _{\pm }\right\Vert _{W_{\alpha
}^{1,p}\left( \Sigma _{\pm }\right) }.
\end{equation*}%
Combining these we have
\begin{eqnarray*}
\left\Vert \xi _{\varepsilon }\right\Vert _{\varepsilon } &=&\left\Vert
\widetilde{\xi _{\varepsilon }}\right\Vert _{W_{\beta _{\delta ,\varepsilon
}}^{1,p}\left( \mathbb{R}\times S^{1}\right) }+\left\Vert \left( \xi
_{\varepsilon }\right) _{0}\right\Vert _{W_{\varepsilon }^{1,p}\left( \left[
-l/\varepsilon ,l/\varepsilon \right] \right) }+\left\vert \left( \xi
_{\varepsilon }\right) _{0}\left( \pm l/\varepsilon \right) \right\vert \\
&\leq &C\left( \left\Vert \xi _{+}\right\Vert _{W_{\alpha }^{1,p}\left(
\Sigma _{+}\right) }+\left\Vert \xi _{-}\right\Vert _{W_{\alpha
}^{1,p}\left( \Sigma _{-}\right) }+\left\Vert \xi _{\chi _{\varepsilon
}}\right\Vert _{W_{_{\rho _{\varepsilon }}}^{1,p}\left( [-l/\varepsilon
,l/\varepsilon ]\times S^{1}\right) }\right) +C\varepsilon \left\vert \mu
\right\vert \\
&\leq &C\left( \left\Vert \eta _{+}\right\Vert _{L_{\alpha }^{p}\left(
\Sigma _{+}\right) }+\left\Vert \eta _{-}\right\Vert _{L_{\alpha }^{p}\left(
\Sigma _{-}\right) }+\left\Vert \eta _{\chi _{\varepsilon }}\right\Vert
_{L_{_{\rho _{\varepsilon }}}^{p}\left( [-l/\varepsilon ,l/\varepsilon
]\times S^{1}\right) }\right) \\
&=&C\left( \left\Vert \kappa _{+}^{\varepsilon }\eta \right\Vert _{L_{\alpha
}^{p}\left( \Sigma _{+}\right) }+\left\Vert \kappa _{-}^{\varepsilon }\eta
\right\Vert _{L_{\alpha }^{p}\left( \Sigma _{-}\right) }+\left\Vert \kappa
_{0}^{\varepsilon }\eta \right\Vert _{L_{_{\rho _{\varepsilon }}}^{p}\left(
[-l/\varepsilon ,l/\varepsilon ]\times S^{1}\right) }\right) \\
&\leq &C\left\Vert \eta \right\Vert _{\varepsilon },
\end{eqnarray*}
Thus we have obtained
\begin{equation*}
\left\Vert Q^{app;\varepsilon}_{para}\eta \right\Vert _{\varepsilon
}=\left\Vert \left( \xi _{\varepsilon },\mu \right) \right\Vert
_{\varepsilon }\leq C\left\Vert \eta \right\Vert _{\varepsilon }.
\end{equation*}
The proposition is now proved.
\end{proof}
Finally it remains to prove \eqref{eq:xichie} which is in order.
\begin{lem}
\label{lem:xie0} We have%
\begin{equation*}
\left\Vert \left( \xi _{\chi _{\varepsilon }}\right) _{0}\right\Vert
_{W_{_{\varepsilon }}^{1,p}\left( \mathbb{R}\right) }\leq C\left\Vert \left(
\xi _{\chi _{\varepsilon }}\right) _{0}\right\Vert _{W_{_{\rho _{\varepsilon
}}}^{1,p}\left( [-l/\varepsilon ,l/\varepsilon ]\right) }.
\end{equation*}
\end{lem}
\begin{proof}
Since the right inverse $L_{_{\varepsilon }}^{p}\left( \mathbb{R}\right)
\rightarrow W_{_{\varepsilon }}^{1,p}\left( \mathbb{R}\right) ,\ \left( \eta
_{\chi _{\varepsilon }}\right) _{0}\rightarrow \ \left( \xi _{\chi
_{\varepsilon }}\right) _{0}\ \ $ is uniformly bounded, and $\left( \eta
_{\chi _{\varepsilon }}\right) _{0}$ is supported in $[-l/\varepsilon
,l/\varepsilon ]$, we have%
\begin{eqnarray*}
\left\Vert \left( \xi _{\chi _{\varepsilon }}\right) _{0}\right\Vert
_{W_{_{\varepsilon }}^{1,p}\left( \mathbb{R}\right) } &\leq &C\left\Vert
\left( \eta _{\chi _{\varepsilon }}\right) _{0}\right\Vert
_{L_{_{\varepsilon }}^{p}\left( \mathbb{R}\right) }=C\left\Vert \left( \eta
_{\chi _{\varepsilon }}\right) _{0}\right\Vert _{L_{_{\varepsilon
}}^{p}\left( [-l/\varepsilon ,l/\varepsilon ]\right) } \\
&=&C\left\Vert D^{\varepsilon }\circ \left( \xi _{\chi _{\varepsilon
}}\right) _{0}\right\Vert _{L_{_{\varepsilon }}^{p}\left( [-l/\varepsilon
,l/\varepsilon ]\right) }\leq C\left\Vert \left( \xi _{\chi _{\varepsilon
}}\right) _{0}\right\Vert _{W_{_{\varepsilon }}^{1,p}\left( [-l/\varepsilon
,l/\varepsilon ]\right) } \\
&\leq &C\left\Vert \left( \xi _{\chi _{\varepsilon }}\right) _{0}\right\Vert
_{W_{_{\rho _{\varepsilon }}}^{1,p}\left( [-l/\varepsilon ,l/\varepsilon
]\right) },
\end{eqnarray*}%
where the last inequality is because $W_{\varepsilon }^{1,p}$ is a component
of $W_{\rho _{\varepsilon }}^{1,p}$.
\end{proof}
Similarly we prove the following
\begin{lem}
\label{lem:tildexie} We have%
\begin{equation*}
\left\Vert \widetilde{\xi _{\chi _{\varepsilon }}}\right\Vert _{W_{\beta
_{\delta ,\varepsilon }}^{1,p}\left( \mathbb{R\times }S^{1}\right) }\leq
C\left\Vert {\Greekmath 0272} \widetilde{\xi _{\chi _{\varepsilon }}}\right\Vert
_{L_{\beta _{\delta ,\varepsilon }}^{p}\left( \left[ -l/\varepsilon
,l/\varepsilon \right] \mathbb{\times }S^{1}\right) .}
\end{equation*}
\end{lem}
\begin{proof}
By previous proposition, $Q_{app}^{\varepsilon }$ is a uniformly bounded
operator, so for the higher mode $\widetilde{\xi _{\chi _{\varepsilon }}}$
of $\xi _{\chi _{\varepsilon }}$, we have
\begin{eqnarray*}
\left\Vert \widetilde{\xi _{\chi _{\varepsilon }}}\right\Vert _{W_{\beta
_{\delta ,\varepsilon }}^{1,p}\left( \mathbb{R\times }S^{1}\right) } &\leq
&C\left\Vert D_{para}^{\varepsilon }\widetilde{\xi _{\chi _{\varepsilon }}}%
\right\Vert _{L_{\beta _{\delta ,\varepsilon }}^{p}\left( \mathbb{R\times }%
S^{1}\right) } \\
&=&C\left\Vert \kappa _{0}^{\varepsilon }\widetilde{\eta }_{\chi
_{\varepsilon }}\right\Vert _{L_{\beta _{\delta ,\varepsilon }}^{p}\left(
\mathbb{R\times }S^{1}\right) } \\
&=&C\left\Vert D^{\varepsilon }\widetilde{\xi _{\chi _{\varepsilon }}}%
\right\Vert _{L_{\beta _{\delta ,\varepsilon }}^{p}\left( \left[
-l/\varepsilon ,l/\varepsilon \right] \mathbb{\times }S^{1}\right) } \\
&\leq &C\left\Vert {\Greekmath 0272} \widetilde{\xi _{\chi _{\varepsilon }}}\right\Vert
_{L_{\beta _{\delta ,\varepsilon }}^{p}\left( \left[ -l/\varepsilon
,l/\varepsilon \right] \mathbb{\times }S^{1}\right) .}
\end{eqnarray*}
\end{proof}
This finishes the proof of Proposition \ref{prop:uniform-inverse-bound},
which establishes the construction of the right inverse with uniform bound
as $\varepsilon \rightarrow 0$ for $L\geq l\geq l_{0}$ for any given $%
L,l_{0}>0$.
We now justify that $Q^{app;\varepsilon}_{para}$ is indeed an approximate
right inverse.
\begin{prop}
\label{prop:approx-right-inverse}$\left\Vert \left(
D^\varepsilon_{para}\circ Q^{app;\varepsilon}_{para} -I\right) \eta
\right\Vert _{\varepsilon }<\frac{1}{2}\left\Vert \eta \right\Vert
_{\varepsilon }.$
\end{prop}
\begin{proof}
By the definition of $Q_{para}^{app;\varepsilon }$, $\left(
D_{para}^{\varepsilon }\circ Q_{para}^{app;\varepsilon }-I\right) \eta =0$
except on the intervals $\pm \left[ l/\varepsilon -T\left( \varepsilon
\right) ,l/\varepsilon +T\left( \varepsilon \right) \right] $. Let's
consider $\tau \in \left[ l/\varepsilon -T\left( \varepsilon \right)
,l/\varepsilon +T\left( \varepsilon \right) \right] $. The other interval is
the same. Recall $\widetilde{\xi _{\chi _{\varepsilon }}}{}=\xi _{\chi
_{\varepsilon }}-\left( \xi _{\chi _{\varepsilon }}\right) _{0}$, $%
\widetilde{\xi _{+}}=\xi _{+}-v_{+}$ and
\begin{equation*}
\xi _{\varepsilon }=\left( \xi _{\chi _{\varepsilon }}\right) _{0}+\phi
_{0}^{l/\varepsilon +T\left( \varepsilon \right) }\left( \tau \right)
\widetilde{\xi _{\chi _{\varepsilon }}}+\phi _{+}^{\left( l/\varepsilon
-T\left( \varepsilon \right) \right) }\left( \tau \right) \widetilde{\xi
_{+}^{\varepsilon }}.
\end{equation*}%
We compute%
\begin{equation*}
\begin{tabular}{ll}
& $D_{para}^{\varepsilon }\circ Q_{para}^{app;\varepsilon }\eta -\eta $ \\
$=$ & $D_{para}^{\varepsilon }\left( \left( \xi _{\chi _{\varepsilon
}}\right) _{0}{},\mu \right) +\phi _{0}^{l/\varepsilon +T\left( \varepsilon
\right) }D_{para}^{\varepsilon }\widetilde{\xi _{\chi _{\varepsilon }}}{}%
+\phi _{+}^{l/\varepsilon -T\left( \varepsilon \right)
}D_{para}^{\varepsilon }\widetilde{\xi _{+}^{\varepsilon }}-\left( \eta _{0}+%
\widetilde{\eta }\right) $ \\
& $+\left( \phi _{0}^{l/\varepsilon +T\left( \varepsilon \right) }\right)
^{\prime }\widetilde{\xi _{\chi _{\varepsilon }}}{}+\left( \phi
_{+}^{l/\varepsilon -T\left( \varepsilon \right) }\right) ^{\prime }%
\widetilde{\xi _{+}}$ \\
$=$ & $D_{para}^{\varepsilon }\left( \left( \xi _{\chi _{\varepsilon
}}\right) _{0},\mu \right) +\phi _{0}^{l/\varepsilon +T\left( \varepsilon
\right) }D_{para}^{\varepsilon }\widetilde{\xi _{\chi _{\varepsilon }}}{}%
+\phi _{+}^{l/\varepsilon -T\left( \varepsilon \right) }\left[
D_{u_{+}^{\varepsilon }}\overline{\partial }_{\left( K_{+}^{\varepsilon
},J_{+}^{\varepsilon }\right) }+A_{\varepsilon }\right] \left( \xi
_{+}^{\varepsilon }-v_{+}\right) $ \\
& $-\left( \eta _{0}+\widetilde{\eta }\right) +\left( \phi
_{0}^{l/\varepsilon +T\left( \varepsilon \right) }\right) ^{\prime }%
\widetilde{\xi _{\chi _{\varepsilon }}}+\left( \phi _{+}^{l/\varepsilon
-T\left( \varepsilon \right) }\right) ^{\prime }\widetilde{\xi
_{+}^{\varepsilon }}$ \\
$=$ & $D_{para}^{\varepsilon }\left( \left( \xi _{\chi _{\varepsilon
}}\right) _{0},\mu \right) +\phi _{0}^{l/\varepsilon +T\left( \varepsilon
\right) }D_{para}^{\varepsilon }\widetilde{\xi _{\chi _{\varepsilon }}}{}%
+\phi _{+}^{l/\varepsilon -T\left( \varepsilon \right)
}D_{u_{+}^{\varepsilon }}\overline{\partial }_{\left( K_{+}^{\varepsilon
},J_{+}^{\varepsilon }\right) }\xi _{+}^{\varepsilon }-\left( \eta _{0}+%
\widetilde{\eta }\right) $ \\
& $+\phi _{+}^{l/\varepsilon -T\left( \varepsilon \right) }\left(
A_{\varepsilon }\widetilde{\xi _{+}^{\varepsilon }}-D_{u_{+}^{\varepsilon }}%
\overline{\partial }_{\left( K_{+}^{\varepsilon },J_{+}^{\varepsilon
}\right) }v_{+}\right) +\left( \phi _{0}^{l/\varepsilon +T\left( \varepsilon
\right) }\right) ^{\prime }\widetilde{\xi _{\chi _{\varepsilon }}}+\left(
\phi _{+}^{l/\varepsilon -T\left( \varepsilon \right) }\right) ^{\prime }%
\widetilde{\xi _{+}^{\varepsilon }},$%
\end{tabular}%
\end{equation*}%
where in the second identity we have used the notation $\widetilde{\xi
_{+}^{\varepsilon }}=\xi _{+}^{\varepsilon }-v_{+}$. By our construction
\begin{equation*}
D_{para}^{\varepsilon }\left( \left( \xi _{\chi _{\varepsilon }}\right)
_{0},\mu \right) =\eta _{0},~~D_{para}^{\varepsilon }\widetilde{\xi _{\chi
_{\varepsilon }}}{}=\kappa _{0}^{\varepsilon }\widetilde{\eta ,}~\RIfM@\expandafter\text@\else\expandafter\mbox\fi{and}%
~D_{u_{+}^{\varepsilon }}\overline{\partial }_{\left( K_{+}^{\varepsilon
},J_{+}^{\varepsilon }\right) }\xi _{+}^{\varepsilon }=\kappa
_{+}^{\varepsilon }\widetilde{\eta }
\end{equation*}%
in $\left[ l/\varepsilon -T\left( \varepsilon \right) ,l/\varepsilon
+T\left( \varepsilon \right) \right] $. Then substitution of these into the
above and canceling out makes the second to the last row in the above
identity become $0$.
For the term $A_{\varepsilon }\left( \tau \right) \widetilde{\xi
_{+}^{\varepsilon }}$, using $\left\Vert A_{\varepsilon }\right\Vert
_{C^{1}}\leq C\varepsilon $, we obtain
\begin{eqnarray*}
\left\Vert \phi _{+}^{l/\varepsilon -T\left( \varepsilon \right) }\left(
\tau \right) A_{\varepsilon }\left( \tau \right) \widetilde{\xi
_{+}^{\varepsilon }}\right\Vert _{\varepsilon } &\leq &C\varepsilon
\left\Vert \widetilde{\xi _{+}^{\varepsilon }}\right\Vert _{\varepsilon
}\leq C\varepsilon \left\Vert \xi _{+}\right\Vert _{\varepsilon } \\
&\leq &C\varepsilon \left\Vert \eta _{+}\right\Vert _{\varepsilon }\leq
C\varepsilon \left\Vert \eta \right\Vert _{\varepsilon },
\end{eqnarray*}%
where the first inequality holds because $\xi _{+}^{\varepsilon }$ is a
shift of $\xi _{+}$, the second holds by the definition of the norm $%
\left\Vert \cdot \right\Vert _{\varepsilon }$ and the third comes from the
boundedness of the right inverse of $D_{u_{+}}\overline{\partial }_{\left(
K_{+},J_{+}\right) }$.
For $D_{u_{+}^{\varepsilon }}\overline{\partial }_{\left( K_{+}^{\varepsilon
},J_{+}^{\varepsilon }\right) }v_{+}$, by the fact that on
\begin{equation*}
\tau \in \left[ l/\varepsilon -T\left( \varepsilon \right) -\tau \left(
\varepsilon \right) ,l/\varepsilon +T\left( \varepsilon \right) -\tau \left(
\varepsilon \right) \right] \subset \lbrack -\infty ,-1),
\end{equation*}%
$J_{+}=J_{0},$ and $v_{+}$ is a vector field obtained by parallel transport
of $\xi _{+}\left( o_{+}\right) ,$
\begin{equation*}
\left\vert v_{+}\right\vert =\left\vert \xi _{+}\left( o_{+}\right)
\right\vert ,\left\vert {\Greekmath 0272} v_{+}(o_{+})\right\vert =0
\end{equation*}%
and
\begin{eqnarray*}
\left\vert D_{u_{+}}\overline{\partial }_{\left( K_{+},J_{+}\right)
}v_{+}\right\vert &=&\left\vert \left( {\Greekmath 0272} v_{+}\right) ^{0,1}+\frac{1}{2}%
J_{0}{\Greekmath 0272} _{v_{+}}J_{0}\partial u_{+}\right\vert \\
&\leq &C\left\vert du_{+}\right\vert \left\vert v_{+}\right\vert
=C\left\vert du_{+}\right\vert \left\vert \xi _{+}\left( o_{+}\right)
\right\vert \RIfM@\expandafter\text@\else\expandafter\mbox\fi{.}
\end{eqnarray*}%
Using the uniform exponential decay $\left\vert du_{+}\right\vert \leq
Ce^{2\pi \tau }$ as $\tau \rightarrow -\infty $, and noticing $\varepsilon
^{1-p}\leq e^{-2\pi \delta \left( \tau -\tau (\varepsilon )\right) }$ for $%
\tau \in $ $\left[ l/\varepsilon -T\left( \varepsilon \right) ,l/\varepsilon %
\right] $, we have
\begin{equation*}
|du_{+}(\tau )|\leq Ce^{2\pi \left( -l/\varepsilon +T(\varepsilon )\right) }
\end{equation*}%
for $\tau \in \lbrack l/\varepsilon -T(\varepsilon ),l/\varepsilon ]$, hence%
\begin{eqnarray*}
\left\Vert D_{u_{+}}\overline{\partial }_{\left( K_{+},J_{+}\right)
}v_{+}\right\Vert _{\varepsilon } &\leq &C|du_{+}(\tau )|\left\vert \xi
_{+}\left( o_{+}\right) \right\vert \\
&\leq &Ce^{2\pi \left( -l/\varepsilon +T\left( \varepsilon \right) \right)
}\left\Vert \xi _{+}\right\Vert _{W_{\alpha }^{1,p}\left( \Sigma _{+}\right)
} \\
&\leq &Ce^{2\pi \left( -l/\varepsilon +T\left( \varepsilon \right) \right)
}\left\Vert \eta \right\Vert _{L_{\alpha }^{p}\left( \Sigma _{+}\right) }.
\end{eqnarray*}%
Here we have used that $\left\vert \xi _{+}\left( o_{+}\right) \right\vert $
is part of the norm $\left\Vert \xi _{+}\right\Vert _{W_{\alpha
}^{1,p}\left( \Sigma _{+}\right) }$.
We estimate the remaining two terms%
\begin{equation*}
\left( \phi _{0}^{l/\varepsilon +T\left( \varepsilon \right) }\right)
^{\prime }\widetilde{\xi _{\chi _{\varepsilon }}},\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ \ }\left( \phi
_{+}^{l/\varepsilon -T\left( \varepsilon \right) }\right) ^{\prime }%
\widetilde{\xi _{+}^{\varepsilon }}.
\end{equation*}
For $\left( \phi _{+}^{l/\varepsilon -T\left( \varepsilon \right) }\right)
^{\prime }\left( \tau \right) \widetilde{\xi _{+}^{\varepsilon }}$, it is
supported in $\left[ l/\varepsilon -T\left( \varepsilon \right)
,l/\varepsilon -T\left( \varepsilon \right) +1\right] $, where both the
power order weight and $\varepsilon $-adiabatic weight are dominated by the
exponential weight as the following%
\begin{equation*}
\varepsilon ^{1-p+\delta }\left( 1+\left\vert \tau \right\vert \right)
^{\delta }\leq \varepsilon ^{1-p}\leq e^{-2\pi \delta T\left( \varepsilon
\right) }\cdot e^{2\pi \delta \left\vert \tau -\tau (\varepsilon
)\right\vert }.
\end{equation*}%
Therefore we do not need to separate the $0$-mode and the higher mode parts
of $\phi _{l/\varepsilon -T\left( \varepsilon \right) }^{\prime }\left( \xi
_{+}-v_{+}\right) $ and immediately see
\begin{eqnarray*}
&{}&\left\Vert \left( \phi _{+}^{l/\varepsilon -T\left( \varepsilon \right)
}\right) ^{\prime }\left( \tau \right) \widetilde{\xi _{+}^{\varepsilon }}%
\right\Vert _{\varepsilon } \\
&\leq &e^{\frac{-2\pi \delta T\left( \varepsilon \right) }{p}}\left\Vert
\left( \phi _{+}^{l/\varepsilon -T\left( \varepsilon \right) }\right)
^{\prime }\left( \tau \right) \widetilde{\xi _{+}^{\varepsilon }}\right\Vert
_{W_{\delta }^{1,p}\left[ l/\varepsilon -T\left( \varepsilon \right)
,l/\varepsilon -T\left( \varepsilon \right) +1\right] } \\
&\leq &e^{\frac{-2\pi \delta T\left( \varepsilon \right) }{p}}\left\Vert \xi
_{+}\right\Vert _{W_{\alpha }^{1,p}\left( \Sigma _{+}\right) }\leq e^{\frac{%
-2\pi \delta T\left( \varepsilon \right) }{p}}C\left\Vert \eta
_{+}\right\Vert _{L_{\alpha }^{p}\left( \Sigma _{+}\right) }\leq Ce^{\frac{%
-2\pi \delta T\left( \varepsilon \right) }{p}}\left\Vert \eta \right\Vert
_{\varepsilon }.
\end{eqnarray*}
For $\left( \phi _{+}^{l/\varepsilon -T\left( \varepsilon \right) }\right)
^{\prime }\widetilde{\xi _{+}^{\varepsilon }}$, it is supported in $\left[
l/\varepsilon +T\left( \varepsilon \right) ,l/\varepsilon +T\left(
\varepsilon \right) +1\right] $. From the Sobolev embedding, we obtain
\begin{eqnarray*}
&{}&\left\vert \widetilde{\xi _{+}^{\varepsilon }}\right\vert \\
&\leq &C\varepsilon ^{\frac{p-1-\delta }{p}}\left( 1+l/\varepsilon +T\left(
\varepsilon \right) \right) ^{-\frac{\delta }{p}}\left\Vert \widetilde{\xi
_{+}^{\varepsilon }}\right\Vert _{W_{\rho _{\varepsilon }}^{1,p}\left( \left[
l/\varepsilon +T\left( \varepsilon \right) ,l/\varepsilon +T\left(
\varepsilon \right) +1\right] \times S^{1}\right) } \\
&\leq &C\varepsilon ^{\frac{p-1-\delta }{p}}\left( \varepsilon /l\right) ^{%
\frac{\delta }{p}}\left\Vert \widetilde{\xi _{+}^{\varepsilon }}\right\Vert
_{W_{\rho _{\varepsilon }}^{1,p}\left( \left[ l/\varepsilon +T\left(
\varepsilon \right) ,l/\varepsilon +T\left( \varepsilon \right) +1\right]
\times S^{1}\right) } \\
&\leq &C\varepsilon ^{\frac{p-1-\delta }{p}}\left( \varepsilon /l\right) ^{%
\frac{\delta }{p}}\cdot C\left\Vert \widetilde{\eta }_{\chi _{\varepsilon
}}\right\Vert _{L_{\rho _{\varepsilon }}^{p}\left( \mathbb{R}\times
S^{1}\right) } \\
&\leq &C\varepsilon ^{\frac{p-1-\delta }{p}}\left( \varepsilon /l\right) ^{%
\frac{\delta }{p}}\left\Vert \eta \right\Vert _{\varepsilon },
\end{eqnarray*}%
Since $\left[ l/\varepsilon +T\left( \varepsilon \right) ,l/\varepsilon
+T\left( \varepsilon \right) +1\right] $ is contained outside $\left[
-l/\varepsilon ,l/\varepsilon \right] $, we needn't to distinguish its $0$%
-mode and higher mode in computing the weighted Sobolev norm of $\left( \phi
_{+}^{l/\varepsilon -T\left( \varepsilon \right) }\right) ^{\prime }\left(
\tau \right) \widetilde{\xi _{+}^{\varepsilon }}$. Note
\begin{equation*}
\left. \left( \phi _{+}^{l/\varepsilon -T\left( \varepsilon \right) }\right)
^{\prime }\left( \tau \right) \widetilde{\xi _{+}^{\varepsilon }}\right\vert
_{\tau =l/\varepsilon }=\left( \phi _{+}^{l/\varepsilon -T\left( \varepsilon
\right) }\right) ^{\prime }\left( \tau \right) \left( \xi _{\chi
_{\varepsilon }}-\left( \xi _{\chi _{\varepsilon }}\right) _{0}\right) \Big |%
_{\tau =l/\varepsilon }=0,
\end{equation*}%
and the weight there is
\begin{equation*}
e^{-2\pi \delta \left( \tau -l/\varepsilon -\frac{p-1}{\delta }S\left(
\varepsilon \right) \right) }\leq e^{-2\pi \delta \left( T\left( \varepsilon
\right) -\frac{p-1}{\delta }S(\varepsilon )\right) }=\left( 1+l/\varepsilon
\right) ^{p-1}e^{-2\pi \delta T\left( \varepsilon \right) },
\end{equation*}%
therefore%
\begin{eqnarray*}
\left\Vert \left( \phi _{+}^{l/\varepsilon -T\left( \varepsilon \right)
}\right) ^{\prime }\left( \tau \right) \widetilde{\xi _{+}^{\varepsilon }}%
\right\Vert _{\varepsilon } &\leq &C\varepsilon ^{\frac{p-1-\delta }{p}%
}\left( \varepsilon /l\right) ^{\frac{\delta }{p}}\left\Vert \eta
\right\Vert _{\varepsilon }\cdot \left( 1+l/\varepsilon \right) ^{\frac{p-1}{%
p}}e^{\frac{-2\pi \delta T\left( \varepsilon \right) }{p}} \\
&=&Ce^{\frac{-2\pi \delta T\left( \varepsilon \right) }{p}}\left\Vert \eta
\right\Vert _{\varepsilon }.
\end{eqnarray*}%
Combining these estimates we obtain%
\begin{eqnarray*}
\left\Vert D_{para}^{\varepsilon }\overline{\partial }_{\left(
K_{\varepsilon },J_{\varepsilon }\right) }\circ Q_{para}^{app;\varepsilon
}(\eta )-\eta \right\Vert _{\varepsilon } &\leq &C\varepsilon \left\Vert
\eta \right\Vert _{\varepsilon }+Ce^{\frac{-2\pi \delta T\left( \varepsilon
\right) }{p}}\left\Vert \eta \right\Vert _{\varepsilon }+Ce^{\frac{-2\pi
\delta T\left( \varepsilon \right) }{p}}\left\Vert \eta \right\Vert
_{\varepsilon } \\
&{}&\quad +Ce^{2\pi \left( -l/\varepsilon +T\left( \varepsilon \right)
\right) }\left\Vert \eta \right\Vert _{L_{\alpha }^{p}\left( \Sigma
_{+}\right) }
\end{eqnarray*}%
Since $T\left( \varepsilon \right) =\frac{p-1}{3\delta }S(\varepsilon
)\rightarrow \infty $ and $\varepsilon S(\varepsilon )\rightarrow 0$ as $%
\varepsilon \rightarrow 0$, we have for small enough $\varepsilon $,
\begin{equation*}
\left\Vert D_{para}^{\varepsilon }\overline{\partial }_{\left(
K_{\varepsilon },J_{\varepsilon }\right) }\circ Q_{para}^{app;\varepsilon
}(\eta )-\eta \right\Vert _{\varepsilon }\leq \frac{1}{2}\left\Vert \eta
\right\Vert _{\varepsilon }.
\end{equation*}%
The proposition follows.
\end{proof}
By the above proposition, $D_{para}^{\varepsilon }\overline{\partial }%
_{\left( K_{\varepsilon },J_{\varepsilon }\right) }\circ
Q_{para}^{app;\varepsilon }$ is invertible, and
\begin{eqnarray*}
\left\Vert \left( D_{para}^{\varepsilon }\overline{\partial }_{\left(
K_{\varepsilon },J_{\varepsilon }\right) }\circ Q_{para}^{app;\varepsilon
}\right) ^{-1}\right\Vert &\leq &\left\Vert \Sigma _{k=0}^{\infty }\left(
D_{para}^{\varepsilon }\overline{\partial }_{\left( K_{\varepsilon
},J_{\varepsilon }\right) }\circ Q_{para}^{app;\varepsilon }-id\right)
^{k}\right\Vert \\
&\leq &\Sigma _{k=0}^{\infty }\left( \frac{1}{2}\right) ^{k}=1,
\end{eqnarray*}%
so we can construct the true right inverse of \ $D_{para}^{\varepsilon }%
\overline{\partial }_{\left( K_{\varepsilon },J_{\varepsilon }\right) }$ as
\begin{equation*}
Q_{para}^{\varepsilon }:=Q_{para}^{app;\varepsilon }\circ \left(
D_{para}^{\varepsilon }\overline{\partial }_{\left( K_{\varepsilon
},J_{\varepsilon }\right) }\circ Q_{para}^{app;\varepsilon }\right) ^{-1}.
\end{equation*}%
Since $Q_{para}^{app;\varepsilon }$ is uniformly bounded, so is $%
Q_{para}^{\varepsilon }$.
\begin{rem}
If we examine the proof of the right inverse bound of $Q_{para}^{\varepsilon
}$, , we get $\left\Vert Q_{para}^{\varepsilon }\right\Vert \leq Cl^{\frac{%
p-1}{p}}$. It will increase as $l\rightarrow \infty $. But the $\overline{%
\partial }_{\left( K_{\varepsilon },J_{\varepsilon }\right) }$-error
estimate will have faster order decay $\left( Ce^{-cl}+l^{-\frac{p-1}{\delta
}}\right) l^{\frac{p-1}{p}}\varepsilon ^{\frac{1}{p}}$, and the quadratic
estimate in next section has the uniform constant $C$, so we an still apply
the implicit function theorem for all $l\geq l_{0}$.
\end{rem}
\section{\protect\bigskip Uniform quadratic estimate and implicit function
theorem\label{sec:quadratic}}
Consider the Banach spaces
\begin{eqnarray*}
X &=&\left\{ \left( \xi ,\mu \right) \,\Big|\,\xi \in \Gamma \left( \left(
u_{app}^{\varepsilon }\right) ^{\ast }TM\right) ,\mu \in T_{l/\varepsilon }%
\mathbb{R},\left\Vert \left( \xi ,\mu \right) \right\Vert _{\varepsilon
}=\left\Vert \xi \right\Vert _{\varepsilon }+\left\vert \mu \right\vert
<\infty \right\} \\
Y &=&\left\{ \eta \,\Big|\,\eta \in \Gamma \left( \left(
u_{app}^{\varepsilon }\right) ^{\ast }TM\right) \otimes \Lambda ^{0,1}\left(
\mathbb{R\times }S^{1}\right) ,\left\Vert \eta \right\Vert _{\varepsilon
}<\infty \right\}
\end{eqnarray*}%
with the Banach norms $\left\Vert \cdot \right\Vert _{\varepsilon }$ defined
for $\xi $ and $\eta $ in section \ref{sec:off-shell}. $\ $ For sections $%
\xi ,\xi ^{\prime }$ in $\left( u_{app}^{\varepsilon }\right) ^{\ast }TM$
and $\mu ,\mu ^{\prime }$ in $T_{l/\varepsilon }\mathbb{R}$, let \
\begin{equation*}
\mathcal{F}_{u_{app}^{\varepsilon }}:\Gamma \left( \left(
u_{app}^{\varepsilon }\right) ^{\ast }TM\right) \times T_{l/\varepsilon }%
\mathbb{R\rightarrow }\Gamma \left( \left( u_{app}^{\varepsilon }\right)
^{\ast }TM\right) \otimes \Lambda ^{0,1}\left( \mathbb{R\times }S^{1}\right)
\end{equation*}%
that%
\begin{equation*}
\mathcal{F}_{u_{app}^{\varepsilon }}\left( \xi ,\mu \right) \left( \tau
,t\right) =Pal_{\xi }^{-1}\left[ \overline{\partial }_{\left( K_{\varepsilon
},J_{\varepsilon }\right) }\exp _{u_{app}^{\varepsilon }}\left( \xi \right) %
\right] \left( P_{\mu }\tau ,t\right)
\end{equation*}%
where $Pal_{\xi }^{-1}$ is the parallel transport from $\exp
_{u_{app}^{\varepsilon }}\xi \left( \tau ,t\right) $ to $u_{app}^{%
\varepsilon }\left( \tau ,t\right) $ along the shortest geodesic, and the
reparametrization map
\begin{equation*}
P_{\mu }:\left( \mathbb{R};-l/\varepsilon ,l/\varepsilon \right) \rightarrow
\left( \mathbb{R};-\left( l/\varepsilon -\mu \right) ,l/\varepsilon -\mu
\right) .
\end{equation*}%
between marked real lines is
\begin{equation*}
P_{\mu }\left( \tau \right) =%
\begin{cases}
\tau +\mu & \mbox{for }\, \tau <-l/\varepsilon , \\
\frac{l}{l-\varepsilon \mu }\tau & \mbox{for }\, \left\vert \tau \right\vert
\leq l/\varepsilon , \\
\tau -\mu & \mbox{for }\, \tau >l/\varepsilon ,%
\end{cases}%
\end{equation*}%
The above definition of $\mathcal{F}_{u_{app}^{\varepsilon }}$ is slightly
imprecise since $P_{\mu }$ is only piecewise differentiable and we need to
smooth it to a sufficiently close diffeomorphism, but that can be done. Then
\begin{eqnarray*}
&&d\mathcal{F}_{u_{app}^{\varepsilon }}\left( \xi ,\mu \right) \left( \xi
^{\prime },\mu ^{\prime }\right) \left( \tau ,t\right) \\
&=&\left. \frac{d}{ds}\right\vert _{s=0}\mathcal{F}_{u_{app}^{\varepsilon
}}\left( \xi +s\xi ^{\prime },\mu +s\mu ^{\prime }\right) \left( \tau
,t\right) \\
&=&\left.\frac{d}{ds}\right\vert _{s=0}Pal_{\xi +s\xi ^{\prime }}^{-1}\left[
\overline{\partial }_{\left( K_{\varepsilon },J_{\varepsilon }\right) }\exp
_{u_{app}^{\varepsilon }}\left( \xi +s\xi ^{\prime }\right) \right] \left(
P_{\mu +s\mu ^{\prime }}\left( \tau \right) ,t\right) \\
&=&
\begin{cases}
\left.\frac{d}{ds}\right\vert _{s=0}Pal_{\xi +s\xi ^{\prime }}^{-1}\left[
\overline{\partial }_{\left( K_{\varepsilon },J_{\varepsilon }\right) }\exp
_{u_{app}^{\varepsilon }}\left( \xi +s\xi ^{\prime }\right) \right] \left(
\frac{l\tau }{l-\varepsilon \left( \mu +s\mu ^{\prime }\right) },t\right) \,
& \mbox{if }\, \left\vert \tau \right\vert \leq l/\varepsilon \\
\left.\frac{d}{ds}\right\vert _{s=0}Pal_{\xi +s\xi ^{\prime }}^{-1}\left[
\overline{\partial }_{\left( K_{\varepsilon },J_{\varepsilon }\right) }\exp
_{u_{app}^{\varepsilon }}\left( \xi +s\xi ^{\prime }\right) \right] \left(
\tau \pm \left( \mu +s\mu ^{\prime }\right) ,t\right)\, & \mbox{if }\,
\left\vert \tau \right\vert >l/\varepsilon%
\end{cases}%
\end{eqnarray*}%
and
\begin{equation*}
d\mathcal{F}_{u_{app}^{\varepsilon }}\left( 0,0\right) \left( \xi ^{\prime
},\mu ^{\prime }\right) =D_{u_{app}^{\varepsilon }}\overline{\partial }%
_{\left( K_{\varepsilon },J_{\varepsilon }\right) }\left( \xi ^{\prime },\mu
^{\prime }\right) .
\end{equation*}
\begin{prop}
\label{prop:quadratic}\bigskip For each given pair $\left( K_{\varepsilon
},J_{\varepsilon }\right) $, there exists uniform constants $C$ (depending
only on $l_{0}$ and $p$ but independent of $\varepsilon$) and $h_{0}$ such
that for all $\left( \xi ,\mu \right) $ and $\left( \xi ^{\prime },\mu
^{\prime }\right) $ in $\Gamma \left( \left( u_{app}^{\varepsilon }\right)
^{\ast }TM\right) \times T_{l/\varepsilon }\mathbb{R}$ with $0\leq
\left\Vert \left( \xi ,\mu \right) \right\Vert _{\varepsilon }\leq h_{0}\,$,
\begin{equation*}
\left\Vert d\mathcal{F}_{u_{app}^{\varepsilon }}\left( \xi ,\mu \right)
\left( \xi ^{\prime },\mu ^{\prime }\right) -D_{u_{app}^{\varepsilon }}%
\overline{\partial }_{\left( K_{\varepsilon },J_{\varepsilon }\right)
}\left( \xi ^{\prime },\mu ^{\prime }\right) \right\Vert _{\varepsilon }\leq
C\left\Vert \left( \xi ,\mu \right) \right\Vert _{\varepsilon }\left\Vert
\left( \xi ^{\prime },\mu ^{\prime }\right) \right\Vert _{\varepsilon }.
\end{equation*}
\end{prop}
\begin{proof}
We first consider the case when $\mu =\mu ^{\prime }=0$. For $\left\vert
\tau \right\vert \leq l/\varepsilon $, and any $t\in S^{1}$, we have%
\begin{eqnarray*}
\left\vert \xi \left( \tau ,t\right) \right\vert &\leq &\left\vert \xi
\left( \tau ,t\right) -\xi _{0}\left( \tau \right) \right\vert +\left\vert
\xi _{0}\left( \tau \right) \right\vert \\
&\leq &C\left\Vert \tilde{\xi}\right\Vert _{W_{\beta _{\delta ,\varepsilon
}}^{1,p}\left( \left[ -l/\varepsilon ,l/\varepsilon \right] \times
S^{1}\right) }+C\left\Vert \xi _{0}\right\Vert _{W_{\varepsilon
}^{1,p}\left( \left[ -l/\varepsilon ,l/\varepsilon \right] \times
S^{1}\right) } \\
&\leq &C\left\Vert \xi \right\Vert _{\varepsilon },
\end{eqnarray*}%
where the second row is by $W^{1,p}\hookrightarrow C^{0}$ Sobolev embedding,
and the facts that the weight $\beta _{\delta ,\varepsilon }\,$\ is nowhere
less than $1$ and the length $l\geq l_{0}>0$. \ For $l/\varepsilon
<\left\vert \tau \right\vert \leq l/\varepsilon +\frac{p-1}{\delta }S\left(
\varepsilon \right) $, again by Sobolev embedding we have%
\begin{eqnarray*}
\left\vert \xi \left( \tau ,t\right) \right\vert &\leq &\left\vert \xi
\left( \tau ,t\right) -\xi _{0}\left( o_{\pm }\right) \right\vert
+\left\vert \xi _{0}\left( o_{\pm }\right) \right\vert \\
&\leq &C\left\Vert \tilde{\xi}\right\Vert _{W_{\beta _{\delta ,\varepsilon
}}^{1,p}\left( \pm \left[ l/\varepsilon ,l/\varepsilon +\frac{p-1}{\delta }%
S\left( \varepsilon \right) \right] \times S^{1}\right) }+\left\vert \xi
_{0}\left( o_{\pm }\right) \right\vert \\
&\leq &C\left\Vert \xi \right\Vert _{\varepsilon }.
\end{eqnarray*}%
For $\left\vert \tau \right\vert >\tau (\varepsilon )$ the weight $\beta
_{\delta ,\varepsilon }\left( \tau \right) $ is $1$ so%
\begin{equation}
\left\vert \xi \left( \tau ,t\right) \right\vert \leq C\left\Vert \xi
\right\Vert _{\varepsilon } \label{e-Sobolev}
\end{equation}%
is standard Sobolev embedding. In the above the constants $C$ only depends
on $l_{0}$ and $p$. Combining these we have the uniform Sobolev constant
\begin{equation*}
C\left( l_{0},p\right) :=\sup_{\xi \neq 0}\frac{\left\vert \xi \right\vert
_{\infty }}{\left\Vert \xi \right\Vert _{\varepsilon }},
\end{equation*}%
for our Banach norm $\left\Vert \cdot \right\Vert _{\varepsilon }$ of all $%
\varepsilon $ for $\xi \in \Gamma \left( \left( u_{app}^{\varepsilon
}\right) ^{\ast }TM\right) $ on $\mathbb{R\times }S^{1}$. The point estimate
in the proof of Proposition 3.5.3 in \cite{MS} yields
\begin{equation*}
\left\vert d\mathcal{F}_{u_{app}^{\varepsilon }}\left( \xi \right) \xi
^{\prime }-D_{u_{app}^{\varepsilon }}\overline{\partial }_{\left(
J_{0},\varepsilon f\right) }\xi ^{\prime }\right\vert \leq A\left(
\left\vert du_{app}^{\varepsilon }\right\vert \left\vert \xi \right\vert
\left\vert \xi ^{\prime }\right\vert +\left\vert \xi \right\vert \left\vert
{\Greekmath 0272} \xi ^{\prime }\right\vert +\left\vert {\Greekmath 0272} \xi \right\vert
\left\vert \xi ^{\prime }\right\vert \right) .
\end{equation*}%
Taking the $p$-th power and integrating $\left\vert {\Greekmath 0272} \xi \right\vert $
and $\left\vert {\Greekmath 0272} \xi ^{\prime }\right\vert $ over $\mathbb{R\times }%
S^{1}$ with respect to the weight $\beta _{\delta ,\varepsilon }$, while
using the Sobolev inequality (from the above definition of $C\left(
l_{0},p\right) $)
\begin{equation*}
\left\vert \xi \right\vert _{\infty }\leq C\left( l_{0},p\right) \left\Vert
\xi \right\Vert _{\varepsilon }
\end{equation*}%
for the terms $\left\vert \xi \right\vert $ and $\left\vert \xi ^{\prime
}\right\vert \,$, we obtain the uniform quadratic estimate%
\begin{equation*}
\left\Vert d\mathcal{F}_{u_{app}^{\varepsilon }}\left( \xi \right) \xi
^{\prime }-D_{u_{app}^{\varepsilon }}\overline{\partial }_{\left(
J,\varepsilon f\right) }\xi ^{\prime }\right\Vert _{\varepsilon }\leq
C\left\Vert \xi \right\Vert _{\varepsilon }\left\Vert \xi ^{\prime
}\right\Vert _{\varepsilon }
\end{equation*}%
where $C$ is dependent on $l_{0},p$ but independent on $\varepsilon $.
Then we include the $\mu $ and $\mu ^{\prime }$ in the quadratic estimate.
For simplicity of notation we let
\begin{equation*}
u_{\xi }=\exp _{u_{app}^{\varepsilon }}\xi
\end{equation*}%
and write the pair $\left( u_{app}^{\varepsilon },l/\varepsilon \right) $ to
emphasize the parameter $l$ in the construction of $u_{app}^{\varepsilon }$.
Then for $\left\vert \tau \right\vert \leq l/\varepsilon $,
\begin{eqnarray}
&&\left\Vert d\mathcal{F}_{\left( u_{app}^{\varepsilon },\frac{l}{%
\varepsilon }\right) }\left( \xi ,\mu \right) \left( \xi ^{\prime },\mu
^{\prime }\right) -d\mathcal{F}_{\left( u_{app}^{\varepsilon },\frac{l}{%
\varepsilon }\right) }\left( 0,0\right) \left( \xi ^{\prime },\mu ^{\prime
}\right) \right\Vert \notag \\
&=&\left\Vert \left( d\mathcal{F}_{\left( u_{app}^{\varepsilon },\frac{l}{%
\varepsilon }-\mu \right) }\left( \xi \right) \xi ^{\prime }+\frac{\mu
^{\prime }l\varepsilon \tau }{\left( l-\varepsilon \mu \right) ^{2}}Pal_{\xi
}^{-1}\frac{\partial u_{\xi }}{\partial \tau }\right) -\left( d\mathcal{F}%
_{\left( u_{app}^{\varepsilon },\frac{l}{\varepsilon }\right) }\left(
0\right) \xi ^{\prime }+\frac{\mu ^{\prime }\varepsilon \tau }{l}\frac{%
\partial u}{\partial \tau }\right) \right\Vert \notag \\
&\leq &\left\Vert d\mathcal{F}_{\left( u_{app}^{\varepsilon },\frac{l}{%
\varepsilon }-\mu \right) }\left( \xi \right) \xi ^{\prime }-d\mathcal{F}%
_{\left( u_{app}^{\varepsilon },\frac{l}{\varepsilon }-\mu \right) }\left(
0\right) \xi ^{\prime }\right\Vert \notag \\
&&+\left\Vert d\mathcal{F}_{\left( u_{app}^{\varepsilon },\frac{l}{%
\varepsilon }-\mu \right) }\left( 0\right) \xi ^{\prime }-d\mathcal{F}%
_{\left( u_{app}^{\varepsilon },\frac{l}{\varepsilon }\right) }\left(
0\right) \xi ^{\prime }\right\Vert +\left\Vert \frac{\mu ^{\prime
}l\varepsilon \tau }{\left( l-\varepsilon \mu \right) ^{2}}Pal_{\xi }^{-1}%
\frac{\partial u_{\xi }}{\partial \tau }-\frac{\mu ^{\prime }\varepsilon
\tau }{l}\frac{\partial u}{\partial \tau }\right\Vert \notag \\
&\leq &C\left\Vert \xi \right\Vert _{\varepsilon }\left\Vert \xi ^{\prime
}\right\Vert _{\varepsilon }+C\left\vert \mu \right\vert \left\Vert \xi
^{\prime }\right\Vert _{\varepsilon }+C\left\vert \mu ^{\prime }\right\vert
\left\Vert \xi \right\Vert _{\varepsilon }, \label{quadratic-C}
\end{eqnarray}%
where in the last inequality, the first term is by the $\mu =\mu ^{\prime
}=0 $ case, the second term holds because the change of the conformal
structure on $\left[ -l/\varepsilon ,l/\varepsilon \right] \times S^{1}$ by $%
\mu $ affects $\overline{\partial }_{\left( K_{\varepsilon },J_{\varepsilon
}\right) }$ in a linear way, and the third term is by the property of
exponential map, and
\begin{equation*}
\left\vert \frac{l\varepsilon \tau }{\left( l-\varepsilon \mu \right) ^{2}}-%
\frac{\varepsilon \tau }{l}\right\vert =\left\vert \frac{\varepsilon \tau }{l%
}\right\vert \left\vert \left( \frac{1}{1-\frac{\varepsilon }{l}\mu }\right)
^{2}-1\right\vert \leq 1\cdot \left\vert \left( \frac{1}{1\pm \frac{%
\varepsilon }{l_{0}}h_{0}}\right) ^{2}-1\right\vert \leq C\frac{%
h_{0}\varepsilon}{l_{0}}
\end{equation*}%
for $\left\vert \varepsilon \tau \right\vert \leq l,$ $\left\vert \mu
\right\vert \leq h_{0}$ and $l\geq l_{0}>0$. \
For $\left\vert \tau \right\vert >l/\varepsilon $, the estimate to get $%
\left( \ref{quadratic-C}\right) $ is similar (actually simpler) since
\begin{equation}
d\mathcal{F}_{\left( u_{app}^{\varepsilon },\frac{l}{\varepsilon }\right)
}\left( \xi ,\mu \right) \left( \xi ^{\prime },\mu ^{\prime }\right) =d%
\mathcal{F}_{\left( u_{app}^{\varepsilon },\frac{l}{\varepsilon }-\mu
\right) }\left( \xi \right) \xi ^{\prime }+\mu ^{\prime }Pal_{\xi }^{-1}%
\frac{\partial u_{\xi }}{\partial \tau }. \label{outter-variation}
\end{equation}
Clearly
\begin{eqnarray*}
C\left\Vert \xi \right\Vert _{\varepsilon }\left\Vert \xi ^{\prime
}\right\Vert _{\varepsilon }+C\left\vert \mu \right\vert \left\Vert \xi
^{\prime }\right\Vert _{\varepsilon }+C\left\vert \mu ^{\prime }\right\vert
\left\Vert \xi \right\Vert _{\varepsilon } &\leq &C\left( \left\Vert \xi
\right\Vert _{\varepsilon }+\left\vert \mu \right\vert \right) \left(
\left\Vert \xi ^{\prime }\right\Vert _{\varepsilon }+\left\vert \mu ^{\prime
}\right\vert \right) \\
&=&C\left\Vert \left( \xi ,\mu \right) \right\Vert _{\varepsilon }\left\Vert
\left( \xi ^{\prime },\mu ^{\prime }\right) \right\Vert _{\varepsilon }.
\end{eqnarray*}%
so the proposition follows.
\end{proof}
\begin{rem}
When $l\rightarrow 0$ the Sobolev constant $C\left( l_{0},p\right) $ blows
up so we can not get uniform constant $C$ in the above quadratic estimate.
Different argument (blowing up the target) is needed for gluing. This is the
nodal Floer case and was treated in \cite{oh-zhu}. When $l_{0}\leq
l\rightarrow \infty $ the constant $C$ remains uniform.
\end{rem}
To perturb $u_{app}^{\varepsilon }$ to be a true solution of the Floer
equation $\overline{\partial }_{\left( K_{\varepsilon },J_{\varepsilon
}\right) }u=0$, we need the following abstract implicit function theorem in
\cite{MS}
\begin{prop}
\label{prop:abstractglue} Let $X,Y$ be Banach spaces and $U$ be an open set
in $X$. The map $F:X\rightarrow Y$ is continuous differentiable. For $%
x_{0}\in U,D:=dF(x_{0}):X\rightarrow Y$ is surjective and has a bounded
linear right inverse $Q:Y\rightarrow X$, with $\left\Vert Q\right\Vert \leq
C $. Suppose that there exists $h>0$ such that $x\in B_{h}(x_{0})\subset U$%
\begin{equation*}
x\in B_{h}(x_{0})\subset U\implies \left\Vert dF(x)-D\right\Vert \leq \frac{1%
}{2C}.
\end{equation*}%
Suppose
\begin{equation*}
\left\Vert F(x_{0})\right\Vert \leq \frac{h}{4C},
\end{equation*}%
then there exists a unique $x\in B_{h}(x_{0})$ such that%
\begin{equation*}
F(x)=0,~x-x_{0}\in \operatorname{Image}Q,~\left\Vert x-x_{0}\right\Vert \leq
2C\left\Vert F(x_{0})\right\Vert .
\end{equation*}
\end{prop}
For the remaining section, we will wrap up the gluing construction by
identifying the corresponding Banach spaces $X,\, Y$ and the nonlinear map $%
F $, the point $x_0$ and the right inverse $Q$ for the purpose of applying
this proposition.
For a fixed sufficiently small $\varepsilon _{0}>0$, let $\varepsilon \in
(0,\varepsilon _{0}]$. For any $u_{app}^{\varepsilon }$, we choose Banach
spaces
\begin{equation*}
X=\left( \Gamma (\left( u_{app}^{\varepsilon }\right) ^{\ast }TM)\times T_{l}%
\mathbb{R},\left\Vert \cdot \right\Vert _{\varepsilon }\right) ,\,Y=\left(
\Gamma \left( u_{app}^{\varepsilon }\right) ^{\ast }TM)\otimes \Lambda
^{0,1}\left( \mathbb{R\times }S^{1}\right) ,\left\Vert \cdot \right\Vert
_{\varepsilon }\right)
\end{equation*}%
with the Banach norms $\left\Vert \cdot \right\Vert _{\varepsilon }$ for $%
\xi $ and $\eta $ defined in section \ref{sec:off-shell}. We choose
\begin{equation*}
U_{\varepsilon }=\{\xi \in X\mid \Vert \xi \Vert _{\varepsilon
}\}<C(\varepsilon _{0}),\,\quad x_{0}=0\in U_{\varepsilon }
\end{equation*}%
for a constant $C(\varepsilon _{0})$ depending only on $\varepsilon _{0}$
such that $C(\varepsilon _{0})\rightarrow 0$ as $\varepsilon _{0}\rightarrow
0$. We emphasize that this constant $C(\varepsilon _{0})$ does not depend on
the choice of $0<\varepsilon \leq \varepsilon _{0}$.
Let $F$ be the map $\mathcal{F}_{u_{app}^{\varepsilon }}$ defined in the
beginning of the section. By Proposition \ref{prop:uniform-inverse-bound}
and Proposition \ref{prop:quadratic} the $C,h$ in our case are uniform while
$\left\Vert F\left( x_{0}\right) \right\Vert \leq C\varepsilon ^{\frac{1}{p}%
} $,\ so we can apply Proposition \ref{prop:abstractglue} to find a
perturbation $\left( \xi ,\mu \right) $ such that
\begin{equation}
\overline{\partial }_{(K_{\varepsilon },J_{\varepsilon })}(\exp
_{u_{app}^{\varepsilon }}\xi )\left( P_{\mu }\left( \tau \right) ,t\right) =0
\label{eq:F}
\end{equation}%
i.e., $\left( \exp _{u_{app}^{\varepsilon }}\xi \right) \left( P_{\mu
}\left( \tau \right) ,t\right) $ is a genuine solution for \eqref{eq:KJe}.
We denote
\begin{eqnarray*}
&{}&{\mathcal{M}}^{dfd}(K_{-},J_{-};f,K_{+},J_{+};z_{-},z_{+};A_{-}\#A_{+})
\\
&=&{\mathcal{M}}(K_{-},J_{-};z_{-};A_{-}){}_{ev_{+}}\times _{ev_{0}}{%
\mathcal{M}}(f;[0,\ell ]){}_{ev_{\ell }}\times _{ev_{-}}{\mathcal{M}}%
(K_{+},J_{+};z_{+};A_{+})
\end{eqnarray*}%
and
\begin{equation*}
{\mathcal{M}}_{\leq \varepsilon
_{0}}^{para}(K,J;z_{-},z_{+};A_{-}\#A_{+})=\bigcup _{\varepsilon \in
(0,\varepsilon _{0}]}{\mathcal{M}}^{\varepsilon }(K_{\varepsilon
},J_{\varepsilon };z_{-},z_{+};A_{-}\#A_{+}).
\end{equation*}%
We denote by the smooth map
\begin{eqnarray}
&{}&\RIfM@\expandafter\text@\else\expandafter\mbox\fi{Glue}:(0,\varepsilon _{0}]\times {\mathcal{M}}%
_{(0;1,1)}^{dfd}(K_{-},J_{-};f,K_{+},J_{+};z_{-},z_{+};A_{-}\#A_{+}) \notag
\label{eq:Glue} \\
&{}&\hskip1in\rightarrow \bigcup_{0<\varepsilon \leq \varepsilon _{0}}{%
\mathcal{M}}^{\varepsilon }(K_{\varepsilon },J_{\varepsilon
};z_{-},z_{+};A_{-}\#A_{+})
\end{eqnarray}%
the composition of $\RIfM@\expandafter\text@\else\expandafter\mbox\fi{preG}$ followed by the map
\begin{equation*}
u_{app}^{\varepsilon }\rightarrow \left( \exp _{u_{app}^{\varepsilon }}\xi
\right) \left( P_{\mu }\cdot ,\cdot \right)
\end{equation*}%
obtained by solving the equation \eqref{eq:F} applying Proposition \ref%
{prop:abstractglue}. We also denote by $\RIfM@\expandafter\text@\else\expandafter\mbox\fi{Glue}_{\varepsilon }$ the
slice of $\RIfM@\expandafter\text@\else\expandafter\mbox\fi{Glue}$ for $\varepsilon $.
The main gluing theorem is the following
\begin{thm}
\label{thm:maingluing} Let $(K_{\varepsilon},J_{\varepsilon})$ be the family
of Floer data defined in \eqref{eq:KR}. Then
\begin{enumerate}
\item there exists a topology on
\begin{equation*}
{\mathcal{M}}_{\leq \varepsilon
_{0}}^{para}(K,J;z_{-},z_{+};A_{-}\#A_{+})=\bigcup_{0<\varepsilon \leq
\varepsilon _{0}}{\mathcal{M}}^{\varepsilon }(K_{\varepsilon
},J_{\varepsilon };z_{-},z_{+};A_{-}\#A_{+})
\end{equation*}
with respect to which the gluing construction defines a proper embedding
\begin{eqnarray*}
Glue & : &(0,\varepsilon _{0}]\times {\mathcal{M}}%
_{(0;1,1)}^{dfd}(K_{-},J_{-};f,K_{+},J_{+};z_{-},z_{+};A_{-}\#A_{+}) \\
&{}& \qquad \rightarrow {\mathcal{M}}_{\leq \varepsilon
_{0}}^{para}(K,J;z_{-},z_{+};A_{-}\#A_{+})
\end{eqnarray*}
for sufficiently small $\varepsilon _{0}$.
\item the above mentioned topology can be embedded into
\begin{equation*}
{\mathcal{M}}^{dfd}(K_{-},J_{-};f,K_{+},J_{+};z_{-},z_{+};A_{-}\#A_{+})%
\bigcup {\mathcal{M}}_{\leq \varepsilon
_{0}}^{para}(K,J;z_{-},z_{+};A_{-}\#A_{+})
\end{equation*}
as a set,
\item the embedding $\RIfM@\expandafter\text@\else\expandafter\mbox\fi{Glue}$ smoothly extends to the embedding
\begin{eqnarray*}
\overline{\RIfM@\expandafter\text@\else\expandafter\mbox\fi{Glue}} &:&[0,\varepsilon _{0})\times {\mathcal{M}}%
^{dfd}(K_{-},J_{-};f,K_{+},J_{+};z_{-},z_{+};A_{-}\#A_{+}) \\
&\rightarrow &\overline{{\mathcal{M}}_{\leq \varepsilon _{0}}^{para}}%
(K,J;z_{-},z_{+};A_{-}\#A_{+})
\end{eqnarray*}%
that satisfies $\overline{\RIfM@\expandafter\text@\else\expandafter\mbox\fi{Glue}}(0,u_{-},\chi ,u_{+})=(u_{-},\chi
,u_{+})$.
\end{enumerate}
\end{thm}
One essential ingredient to establish to complete the proof of this theorem
is the following surjectivity whose proof we give in the next section.
\begin{prop}
\label{prop:surjective} There exists some $\varepsilon _{0}>0,$ $\zeta
_{0}>0 $ and a function $\delta :\left( 0,\varepsilon _{0}\right)
\rightarrow \mathbb{R}_{+}$ such that the gluing map
\begin{equation*}
\RIfM@\expandafter\text@\else\expandafter\mbox\fi{Glue}:(0,\varepsilon _{0}]\times {\mathcal{M}}%
^{dfd}(K_{-},J_{-};f,K_{+},J_{+};z_{-},z_{+};A_{-}\#A_{+})
\end{equation*}%
is surjective onto
\begin{equation*}
{\mathcal{M}}_{\leq \varepsilon
_{0}}^{para}(K,J;z_{-},z_{+};A_{-}\#A_{+})\cap \bigcup_{\substack{ %
0<\varepsilon \leq \varepsilon _{0} \\ 0<\zeta \leq \zeta _{0}}}V_{\zeta
,\delta \left( \varepsilon \right) }^{\varepsilon }.
\end{equation*}%
Here $V_{\zeta ,\delta }^{\varepsilon }$ is the open subset given in %
\eqref{eq:Ve}.
\end{prop}
Once this is established, the standard argument employed by Donaldson will
complete the proof of this main theorem.
\section{Surjectivity}\label{sec:surjectivity}
In this section, we give the proof of Proposition %
\ref{prop:surjective}.
To prove surjectivity, we need to show that there exists $\varepsilon _{0}>0$%
, $0<\zeta _{0}<1$ and a function $\delta :(0,\varepsilon _{0}]\rightarrow
\mathbb{R}_{+}$ such that for $0<\varepsilon \leq \varepsilon _{0}$ and $%
0<\zeta \leq \zeta _{0}$, any pair $(\varepsilon ,u)\in {\mathcal{M}}_{\leq
\varepsilon _{0}}^{para}(K,J;z_{-},z_{+};A_{-}\#A_{+})$ with
\begin{equation}
d_{adia}^{\varepsilon ,\zeta }(u,(u_{-},\chi ,u_{+}))<\delta \left(
\varepsilon \right) \label{adia-distance-delta}
\end{equation}%
lies in the image of the gluing map
\begin{equation*}
\RIfM@\expandafter\text@\else\expandafter\mbox\fi{Glue}_{\varepsilon }:{\mathcal{M}}%
^{dfd}(K_{-},J_{-};f,K_{+},J_{+};z_{-},z_{+};A_{-}\#A_{+})\rightarrow {%
\mathcal{M}}(K_{\varepsilon },J_{\varepsilon };z_{-},z_{+};A_{-}\#A_{+}).
\end{equation*}%
The above condition $\left( \ref{adia-distance-delta}\right) $ implies that
\begin{enumerate}
\item $E_{J,\Theta _{\varepsilon }}(u)<\delta \left( \varepsilon \right) ,$
\item $d_{\RIfM@\expandafter\text@\else\expandafter\mbox\fi{\textrm{H}}}\left( u([-R(\varepsilon ),R(\varepsilon
)]\times S^{1}),\chi ([-l,l])\right) <\delta \left( \varepsilon \right) ,$
\item $d_{C^{\infty }\left( \pm \lbrack \frac{1}{2\pi }\ln \zeta _{0},\infty
)\times S^{1}\right) }\left( u(\cdot \pm \tau \left( \varepsilon \right)
,\cdot ),u_{\pm }\right) <\delta \left( \varepsilon \right) ,$
\item $\operatorname{diam}\left( u\left( \pm \left[ R\left( \varepsilon \right)
,\tau \left( \varepsilon \right) +\frac{1}{2\pi }\ln \zeta _{0}\right]
\times S^{1}\right) \right) <\delta \left( \varepsilon \right) .$
\end{enumerate}
By definition of $V_{\zeta ,\delta }^{\varepsilon }$, any element $u\in
V_{\zeta ,\delta }^{\varepsilon }$ can be expressed as
\begin{equation}
u\left( \tau ,t\right) =\exp _{u_{app}^{\varepsilon }}(\xi )\left( \tau
,t\right) \label{eq:uexpxi}
\end{equation}%
with $u_{app}^{\varepsilon }=\RIfM@\expandafter\text@\else\expandafter\mbox\fi{preG}(\varepsilon ,u_{-},\chi ,u_{+})$
for some
\begin{equation*}
(u_{-},\chi ,u_{+})\in {\mathcal{M}}%
^{dfd}(K_{-},J_{-};f,K_{+},J_{+};z_{-},z_{+};A_{-}\#A_{+})
\end{equation*}%
and $\xi \in \Gamma \left( \left( u_{app}^{\varepsilon }\right) ^{\ast
}TM\right) $ for some $\xi $. More precisely we introduce the off-shell
version of the pre-gluing map
\begin{equation*}
\widetilde{\RIfM@\expandafter\text@\else\expandafter\mbox\fi{preG}}:((u_{-},\xi _{-}),(\chi ,\xi _{0}),(u_{+},\xi
_{+}),\mu )\mapsto \RIfM@\expandafter\text@\else\expandafter\mbox\fi{preG}\left( \exp _{u_{-}}(\xi _{-}),\exp _{\chi
}(\xi _{0}),\exp _{u_{+}}(\xi _{+})\right) \left( P_{\mu }\tau ,t\right)
\end{equation*}%
by the same formula for preG as \eqref{eq:uappe} with $u_{\pm }$ and $\chi $
replaced by $\exp _{u_{\pm }}(\xi _{\pm })$ and $\exp _{\chi }(\xi _{0})$
respectively, and $\mu \in T_{R\left( \varepsilon \right) }\mathbb{R}_{+}$
corresponds to the reparameterization of $u_{app}^{\varepsilon }$ for $%
\left( \tau ,t\right) \in \left[ -R\left( \varepsilon \right) ,R\left(
\varepsilon \right) \right] \times S^{1}$ by $\left( \tau ^{\prime
},t\right) \in \left[ -R\left( \varepsilon \right) +\mu ,R\left( \varepsilon
\right) -\mu \right] \times S^{1}$ as in $\left( \ref{reparameterize-tau}%
\right) $.
The following is an immediate translation of the stable map convergence
together with adiabatic convergence result in Theorem \ref{thm:centrallimit}.
\begin{lem}
There exists some $\varepsilon _{0}>0$ such that for any $0<\varepsilon \leq
\varepsilon _{0}$ any
\begin{equation*}
u\in \mathcal{M}(K_{\varepsilon },J_{\varepsilon
};z_{-},z_{+};A_{-}\#A_{+})\cap \bigcup_{0<\zeta \leq \zeta _{0}}V_{\zeta
,\delta (\varepsilon )}^{\varepsilon },
\end{equation*}%
can be expressed as
\begin{equation*}
u(\tau ,t)=\widetilde{\RIfM@\expandafter\text@\else\expandafter\mbox\fi{preG}}((u_{-},\xi _{-}),(\chi ,\xi
_{0}),(u_{+},\xi _{+}))(P_{\mu }\tau ,t)
\end{equation*}%
for some $(u_{-},\chi ,u_{+})$ such that
\begin{equation}
\Vert \xi _{\pm }\Vert _{L^{\infty }}\leq \delta ,\quad \Vert \xi _{0}\Vert
_{L^{\infty }}\leq \delta ,\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ \ }\left\vert \mu \right\vert \leq \delta .
\label{eq:Linftydelta}
\end{equation}
\end{lem}
By the above lemma, for $u\in \mathcal{M}(K_{\varepsilon },J_{\varepsilon
};z_{-},z_{+};A_{-}\#A_{+})\cap \bigcup_{0<\zeta \leq \zeta _{0}}V_{\zeta
,\delta (\varepsilon )}^{\varepsilon }$ with $0<\varepsilon \leq \varepsilon
_{0}$, we can represent $u$ by tangent vectors $\left( \xi _{-},\xi _{0},\xi
_{+},\mu \right) $ via%
\begin{equation*}
u(\tau ,t)=\widetilde{\RIfM@\expandafter\text@\else\expandafter\mbox\fi{preG}}((u_{-},\xi _{-}),(\chi ,\xi
_{0}),(u_{+},\xi _{+}))(P_{\mu }\tau ,t).
\end{equation*}%
For notation brevity, we let $\xi =\left( \xi _{-},\xi _{0},\xi _{+}\right) $%
.
>From now on, $u$ is represented by $\left( \xi ,\mu \right) $. To prove
Proposition \ref{prop:surjective}, it is enough to prove the following via
the local uniquenss property of the gluing construction.
\begin{prop}[Norm-convergence]
\label{prop:norm-converge} \ Let $\Vert \left( \xi ,\mu \right) \Vert
_{\varepsilon }$ be the Banach norm as defined in \eqref{eq:xienorm}. Then
there exists some $0<\zeta _{0}<1$ and a function $\delta :(0,\varepsilon
_{0}]\rightarrow {\mathbb{R}}_{+}$ such that
\begin{equation*}
\Vert \left( \xi ,\mu \right) \Vert _{\varepsilon }\rightarrow 0
\end{equation*}%
uniformly over $u\in \mathcal{M}(K_{\varepsilon },J_{\varepsilon
};z_{-},z_{+};A_{-}\#A_{+})\cap \bigcup_{0<\zeta \leq \zeta _{0}}V_{\zeta
,\delta (\varepsilon )}^{\varepsilon }$ as $\varepsilon \rightarrow 0$.
\end{prop}
The rest of this section is devoted to the proof of this norm convergence.
We can actually take the function $\delta \left( \varepsilon \right) $ to be
$\delta \left( \varepsilon \right) =\varepsilon $ in this section.
We prove this by contradiction. Suppose that the Proposition is not true,
then there exist $0<\varepsilon _{j}\leq \varepsilon _{0},$ $0<\zeta
_{j}\leq \zeta _{0}$ and solutions $u^{\varepsilon _{j}}\in \mathcal{M}%
(K_{\varepsilon _{j}},J_{\varepsilon _{j}};z_{-},z_{+};A_{-}^{j}\#A_{+}^{j})$
represented by $\left( \xi _{j},\mu _{j}\right) $ such that
\begin{equation*}
d_{adia}^{\varepsilon _{j},\zeta _{j}}\left( u^{\varepsilon _{j}},\left(
u_{-},\chi ,u_{+}\right) \right) <\delta \left( \varepsilon _{j}\right)
=\varepsilon _{j}\rightarrow 0,
\end{equation*}%
but $\Vert \left( \xi _{j},\mu _{j}\right) \Vert _{\varepsilon
_{j}}\nrightarrow 0$. By the assumption $u^{\varepsilon _{j}}$, $%
u^{\varepsilon _{j}}$ cannot develop bubbles, so we may assume $%
u^{\varepsilon _{j}}\in \mathcal{M}(K_{\varepsilon _{j}},J_{\varepsilon
_{j}};z_{-},z_{+};A_{-}\#A_{+})$ for a fixed homology class $A_{-}\#A_{+}$,
and\ has the uniform $C^{1}$ estimate
\begin{equation}
|du^{\varepsilon _{j}}(\tau ,t)|\leq C<\infty . \label{eq:du<C}
\end{equation}%
Therefore the sequence $u^{\varepsilon _{j}}$ is pre-compact on any given
compact interval $[a,b]\times S^{1}$.
\subsection{Exponential map and adiabatic renormalization}
In general, on the whole ${\mathbb{R}}\times S^{1}$, \ we define $\xi
_{j}^{app}\left( \tau ,t\right) \in T_{u_{app}^{\varepsilon _{j}}\left( \tau
,t\right) }M$ $\ $by
\begin{equation*}
u^{\varepsilon _{j}}\left( \tau ,t\right) =\exp _{u_{app}^{\varepsilon
_{j}}\left( \tau ,t\right) }\xi _{j}^{app}\left( \tau ,t\right) .
\end{equation*}%
The exponential map makes sense because of the adiabatic convergence $%
u^{\varepsilon _{j}}\rightarrow \left( u_{-},\chi ,u_{+}\right) $.
For the later purpose, it turns out to be important to use the exponential
map at the \emph{genuine solution}
\begin{equation*}
u_{glue}^{\varepsilon _{j}}:=\operatorname{Glue}(u_{-},\chi ,u_{+};\varepsilon _{j}).
\end{equation*}%
We also express the same sequence $u^{\varepsilon _{j}}$ as
\begin{equation}
u^{\varepsilon _{j}}\left( \tau ,t\right) =\exp _{u_{glue}^{\varepsilon
_{j}}\left( \tau ,t\right) }\xi _{j}\left( \tau ,t\right)
\label{eq:glue-xij}
\end{equation}%
for $\xi _{j}\left( \tau ,t\right) \in T_{u_{glue}^{\varepsilon _{j}}\left(
\tau ,t\right) }M$. By the triangle inequality, the error estimates %
\eqref{eq:error} and the construction of perturbation in implicit function
theorem, we have
\begin{equation}
\left\Vert \xi _{app}^{\varepsilon _{j}}-Pal_{app}^{glue}\left( \varepsilon
_{j}\right) \left( \xi _{glue}^{\varepsilon _{j}}\right) \right\Vert
_{\varepsilon _{j}}\leq \Vert E(u_{app}^{\varepsilon
_{j}},u_{glue}^{\varepsilon _{j}})\Vert _{\varepsilon _{j}}\leq CE(\ell
)\varepsilon _{j}^{\frac{1}{p}} \label{eq:uapp-uglue}
\end{equation}%
where
\begin{equation*}
E(u_{app}^{\varepsilon _{j}},u_{glue}^{\varepsilon _{j}})=\exp
_{u_{app}^{\varepsilon _{j}}}^{-1}(u_{glue}^{\varepsilon _{j}}),
\end{equation*}%
(see \eqref{eq:map-E}), and $Pal_{app}^{glue}\left( \varepsilon _{j}\right)
:T_{u_{glue}^{\varepsilon _{j}}\left( \tau ,t\right) }M\rightarrow
T_{u_{app}^{\varepsilon _{j}}\left( \tau ,t\right) }M$ is the parallel
transport from $u_{glue}^{\varepsilon _{j}}\left( \tau ,t\right) $ to $%
u_{app}^{\varepsilon _{j}}\left( \tau ,t\right) $ along the minimal geodesic
connnecting them. Therefore
\begin{eqnarray*}
\Vert \xi _{app}^{\varepsilon _{j}}\Vert _{\varepsilon _{j}} &\leq
&\left\Vert Pal_{app}^{glue}\left( \varepsilon _{j}\right) \left( \xi
_{glue}^{\varepsilon _{j}}\right) \right\Vert _{\varepsilon _{j}}+\Vert
E(u_{app}^{\varepsilon _{j}},u_{glue}^{\varepsilon _{j}})\Vert _{\varepsilon
_{j}} \\
&\leq &C\Vert \xi _{glue}^{\varepsilon _{j}}\Vert _{\varepsilon
_{j}}+CE(\ell )\varepsilon _{j}^{\frac{1}{p}},
\end{eqnarray*}%
and to prove $\Vert \xi _{app}^{\varepsilon _{j}}\Vert _{\varepsilon
_{j}}\rightarrow 0$, it will be enough to prove
\begin{prop}
Let $\xi _{glue}^{\varepsilon _{j}}$ be as in \eqref{eq:glue-xij}. Then
\begin{equation*}
\Vert \xi _{glue}^{\varepsilon _{j}}\Vert _{\varepsilon _{j}}\rightarrow 0.
\end{equation*}
\end{prop}
The remaining section will be occupied by the proof of this proposition.
We consider the exponential map
\begin{equation*}
\exp :{\mathcal{U}}\subset TM\rightarrow M;\quad \exp (x,\xi ):=\exp
_{x}(\xi )
\end{equation*}%
and denote
\begin{equation*}
D_{1}\exp (x,\xi ):T_{x}M\rightarrow T_{x}M
\end{equation*}%
the (covariant) partial derivative with respect to $x$ and
\begin{equation*}
d_{2}\exp (x,\xi ):T_{x}M\rightarrow T_{x}M
\end{equation*}%
the usual derivative $d_{2}\exp (x,\xi ):=T_{\xi }\exp
_{x}:T_{x}M\rightarrow T_{x}M$. We recall the basic property of the
exponential map
\begin{equation*}
D_{1}\exp (x,0)=d_{2}\exp (x,0)=id.
\end{equation*}%
Denote $\chi _{j}:=\operatorname{cm}(u_{j})$. Then $\chi _{j}\rightarrow \chi $ on $%
[-\ell ,\ell ]$ by Theorem \ref{thm:centrallimit}. It is also useful to
introduce the map
\begin{equation}
E:V_{\Delta }\rightarrow {\mathcal{U}};\quad E(x,y)=\exp _{x}^{-1}(y)
\label{eq:map-E}
\end{equation}%
where $V_{\Delta }$ is the neighborhood of the diagonal of $M\times M$. Then
the following is the standard estimates on this map
\begin{eqnarray}
|d_{2}\exp (x,v)-\Pi _{x}^{\exp (x,v)}| &\leq &C|v| \notag
\label{exp-derivatives} \\
|D_{1}\exp (x,v)-\Pi _{x}^{\exp (x,v)}| &\leq &C|v|
\end{eqnarray}%
for $v\in T_{x}M$ where $C$ is independent of $v$, as long as $|v|$ is
sufficiently small, say smaller than the injectivity radius of the metric on
$M$. (see \cite{katcher}.)
In the following calculations, to simplify the notation, we \emph{suppress
the subindex }$j$\emph{\ from various notations}, i.e. the $\varepsilon ,\xi
$ mean $\varepsilon _{j},\xi _{j}$ respectively. We compute
\begin{equation*}
\frac{\partial u^{\varepsilon }}{\partial \tau }=D_{1}\exp
(u_{glue}^{\varepsilon },\xi )\frac{\partial u_{glue}^{\varepsilon }}{%
\partial \tau }+d_{2}\exp (u_{glue}^{\varepsilon },\xi )\frac{\partial \xi }{%
\partial \tau }.
\end{equation*}%
Similarly we compute
\begin{equation*}
\frac{\partial u^{\varepsilon }}{\partial t}=D_{1}\exp
(u_{glue}^{\varepsilon },\xi )\frac{\partial u_{glue}^{\varepsilon }}{%
\partial t}+d_{2}\exp (u_{glue}^{\varepsilon },\xi )\frac{\partial \xi }{%
\partial t}.
\end{equation*}%
We introduce the following invertible linear operators $P(x,v):T_{x}M%
\rightarrow T_{x}M$ defined by
\begin{equation*}
P(x,v):=(d_{2}\exp (x,v))^{-1}\circ D_{2}\exp (x,v).
\end{equation*}%
Then we have the following inequality
\begin{lem}
\label{lem:P(x,v)}
\begin{equation*}
|P(x,v)-id |\leq C|v|
\end{equation*}%
for a universal constant $C>0$ depending only on the injectivity radius.
\end{lem}
\begin{proof}
This is an immediate consequence of \eqref{exp-derivatives}.
\end{proof}
Now we consider the equation
\begin{equation*}
0=\frac{\partial u^{\varepsilon }}{\partial \tau }+J^{\varepsilon }\left(
\frac{\partial u^{\varepsilon }}{\partial t}-X_{H^{\varepsilon
}}(u^{\varepsilon })\right) .
\end{equation*}%
We re-write
\begin{eqnarray}
0 &=&\left( D_{1}\exp (u_{glue}^{\varepsilon },\xi )\frac{\partial
u_{glue}^{\varepsilon }}{\partial t}+d_{2}\exp (u_{glue}^{\varepsilon },\xi )%
\frac{D\xi }{\partial t}\right) \notag \label{eq:CRue} \\
&{}&\quad +J^{\varepsilon }\left( D_{1}\exp (u_{glue}^{\varepsilon },\xi )%
\frac{\partial u_{glue}^{\varepsilon }}{\partial t}+d_{2}\exp
(u_{glue}^{\varepsilon },\xi )\frac{D\xi }{\partial t}-X_{H^{\varepsilon
}}(u_{glue}^{\varepsilon })\right) \notag \\
&=&d_{2}\exp (u_{glue}^{\varepsilon },\xi )\frac{D\xi }{\partial t}%
+J^{\varepsilon }\left( d_{2}\exp (u_{glue}^{\varepsilon },\xi )\frac{D\xi }{%
\partial t}-DX_{H^{\varepsilon }}(u_{glue}^{\varepsilon })(\xi ))\right)
\notag \\
&{}&+\left( D_{1}\exp (u_{glue}^{\varepsilon },\xi )\frac{\partial
u_{glue}^{\varepsilon }}{\partial t}+J^{\varepsilon }\left( D_{1}\exp
(u_{glue}^{\varepsilon },\xi )\frac{\partial u_{glue}^{\varepsilon }}{%
\partial t}-X_{H^{\varepsilon }}(u_{glue}^{\varepsilon })\right) \right)
\notag \\
&{}&+J^{\varepsilon }N(u_{glue}^{\varepsilon },\xi )
\end{eqnarray}%
where $N(u_{glue}^{\varepsilon },\xi )$ is the higher order term
\begin{eqnarray}
N(u_{glue}^{\varepsilon },\xi ) &=&X_{H^{\varepsilon }}(u^{\varepsilon
})-X_{H^{\varepsilon }}(u_{glue}^{\varepsilon })-DX_{H^{\varepsilon
}}(u_{glue}^{\varepsilon })(\xi ) \notag \label{eq:Nuepsilon} \\
&=&X_{H^{\varepsilon }}(\exp (u_{glue}^{\varepsilon },\xi
))-X_{H^{\varepsilon }}(u_{glue}^{\varepsilon })-DX_{H^{\varepsilon
}}(u_{glue}^{\varepsilon })(\xi )
\end{eqnarray}%
obtained from the (pointwise) Taylor expansion of $X_{H^{\varepsilon }}(\exp
(x,v))$.
Now we denote the pull-back
\begin{equation*}
\widehat{J}^{\varepsilon }:=(d_{2}\exp (u_{glue}^{\varepsilon },\xi
))^{-1}J^{\varepsilon }d_{2}\exp (u_{glue}^{\varepsilon },\xi )
\end{equation*}%
and
\begin{equation*}
\widetilde{J}^{\varepsilon }:=(P(u_{glue}^{\varepsilon },\xi
))^{-1}J^{\varepsilon }P(u_{glue}^{\varepsilon },\xi ).
\end{equation*}%
Then we obtain the pointwise inequalities
\begin{equation}
|\widehat{J}^{\varepsilon }-J^{\varepsilon }|\leq C|\xi |,\quad |\widetilde{J%
}^{\varepsilon }-J^{\varepsilon }|\leq C|\xi |. \label{eq:inequality-Je}
\end{equation}%
We also have
\begin{eqnarray}
|(D_{1}\exp (u_{glue}^{\varepsilon },\xi ))^{-1}(X_{K^{\varepsilon
}}(u_{glue}^{\varepsilon }))-(X_{K^{\varepsilon }}(u_{glue}^{\varepsilon
}))| &\leq &C_{\varepsilon }|\xi |, \notag \\
|(d_{2}\exp (u_{glue}^{\varepsilon },\xi ))^{-1}(DX_{K^{\varepsilon
}}(u_{glue}^{\varepsilon }))-(DX_{K^{\varepsilon }}(u_{glue}^{\varepsilon
}))| &\leq &C_{\varepsilon }|\xi |. \label{eq:inequality-XHe}
\end{eqnarray}%
With this notation, we can simplify and write \eqref{eq:CRue} into
\begin{eqnarray}
&{}&\frac{D\xi }{\partial \tau }+\widehat{J}^{\varepsilon }\left( \frac{D\xi
}{\partial t}-(d_{2}\exp (u_{glue}^{\varepsilon }))^{-1}(DX_{K^{\varepsilon
}}(u_{glue}^{\varepsilon })(\xi ))\right) \notag \\
&=&-P(u_{glue}^{\varepsilon },\xi )\left( \frac{\partial
u_{glue}^{\varepsilon }}{\partial \tau }+\widetilde{J}^{\varepsilon }\left(
\frac{\partial u_{glue}^{\varepsilon }}{\partial t}-(D_{1}\exp
(u_{glue}^{\varepsilon },\xi ))^{-1}(X_{K^{\varepsilon
}}(u_{glue}^{\varepsilon }))\right) \right) \notag \\
&{}&-d_{2}\exp (u_{glue}^{\varepsilon },\xi )^{-1}(J^{\varepsilon }N(u_{glue}^{\varepsilon },\xi )) \label{eq:xi-on-chi}
\end{eqnarray}%
Combining the pointwise inequalities \eqref{eq:inequality-Je}, %
\eqref{eq:inequality-XHe} and the error estimate for $u_{glue}^{\varepsilon
} $, we obtain the differential inequality
\begin{equation*}
\left\vert \frac{D\xi }{\partial t}+\widehat{J}^{\varepsilon }\left( \frac{%
D\xi }{\partial t}-(DX_{K^{\varepsilon }}(u_{glue}^{\varepsilon })(\xi
))\right) \right\vert \leq C(\varepsilon )+C(|\xi |)|\xi |
\end{equation*}%
where $C(\varepsilon )$ is the error term for $u_{glue}^{\varepsilon }$.
For simplicity of exposition, we rewrite this equation (\emph{now with
subindex }$j$) into
\begin{eqnarray}
&{}&\frac{D\xi _{j}}{\partial \tau }+\widehat{J}^{\varepsilon _{j}}\left(
\tau ,t\right) \frac{\partial \xi _{j}}{\partial t}+A_{j}(u_{glue}^{%
\varepsilon _{j}}\left( \tau ,t\right) )\cdot \xi _{j}(\tau ,t) \notag \\
&=&B_{j}\left( u_{glue}^{\varepsilon _{j}}\left( \tau ,t\right) ,\xi
_{j}(\tau ,t)\right) +E_{\varepsilon _{j}}\left( \tau ,t\right)
\label{eq:simplifiedqxi}
\end{eqnarray}%
where $A_{j},\,B_{j}$ and $E_{\varepsilon _{j}}$ are defined by
\begin{eqnarray*}
A_{j}(x)\xi _{j} &=&-\widehat{J}^{\varepsilon _{j}}DX_{H^{\varepsilon
_{j}}}(x)(\xi _{j}),\\
\quad B_{j}(x,\xi _{j})& = &
-d_{2}\exp (u_{glue}^{\varepsilon },\xi )^{-1}(J^{\varepsilon
_{j}}N^{\varepsilon _{j}}(x,\xi _{j})) \\
E_{\varepsilon _{j}}\left( {\tau },{t}\right) &=&-P(u_{glue}^{\varepsilon
_{j}},\xi _{j})\left( \frac{\partial u_{glue}^{\varepsilon _{j}}}{\partial
\tau }+\widetilde{J}^{\varepsilon _{j}}\left( \frac{\partial
u_{glue}^{\varepsilon _{j}}}{\partial t}-(D_{1}\exp (u_{glue}^{\varepsilon
_{j}},\xi ))^{-1}(X_{K^{\varepsilon _{j}}}(u_{glue}^{\varepsilon
_{j}}))\right) \right) .
\end{eqnarray*}%
We have
\begin{equation}
|B_{j}(x,\xi _{j})|\leq C \, |\xi _{j}|^{2}
\label{eq:Bjquadratic}
\end{equation}%
for some uniform constant $C$, provided $\|\xi _{j}\|_{C^0} \leq \RIfM@\expandafter\text@\else\expandafter\mbox\fi{injectivity radius of the metric $g$}$.
(Note that this latter requirement is automatic since $\|\xi_j\|_{\varepsilon_j} \to 0$.)
The pointwise inequality \eqref{eq:Bjquadratic}
follows from the inequality \eqref{exp-derivatives} and the pointwise inequality
$$
|X_{H^{\varepsilon }}(\exp_x(v))-X_{H^{\varepsilon }}(x)-DX_{H^{\varepsilon
}}(x)(v)|\leq C |v|^2
$$
where $C$ depends only on $H$, the injectivity radius of the metric as long as
$|v| \leq \RIfM@\expandafter\text@\else\expandafter\mbox\fi{injectivity radius of the metric $g$}$.
In the transition region $\Omega _{\pm }\left( \varepsilon _{j}\right) :=\pm %
\left[ R(\varepsilon _{j}),\tau \left( \varepsilon _{j}\right) \right]
\times S^{1}$, we do not renormalize but consider \eqref{eq:simplifiedqxi}
itself. From the adiabatic convergence of $u_{j}\rightarrow \left(
u_{-},\chi ,u_{+}\right) $ we have
\begin{equation}
\left\vert \xi _{j}\left( \pm R\left( \varepsilon _{j}\right) ,t\right)
\right\vert ,\,|\xi _{j}(\pm \tau \left( \varepsilon _{j}\right) ,t)|\leq
\delta _{j} \label{eq:xibdyatRejSej}
\end{equation}%
where
\begin{equation*}
\delta _{j}=\varepsilon _{j}\rightarrow 0\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ \ \ as }j\rightarrow \infty .
\end{equation*}
We next consider the region $\Theta \left( \varepsilon _{j}\right)
=[-R(\varepsilon _{j}),R(\varepsilon _{j})]\times S^{1}$ for $\left( \tau
,t\right) $. We recall that on $\Theta \left( \varepsilon _{j}\right) $ we
have $u_{glue}^{\varepsilon _{j}}(\tau ,t)=\chi (\varepsilon _{j}\tau )$ and
so
\begin{equation*}
\widehat{J}^{\varepsilon _{j}}(u_{glue}^{\varepsilon _{j}})\equiv
J_{0}\equiv \widetilde{J}^{\varepsilon _{j}}(u_{glue}^{\varepsilon _{j}}).
\end{equation*}%
By renomalizing the domain%
\begin{equation*}
(\overline{\tau },\overline{t})=(\varepsilon _{j}\tau ,\varepsilon _{j}t),\,%
\RIfM@\expandafter\text@\else\expandafter\mbox\fi{\ }\overline{u_{j}}\left( \overline{\tau },\overline{t}\right)
=u_{j}\left( \frac{\overline{\tau }}{\varepsilon _{j}},\frac{\overline{t}}{%
\varepsilon _{j}}\right) =u_{j}\left( \tau ,t\right) ,
\end{equation*}%
and applying the Hausdorff convergence of $u_{j}|_{[-R(\varepsilon
_{j}),R(\varepsilon _{j})]\times S^{1}}$ to $\chi |_{[-\ell ,\ell ]}$, we
can write $\overline{u}_{j}$
\begin{equation}
\overline{u}_{j}(\overline{\tau },\overline{t})=\exp _{\chi (\overline{\tau }%
)}\overline{\xi }_{j}(\overline{\tau },\overline{t}) \label{eq:baruexpxi}
\end{equation}%
for $(\overline{\tau },\overline{t})\in \lbrack -\ell ,\ell ]\times {\mathbb{%
R}}/2\pi \varepsilon _{j}{\mathbb{Z}}$, a cylinder with radius $\varepsilon
_{j}$. Here we have
\begin{equation*}
\overline{\xi }_{j}(\overline{\tau },\overline{t})\in T_{\chi (\overline{%
\tau })}M.
\end{equation*}%
If we restrict $\left( \tau ,t\right) $ on $[-R(\varepsilon
_{j}),R(\varepsilon _{j})]\times S^{1}$, then $\xi _{j}\left( \tau ,t\right)
=$ $\overline{\xi }_{j}(\overline{\tau },\overline{t})$.
\bigskip Furthermore it easily follows from $\left( \ref{exp-derivatives}%
\right) $ and smoothness of the exponential map that%
\begin{eqnarray*}
&{}&|d_{2}\exp (\chi (\overline{\tau }),\overline{\xi }_{j}(\overline{\tau },%
\overline{t})))^{-1}(D_{1}\exp (\chi (\overline{\tau }),\overline{\xi }_{j}(%
\overline{\tau },\overline{t}))(\dot{\chi}(\overline{\tau }))-\dot{\chi}(%
\overline{\tau })| \\
&\leq &C|\overline{\xi }(\tau ,t)||\dot{\chi}(\overline{\tau })|
\end{eqnarray*}%
and
\begin{equation*}
|\exp _{\chi (\overline{\tau })}^{\ast }(\operatorname{grad}f)(\overline{\xi }_{j}(%
\overline{\tau },\overline{t}))-\operatorname{grad}f(\chi (\overline{\tau }))|\leq C|%
\overline{\xi }_{j}(\overline{\tau },\overline{t})|
\end{equation*}%
and
\begin{equation*}
|(\exp _{\chi (\overline{\tau })})^{\ast }J^{\varepsilon _{j}}(\chi (%
\overline{\tau }))-J^{\varepsilon _{j}}|\leq C|\overline{\xi }_{j}(\overline{%
\tau },\overline{t})|
\end{equation*}%
where the constant $C$ depends only on $M$.
If we take the reparamterization
\begin{equation*}
(\overline{\tau },\overline{t})=(\varepsilon _{j}\tau ,\varepsilon
_{j}t),\,\xi _{j}(\tau ,t)=\overline{\xi }_{j}(\varepsilon _{j}\tau
,\varepsilon _{j}t),
\end{equation*}%
then \eqref{eq:simplifiedqxi} becomes
\begin{equation}
\frac{D\xi _{j}}{\partial \tau }+J_{0}(\chi (\varepsilon _{j}\tau ))\frac{%
\partial \xi _{j}}{\partial t}+\varepsilon _{j}A\left( \chi (\varepsilon
_{j}\tau )\right) \cdot \xi _{j}=\varepsilon _{j}\left( B\left( \chi
(\varepsilon _{j}\tau ),\xi _{j}\right) \right) \label{eq:renormalizedeqxi}
\end{equation}%
where
\begin{eqnarray*}
A\left( \chi \right) \xi &=&({\Greekmath 0272} \operatorname{grad}f)(\chi )(\xi ) \\
B(\chi ,\overline{\xi }_{j}) &=&-d_2\exp (\chi,\xi_j)^{-1}
(J_{0}N(\chi ,\xi _{j})) \\
E_{\varepsilon _{j}}\left( \overline{\tau },\overline{t}\right) &=&-({%
\overline{\partial }}_{J,\varepsilon _{j}f}u_{glue}^{\varepsilon _{j}})(\tau
,t)=0.
\end{eqnarray*}
\subsection{Three-interval method of exponential estimates}
We first study the equation of $\xi _{j}$ on the transition regions
\begin{equation*}
\Omega _{\pm }\left( \varepsilon _{j}\right) :=\pm \left[ R(\varepsilon
_{j}),\tau \left( \varepsilon _{j}\right) \right] \times S^{1}.
\end{equation*}
Let $\delta _{0}>0$ be a number much smaller than the injectivity radius of $%
M$, and%
\begin{equation*}
h\left( \zeta \right) :=\frac{1}{2\pi }\left\vert \ln \zeta \right\vert >0.
\end{equation*}%
By the adiabatic convergence condition (3), (4), there exists small $0<\zeta
_{0}<1$ and integer $j_{0}$, such that for all $j\geq $ $j_{0}$ and $\delta
_{j}=\delta \left( \varepsilon _{j}\right) $,
\begin{eqnarray}
\RIfM@\expandafter\text@\else\expandafter\mbox\fi{dist }\left( u_{j}\left( \pm R\left( \varepsilon _{j}\right) \right)
,p_{\pm }\right) &<&\delta _{j}, \notag \\
d_{C^{\infty }\left( \pm \left[ \tau \left( \varepsilon _{j}\right) -h\left(
\zeta _{0}\right) ,\tau \left( \varepsilon _{j}\right) -h\left( \zeta
_{0}\right) +1\right] \times S^{1}\right) }\left( u_{j},u_{\pm
}^{\varepsilon _{j}}\right) &<&\delta _{j},\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ } \notag \\
\RIfM@\expandafter\text@\else\expandafter\mbox\fi{diam}\left( u_{j}\left( \pm \left[ R(\varepsilon _{j}),\tau \left(
\varepsilon _{j}\right) -h\left( \zeta _{0}\right) +1\right] \times
S^{1}\right) \right) &<&\delta _{j}. \label{3-distances-small}
\end{eqnarray}%
We denote the region
\begin{equation*}
\Omega _{\pm \zeta _{0}}\left( \varepsilon _{j}\right) :=\pm \left[
R(\varepsilon _{j}),\tau \left( \varepsilon _{j}\right) -h\left( \zeta
_{0}\right) +1\right] \times S^{1}\subset \Omega _{\pm }\left( \varepsilon
_{j}\right) .
\end{equation*}%
Without loss of generality we assume $\delta _{j}\leq \delta _{0}$. Then for
$\varepsilon _{j}$ small, from $\left( \ref{3-distances-small}\right) $ we
have%
\begin{equation}
u_{j}\left( \Omega _{\pm \zeta _{0}}\left( \varepsilon _{j}\right) \right)
\subset B_{p_{\pm }}\left( 2\delta _{0}\right) . \label{transition-small}
\end{equation}%
If we identify the neighborhood $B_{p_{\pm }}\left( 2\delta _{0}\right) $
into $T_{p_{\pm }}M$ by exponential map and deform the metric and almost
complex structure to standard $\left( g_{p_{\pm }},J_{p_{\pm }}\right) $,
then the $\xi _{j}$ can be simplified to%
\begin{equation*}
\xi _{j}=u_{j}-u_{glue}^{\varepsilon _{j}},
\end{equation*}%
where the \textquotedblleft $-$\textquotedblright\ is with respect to the
linear space structure on $T_{p_{\pm }}M$. Such simplification will not
affect the validity of the proof, as explained in section 8 and remark 11.1,
for it will only affect the $C^{1}$ pointwise estimate of a term of order $%
C\delta _{0}$.
We decompose the equation \eqref{eq:simplifiedqxi} into those of $0$-mode
and higer modes%
\begin{eqnarray}
\frac{\partial }{\partial \tau }\left( \xi _{j}\right) _{0} &=&\left(
B\right) _{0}\left( u_{glue}^{\varepsilon _{j}}\left( \tau ,t\right) ,\xi
_{j}\right) +\left( E_{\varepsilon _{j}}\right) _{0}, \notag \\
\overline{\partial }_{J}\widetilde{\xi _{j}} &=&\widetilde{B}\left(
u_{glue}^{\varepsilon _{j}}\left( \tau ,t\right) ,\xi _{j}\right) +%
\widetilde{E_{\varepsilon _{j}}}. \label{eq:decouple-Floer}
\end{eqnarray}%
\begin{rem}
In Theorem 1.2 of \cite{MT}, the higher mode exponential decay estimate
\begin{equation}
\left\vert \widetilde{\xi _{j}}\right\vert \leq Ce^{-\sigma \left(
l/\varepsilon _{j}-\left\vert \tau \right\vert \right) }\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ for }%
\left\vert \tau \right\vert \leq l/\varepsilon _{j} \label{high-mode-decay}
\end{equation}%
has been obtained (Their notation for higher mode is $\phi _{0}\left(
t,\theta \right) $ instead of our $\widetilde{\xi _{j}}\left( \tau ,t\right)
$). Their observation was that $\left( \ref{high-mode-decay}\right) $ can be
reduced to a local $L^{2}$ estimate
\begin{equation}
\left\Vert \widetilde{\xi _{j}}\right\Vert _{L^{2}\left( Z_{II}\right) }\leq
\frac{1}{2}\left( \left\Vert \widetilde{\xi _{j}}\right\Vert _{L^{2}\left(
Z_{I}\right) }+\left\Vert \widetilde{\xi _{j}}\right\Vert _{L^{2}\left(
Z_{III}\right) }\right) \label{local-elliptic-estimate}
\end{equation}%
on 3 sequential cylinders $Z_{I},Z_{II},Z_{III}\subset \left[ -l/\varepsilon
_{j},l/\varepsilon _{j}\right] \times $ $S^{1}$ of unit length, namely the
cylinders
\begin{equation*}
\left[ i-1,i\right] \times S^{1},\left[ i,i+1\right] \times S^{1},\left[
i+1,i+2\right] \times S^{1}
\end{equation*}%
for some integer $i$.
\end{rem}
To get the best $\sigma $ in the exponential decay (we need $\sigma $ to be
very close to $2\pi $), we recall in \cite{MT} they defined the constant%
\begin{equation*}
\gamma \left( c\right) =\frac{1}{e^{c}+e^{-c}}.
\end{equation*}%
The importance of $\gamma \left( c\right) $ is due to the identity
\begin{equation}
\int_{0}^{1}e^{c\tau }d\tau =\gamma \left( c\right) \left[
\int_{-1}^{0}e^{c\tau }d\tau +\int_{1}^{2}e^{c\tau }d\tau \right]
\label{identity}
\end{equation}%
which will appear in the $L^{2}$-energy of $\xi _{j}$ on $3$ sequential unit
length cylinders later. Notice that when $c>0$, $\gamma \left( c\right) $ is
a strictly decreasing function of $c$.
We recall the elementary
\begin{lem}[\protect\cite{MT} Lemma 9.4]
For nonnegative numbers $x_{k}$ ($k=0,1,\cdots N$), if $\ $for $1\leq k\leq
N-1$,
\begin{equation*}
x_{k}\leq \gamma \left( x_{k-1}+x_{k+1}\right)
\end{equation*}%
for some fixed constant $\gamma \in (0,1/2)$, then for $1\leq k\leq N-1$,
\begin{equation*}
x_{k}\leq x_{0}\xi ^{-k}+x_{N}\xi ^{-\left( N-k\right) },
\end{equation*}
where $\xi =\frac{1+\sqrt{1-4\gamma ^{2}}}{2\gamma }$.
\end{lem}
\begin{rem}
If $\gamma =\gamma \left( c\right) =\left( e^{c}+e^{-c}\right) ^{-1}$, then
we can check $\xi =e^{c}$ and the above inequality becomes the exponential
decay estimate%
\begin{equation}
x_{k}\leq x_{0}e^{-ck}+x_{N}e^{-c\left( N-k\right) }\RIfM@\expandafter\text@\else\expandafter\mbox\fi{.}
\label{L2-exp-decay}
\end{equation}%
for $1\leq k\leq N-1$.
\end{rem}
\begin{prop}
For any $0<\upsilon <1$, there exists $N_{0}=N_{0}\left( \upsilon \right) $
depending on $\upsilon $, such that for all $j>N_{0}$, on any $3$ sequential
cylinders $Z_{I},Z_{II}$ and $Z_{III}$ in $\left[ -\tau \left( \varepsilon
\right) ,\tau (\varepsilon )\right] \times S^{1}$ we have%
\begin{equation}
\left\Vert d\xi _{j}\right\Vert _{L^{2}\left( Z_{II}\right) }\leq \gamma
\left( 4\pi \upsilon \right) \cdot \left( \left\Vert d\xi _{j}\right\Vert
_{L^{2}\left( Z_{I}\right) }+\left\Vert d\xi _{j}\right\Vert _{L^{2}\left(
Z_{III}\right) }\right) \label{local-elliptic-estimate-g}
\end{equation}%
and on any $3$ sequential cylinders in $\left[ -R\left( \varepsilon \right)
,R(\varepsilon )\right] \times S^{1}$ we have
\begin{equation}
\left\Vert \widetilde{\xi _{j}}\right\Vert _{L^{2}\left( Z_{II}\right) }\leq
\gamma \left( 4\pi \upsilon \right) \cdot \left( \left\Vert \widetilde{\xi
_{j}}\right\Vert _{L^{2}\left( Z_{I}\right) }+\left\Vert \widetilde{\xi _{j}}%
\right\Vert _{L^{2}\left( Z_{III}\right) }\right) .
\label{local-elliptic-estimate-h}
\end{equation}
\end{prop}
\begin{proof}
We prove $\left( \ref{local-elliptic-estimate-g}\right) $ by contradiction.
Suppose that for some sequence $\xi _{j}\rightarrow 0$ $\left( \ref%
{local-elliptic-estimate-g}\right) $ is violated on $3$ sequential cylinders
$Z_{I}^{j},Z_{II}^{j}$ and $Z_{III}^{j}$ in $\left[ -\tau \left( \varepsilon
\right) ,\tau (\varepsilon )\right] \times S^{1}$:
\begin{equation*}
\left\Vert d\xi _{j}\right\Vert _{L^{2}\left( Z_{II}^{j}\right) }>\gamma
\left( 4\pi \upsilon \right) \left( \left\Vert d\xi _{j}\right\Vert
_{L^{2}\left( Z_{I}^{j}\right) }+\left\Vert d\xi _{j}\right\Vert
_{L^{2}\left( Z_{III}^{j}\right) }\right) .
\end{equation*}%
On $Z_{I}^{j}\cup Z_{II}^{j}\cup Z_{III}^{j}$ consider the rescaled sequence%
\begin{equation*}
\widehat{\xi _{j}}=\xi _{j}/\left\Vert \xi _{j}\right\Vert _{L^{\infty
}\left( Z_{I}^{j}\cup Z_{II}^{j}\cup Z_{III}^{j}\right) }\RIfM@\expandafter\text@\else\expandafter\mbox\fi{,}
\end{equation*}%
where the denorminator $\left\Vert \xi _{j}\right\Vert _{L^{\infty }\left(
Z_{I}^{j}\cup Z_{II}^{j}\cup Z_{III}^{j}\right) }$ is never $0$, otherwise $%
\xi _{j}\equiv 0$ on $Z_{I}^{j}\cup Z_{II}^{j}\cup Z_{III}^{j}$,
contradicting our assumption. Then%
\begin{eqnarray*}
\left\vert \widehat{\xi _{j}}\right\vert _{L^{\infty }\left( Z_{I}^{j}\cup
Z_{II}^{j}\cup Z_{III}^{j}\right) } &=&1, \\
\left\Vert d\widehat{\xi _{j}}\right\Vert _{L^{2}\left( Z_{II}^{j}\right) }
&>&\gamma \left( 4\pi \upsilon \right) \cdot \left( \left\Vert d\widehat{\xi
_{j}}\right\Vert _{L^{2}\left( Z_{I}^{j}\right) }+\left\Vert d\widehat{\xi
_{j}}\right\Vert _{L^{2}\left( Z_{III}^{j}\right) }\right) , \\
\left( \frac{\partial }{\partial \tau }+J_{0}\frac{\partial }{\partial t}%
\right) \widehat{\xi _{j}} &=&\left( B_{j}\left( \xi _{j}\right)
+E_{\varepsilon _{j}}\right) /\left\Vert \xi _{j}\right\Vert _{L^{\infty
}\left( Z_{I}^{j}\cup Z_{II}^{j}\cup Z_{III}^{j}\right) }.
\end{eqnarray*}
We need the following lemma
\begin{lem}
For $\left( \tau ,t\right) \in \lbrack -\tau (\varepsilon _{j}),\tau
(\varepsilon _{j})]\times S^{1}$, we have
\begin{equation}
|E_{\varepsilon _{j}}\left( {\tau },{t}\right) |\leq C|\xi _{j}(\tau
,t)|\left( \left\vert \frac{\partial u_{glue}^{\varepsilon _{j}}}{\partial t}%
(\tau ,t)\right\vert +|X_{K_{\varepsilon _{j}}}|\right) . \label{E-error}
\end{equation}%
Especially for $Z_{I}^{j}\cup Z_{II}^{j}\cup Z_{III}^{j}\subset \left[ -\tau
\left( \varepsilon _{j}\right) ,\tau \left( \varepsilon _{j}\right) \right]
\times S^{1}$, we have
\begin{equation*}
\lim_{j\rightarrow \infty }E_{\varepsilon _{j}}\left( \tau ,t\right)
/\left\Vert \xi _{j}\right\Vert _{L^{\infty }\left( Z_{I}^{j}\cup
Z_{II}^{j}\cup Z_{III}^{j}\right) }=0.
\end{equation*}
\end{lem}
\begin{proof}
We recall that $u_{glue}^{\varepsilon _{j}}$ is a genuine solution for
\begin{equation}
{\frac{\partial u}{\partial \tau }}+J^{\varepsilon _{j}}\left( {\frac{%
\partial u}{\partial t}}-X_{K^{\varepsilon _{j}}}(u)\right) =0.
\label{eq:perturbedCRKe}
\end{equation}%
Therefore using Lemma \ref{lem:P(x,v)}, we obtain
\begin{equation*}
|E_{\varepsilon _{j}}\left( {\tau },{t}\right) |\leq C|\widetilde{J}%
^{\varepsilon _{j}}-J^{\varepsilon _{j}}|\left\vert \frac{\partial
u_{glue}^{\varepsilon _{j}}}{\partial t}\right\vert +C_{1}|(D_{1}\exp
_{u_{glue}^{\varepsilon _{j}}})^{-1}(\xi _{\varepsilon
_{j}})-id||X_{K_{\varepsilon _{j}}}(u)|.
\end{equation*}%
for some uniform constants $C$ and $C_{1}$. By \eqref{eq:inequality-Je}, we
obtain $\left( \ref{E-error}\right) $. By the definition \eqref{eq:Kepsilon}
of $K_{\varepsilon_{j}}$,
\begin{equation*}
K_{\varepsilon _{j}}\left( \tau ,t,x\right) =\kappa _{0}^{\varepsilon
_{j}}\left( \tau \right) \cdot \varepsilon _{j}f\left( x\right) \RIfM@\expandafter\text@\else\expandafter\mbox\fi{ for }%
\left\vert \tau \right\vert \leq \tau \left( \varepsilon _{j}\right) ,
\end{equation*}%
so $|X_{K_{\varepsilon _{j}}}|\rightarrow 0$ as $\varepsilon _{j}\rightarrow
0$. We also have the $\xi _{j}:=E\left( u_{app}^{\varepsilon
_{j}},u_{glue}^{\varepsilon _{j}}\right) $ satisfying a Cauchy-Riemann
equation with an inhomogeneous term of order $\varepsilon _{j}$, and $%
\left\vert \xi _{j}\right\vert _{C^{o}}\leq C\varepsilon _{j}^{1/p}$ by the
error estimate$\left( \ref{eq:error}\right) $ and Sobolev embedding $\left( %
\ref{e-Sobolev}\right) $. Therefore by the interior Schauder estimate of $%
\xi _{j}$ on $\left[ \tau -\frac{1}{2},\tau +\frac{1}{2}\right] \times
S^{1}\subset \left[ \tau -1,\tau +1\right] \times S^{1}$ we have $\left\vert
{\frac{\partial }{\partial t}}\xi _{j}\left( \tau ,t\right) \right\vert \leq
C\varepsilon _{j}^{1/p}$. Thus for $\left\vert \tau \right\vert \leq \tau
\left( \varepsilon _{j}\right) $, by the triangle inequality we have
\begin{eqnarray*}
\left\vert \frac{\partial u_{glue}^{\varepsilon _{j}}}{\partial t}\left(
\tau ,t\right) \right\vert &\leq &C\left( \left\vert \frac{\partial
u_{app}^{\varepsilon _{j}}}{\partial t}\left( \tau ,t\right) \right\vert
+\left\vert {\frac{\partial }{\partial t}}\xi _{j}\left( \tau ,t\right)
\right\vert \right) \\
&\leq &C\left( \varepsilon _{j}+\varepsilon _{j}^{1/p}\right) \rightarrow 0%
\RIfM@\expandafter\text@\else\expandafter\mbox\fi{,}
\end{eqnarray*}
where in the last inequality we have used the exponential decay of $%
\left\vert du_{\pm }^{\varepsilon _{j}}\right\vert $ for $\left\vert \tau
\right\vert \leq \tau \left( \varepsilon _{j}\right) $. Therefore
\begin{equation*}
E_{\varepsilon _{j}}\left( \tau ,t\right) /\left\Vert \xi _{j}\right\Vert
_{L^{\infty }\left( Z_{I}^{j}\cup Z_{II}^{j}\cup Z_{III}^{j}\right) }\leq
C\left( \left\vert \frac{\partial u_{glue}^{\varepsilon _{j}}}{\partial t}%
(\tau ,t)\right\vert +|X_{K_{\varepsilon _{j}}}|\right) \rightarrow 0
\end{equation*}%
as $j\rightarrow \infty $. The lemma follows.
\end{proof}
\begin{rem}
Here is the place where we need to use the exponential map around the
genuine solution $u_{glue}^{\varepsilon }$ instead of $u_{app}^{\varepsilon
} $. If we had used the latter, we would not have the estimate given in this
lemma. This is because $u_{app}^{\varepsilon }$ is only an approximate
solution of \eqref{eq:perturbedCRKe} whose error term
\begin{equation*}
\overline{\partial }_{J,K^{\varepsilon _{j}}}(u_{app}^{\varepsilon _{j}})=%
\frac{\partial u}{\partial \tau }+J^{\varepsilon _{j}}\left( \frac{\partial u%
}{\partial t}-X_{K^{\varepsilon _{j}}}(u_{app}^{\varepsilon _{j}})\right) ,
\end{equation*}%
which is not vanishing in general, is hard to compare with $\xi
_{app}^{\varepsilon _{j}}$ when we express $u^{\varepsilon _{j}}=\exp
_{u_{app}^{\varepsilon _{j}}}\xi _{app}^{\varepsilon _{j}}$. Then it seems
much harder to get
\begin{equation*}
E_{\varepsilon _{j}}\left( \tau ,t\right) /\left\Vert \xi
_{app}^{\varepsilon _{j}}\right\Vert _{L^{\infty }\left( Z_{I}^{j}\cup
Z_{II}^{j}\cup Z_{III}^{j}\right) }\rightarrow 0,
\end{equation*}
a key to apply the three-interval method in the following to derive the
desired exponential decay of $\xi _{app}^{\varepsilon _{j}}$. This is the
reason why we first replaced $u_{app}^{\varepsilon _{j}}$ by the genuine
solution $u_{glue}^{\varepsilon _{j}}$ in the exponential map $\left( \ref%
{eq:glue-xij}\right) $ in the beginning of our derivation.
\end{rem}
Using the lemma and \eqref{eq:Bjquadratic}, we obtain
\begin{equation*}
\frac{|B\left( u_{glue}^{\varepsilon _{j}}\left( \tau ,t\right) ,\xi
_{j}(\tau ,t)\right) +E_{\varepsilon _{j}}\left( \tau ,t\right) |}{\Vert \xi
_{j}\Vert _{L^{\infty }(Z_{I}^{j}\cup Z_{II}^{j}\cup Z_{III}^{j})}}\leq
C\left( |\xi _{j}|+\left\vert \frac{\partial u_{glue}^{\varepsilon _{j}}}{%
\partial t}\right\vert \right) \rightarrow 0
\end{equation*}%
uniformly over $Z_{I}^{j}\cup Z_{II}^{j}\cup Z_{III}^{j}\subset \lbrack
-\left( \tau \left( \varepsilon _{j}\right) -h\left( \zeta _{j}\right)
\right) ,\tau (\varepsilon _{j})-h\left( \zeta _{j}\right) ]\times S^{1}$.
After possibly shifting $Z_{I}^{j}\cup Z_{II}^{j}\cup Z_{III}^{j}$ and
taking subsequence of $\widehat{\xi _{j}}$, we can assume $\widehat{\xi _{j}}
$ $C^{1}$-converges to $\widehat{\xi _{\infty }}$ on a \emph{fixed} $%
Z_{I}\cup Z_{II}\cup Z_{III}$ (this is guranteed by $C^{0}$ convergence from
our adiabatic convergence definition, and elliptic estimate on a length $5$
cylinder containing $Z_{I}^{j}\cup Z_{II}^{j}\cup Z_{III}^{j}$), which
satisfies%
\begin{eqnarray*}
\left\vert \widehat{\xi _{\infty }}\right\vert _{L^{\infty }\left( Z_{I}\cup
Z_{II}\cup Z_{III}\right) } &=&1, \\
\left\Vert d\widehat{\xi _{\infty }}\right\Vert _{L^{2}\left( Z_{II}\right)
} &\geq &\gamma \left( 4\pi \upsilon \right) \cdot \left( \left\Vert d%
\widehat{\xi _{\infty }}\right\Vert _{L^{2}\left( Z_{I}\right) }+\left\Vert d%
\widehat{\xi _{\infty }}\right\Vert _{L^{2}\left( Z_{III}\right) }\right) ,
\\
\left( \frac{\partial }{\partial \tau }+J_{0}\frac{\partial }{\partial t}%
\right) \widehat{\xi _{\infty }} &=&0.
\end{eqnarray*}%
Then $\widehat{\xi _{\infty }}$ is a nonzero holomorphic function by the
first and third identity. We write $\widehat{\xi _{\infty }}\left( \tau
,t\right) $ in Fourier series
\begin{equation*}
\widehat{\xi _{\infty }}\left( \tau ,t\right) =\Sigma _{k=-\infty }^{\infty
}a_{k}e^{2\pi k\tau }e^{2\pi kit}\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ with }\left\Vert \widehat{\xi
_{\infty }}\right\Vert _{L^{2}\left( Z_{I}\cup Z_{II}\cup Z_{III}\right)
}\leq 3,
\end{equation*}%
where the $a_{k}$'s are constant vectors in $\mathbb{C}^{n}$. We can
explicitly compute
\begin{equation}
\left\Vert d\widehat{\xi _{\infty }}\right\Vert _{L^{2}\left( \left[ a,b%
\right] \times S^{1}\right) }^{2}=\Sigma _{k=-\infty }^{\infty }4\pi
^{2}k^{2}\left\vert a_{k}\right\vert ^{2}\cdot \int_{a}^{b}e^{4\pi \tau
}d\tau . \label{energy}
\end{equation}%
Multiplying $\left( \ref{identity}\right) $ by $e^{4\pi k}$ and letting $%
c=4\pi $ there, we have
\begin{equation}
\int_{k}^{k+1}e^{4\pi \tau }d\tau =\gamma \left( 4\pi \right) \left[
\int_{k-1}^{k}e^{4\pi \tau }d\tau +\int_{k+1}^{k+2}e^{4\pi \tau }d\tau %
\right] . \label{identity-shift}
\end{equation}%
By $\left( \ref{energy}\right) ,\left( \ref{identity-shift}\right) $ we see%
\begin{equation*}
\left\Vert d\widehat{\xi _{\infty }}\right\Vert _{L^{2}\left( Z_{II}\right)
}=\gamma \left( 4\pi \right) \cdot \left( \left\Vert d\widehat{\xi _{\infty }%
}\right\Vert _{L^{2}\left( Z_{I}\right) }+\left\Vert d\widehat{\xi _{\infty }%
}\right\Vert _{L^{2}\left( Z_{III}\right) }\right) .
\end{equation*}%
This contradicts with
\begin{equation*}
\left\Vert d\widehat{\xi _{\infty }}\right\Vert _{L^{2}\left( Z_{II}\right)
}\geq \gamma \left( 4\pi \upsilon \right) \cdot \left( \left\Vert d\widehat{%
\xi _{\infty }}\right\Vert _{L^{2}\left( Z_{I}\right) }+\left\Vert d\widehat{%
\xi _{\infty }}\right\Vert _{L^{2}\left( Z_{III}\right) }\right)
\end{equation*}%
since $\gamma \left( 4\pi \right) <\gamma \left( 4\pi \upsilon \right) $.
The proof of $\left( \ref{local-elliptic-estimate-h}\right) $ is similar.
The crucial point is that $\widetilde{\xi _{j}}$ contains no $0$-mode, so
the rescaled sequence \ $\widehat{\widetilde{\xi _{j}}}:=$ $\widetilde{\xi
_{j}}/\left\Vert \widetilde{\xi _{j}}\right\Vert _{L^{\infty }\left(
Z_{I}^{j}\cup Z_{II}^{j}\cup Z_{III}^{j}\right) }$ and its limit \ $\widehat{%
\widetilde{\xi _{\infty }}}$ \ are in the higher mode space. Since\ $%
\widehat{\widetilde{\xi _{\infty }}}$ $\ $is holomorphic, writing it in
Fourier series $\widehat{\widetilde{\xi _{\infty }}}\left( \tau ,t\right)
=\Sigma _{k\neq 0}b_{k}e^{2\pi k\tau }e^{2\pi kit}$, we have%
\begin{equation*}
\left\Vert \widehat{\widetilde{\xi _{\infty }}}\right\Vert _{L^{2}\left( %
\left[ a,b\right] \times S^{1}\right) }^{2}=\Sigma _{k\neq 0}\left\vert
b_{k}\right\vert ^{2}\cdot \int_{a}^{b}e^{4\pi \tau }d\tau .
\end{equation*}%
This is similar to $\left( \ref{energy}\right) $, and the remaining steps
are the same. We omit the details.
\end{proof}
Combining the above Lemma and $\left( \ref{L2-exp-decay}\right) $ we have
\begin{cor}
For any $0<\upsilon <1$, there exists $N_{0}=N_{0}\left( \upsilon \right) $
depending on $\upsilon$, such that for all $j>N_{0}$ and $\left[ \tau,\tau+1%
\right] \subset \left[ -\tau \left( \varepsilon _{j}\right) ,\tau \left(
\varepsilon _{j}\right) \right] $, we have
\begin{equation*}
\int_{\left[ \tau,\tau+1\right] \times S^{1}}\left\vert d\xi _{j}\right\vert
^{2}\leq e^{-4\pi \upsilon \left( \tau \left( \varepsilon _{j}\right)
-\left\vert \tau \right\vert \right) }\left[ \int_{\left[ -\tau \left(
\varepsilon _{j}\right) -1,-\tau \left( \varepsilon _{j}\right) \right]
\times S^{1}}\left\vert d\xi _{j}\right\vert ^{2}+\int_{\left[ \tau \left(
\varepsilon _{j}\right) ,\tau \left( \varepsilon _{j}\right) +1\right]
\times S^{1}}\left\vert d\xi _{j}\right\vert ^{2}\right] ,
\end{equation*}
and for $\left[ \tau,\tau+1\right] \subset \left[ -R \left( \varepsilon
_{j}\right) ,R \left( \varepsilon _{j}\right) \right] $, we have
\begin{equation*}
\int_{\left[ \tau,\tau+1\right] \times S^{1}}\left\vert \widetilde\xi
_{j}\right\vert ^{2}\leq e^{-4\pi \upsilon \left( R \left( \varepsilon
_{j}\right) -\left\vert \tau \right\vert \right) }\left[ \int_{\left[ -R
\left( \varepsilon _{j}\right) -1,-R \left( \varepsilon _{j}\right) \right]
\times S^{1}}\left\vert \widetilde\xi _{j}\right\vert ^{2}+\int_{\left[ R
\left( \varepsilon _{j}\right) ,R \left( \varepsilon _{j}\right) +1\right]
\times S^{1}}\left\vert \widetilde\xi _{j}\right\vert ^{2}\right] .
\end{equation*}
\end{cor}
>From these results and standard ellitpic estimate on the cylinder $\left[
\tau -\frac{1}{2},\tau +\frac{1}{2}\right] \times S^{1}$, we obtain the
following pointwise exponential decay estimate of $\xi _{j}$.
\begin{cor}
For any $0<\upsilon <1$, there exists $N_{0}=N_{0}\left( \upsilon \right) $
depending on $\upsilon$, such that for all $j>N_{0}$ and $\tau \in \left[
-\tau \left( \varepsilon _{j}\right) ,\tau \left( \varepsilon _{j}\right) %
\right] $, we have%
\begin{equation*}
\left\vert {\Greekmath 0272} \xi _{j}\right\vert \leq Ce^{-2\pi \upsilon \left( \tau
\left( \varepsilon _{j}\right) -\left\vert \tau \right\vert \right) }\left(
\left\Vert d\xi _{j}\right\Vert _{L^{2}\left( \left[ -\tau \left(
\varepsilon _{j}\right) -1,-\tau \left( \varepsilon _{j}\right) \right]
\times S^{1}\right) }+\left\Vert d\xi _{j}\right\Vert _{L^{2}\left( \left[
\tau \left( \varepsilon _{j}\right) ,\tau \left( \varepsilon _{j}\right) +1%
\right] \times S^{1}\right) }\right) ,
\end{equation*}
and for $\tau \in \left[ -R \left( \varepsilon _{j}\right) ,R \left(
\varepsilon _{j}\right) \right] $, we have
\begin{equation*}
\left\vert \widetilde{\xi _{j}}\right\vert \leq Ce^{-2\pi \upsilon \left(
R\left( \varepsilon _{j}\right) -\left\vert \tau \right\vert \right) }\left(
\left\Vert \widetilde{\xi _{j}}\right\Vert _{L^{2}\left( \left[ -R\left(
\varepsilon _{j}\right) -1,-R\left( \varepsilon _{j}\right) \right] \times
S^{1}\right) }+\left\Vert \widetilde{\xi _{j}}\right\Vert _{L^{2}\left( %
\left[ R\left( \varepsilon _{j}\right) ,R\left( \varepsilon _{j}\right) +1%
\right] \times S^{1}\right) }\right) .
\end{equation*}%
The $\upsilon $ can be made arbitrarily close to $1$.
\end{cor}
We fix a $\upsilon $ in the above lemma, such that $\frac{p-1}{\delta }%
\upsilon \geq 1$ . This is always possible since $\frac{p-1}{\delta }>1$. We
assume $\delta _{j}\leq \varepsilon _{j}$ from now on. We estimate $\xi _{j}$
on 4 regions of ${\mathbb{R}}\times S^{1}$.
\textbf{1. }We first study the Banach norm $\left\Vert \xi _{j}\right\Vert
_{\varepsilon _{j}}$ on the region $\Omega _{\pm \zeta _{0}}\left(
\varepsilon _{j}\right) $. From the exponential decay estimate we have for $%
\left\vert \tau \right\vert \leq \tau \left( \varepsilon _{j}\right)
-h\left( \zeta _{0}\right) $ that
\begin{eqnarray}
\left\vert {\Greekmath 0272} \xi _{j}\left( \tau ,t\right) \right\vert &\leq &Ce^{2\pi
\upsilon \left[ \tau -\left( R\left( \varepsilon _{j}\right) +\frac{p-1}{%
\delta }S\left( \varepsilon _{j}\right) -h\left( \zeta _{0}\right) \right) %
\right] }\left\Vert \xi _{j}\right\Vert _{L^{2}\left( \left[ \tau \left(
\varepsilon _{0}\right) -h\left( \zeta _{0}\right) ,\tau \left( \varepsilon
_{0}\right) \right] \times S^{1}\right) } \notag \\
&\leq &Ce^{2\pi \upsilon \left( \tau -R\left( \varepsilon _{j}\right)
\right) }\left( \frac{\varepsilon _{j}}{l}\right) ^{\frac{p-1}{\delta }%
\upsilon }\left( \frac{1}{\zeta _{0}}\right) ^{\upsilon }\delta _{j}.
\label{high-mode-exp-decay}
\end{eqnarray}%
Integrating ${\Greekmath 0272} \xi _{j}\left( \tau ,t\right) $ from $R\left(
\varepsilon _{j}\right) $ to $\tau $, we have%
\begin{equation}
\left\vert \xi _{j}\left( \tau ,t\right) -\left( \xi _{j}\left(
R(\varepsilon _{j}),t\right) \right) _{0}\right\vert \leq Ce^{2\pi \upsilon
\left( \tau -R\left( \varepsilon _{j}\right) \right) }\left( \frac{%
\varepsilon _{j}}{l}\right) ^{\frac{p-1}{\delta }\upsilon }\left( \frac{1}{%
\zeta _{0}}\right) ^{\upsilon }\delta _{j}. \label{difference-exp-decay}
\end{equation}%
Therefore by $\left( \ \ref{high-mode-exp-decay}\right) $ and $\left( \ref%
{difference-exp-decay}\right) $, the estimate of $\left\Vert \xi
_{j}\right\Vert _{\varepsilon _{j}}$ restricted on the region%
\begin{equation*}
\left[ R\left( \varepsilon _{j}\right) ,R\left( \varepsilon _{j}\right) +%
\frac{p-1}{\delta }S\left( \varepsilon _{j}\right) -h\left( \zeta
_{0}\right) \right] \times S^{1}
\end{equation*}%
is
\begin{eqnarray*}
&&\int_{R(\varepsilon _{j})}^{R(\varepsilon _{j})+\frac{p-1}{\delta }S\left(
\varepsilon _{j}\right) -h\left( \zeta _{0}\right) }\int_{0}^{1}\left(
\left\vert \xi _{j}\left( \tau ,t\right) -\left( \xi _{j}\left( R\left(
\varepsilon _{j}\right) ,t\right) \right) _{0}\right\vert ^{p}+\left\vert
{\Greekmath 0272} \xi _{j}\left( \tau ,t\right) \right\vert ^{p}\right) \\
&&\cdot e^{2\pi \delta \left( -\tau +R(\varepsilon _{j})+\frac{p-1}{\delta }%
S\left( \varepsilon _{j}\right) \right) }dtd\tau \\
&\leq &C\delta _{j}^{p}\left( \frac{1}{\zeta _{0}}\right) ^{\upsilon
p}\int_{0}^{\frac{p-1}{\delta }S\left( \varepsilon _{j}\right) }\left( \frac{%
\varepsilon _{j}}{l}\right) ^{\frac{p-1}{\delta }\upsilon p}e^{\upsilon
ps}\cdot e^{-\delta s}\left( \frac{l}{\varepsilon _{j}}\right) ^{p-1}d\tau
\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ \ \ (where }s=2\pi \left( \tau -R\left( \varepsilon _{j}\right)
\right) \RIfM@\expandafter\text@\else\expandafter\mbox\fi{)} \\
&\leq &C\delta _{j}^{p}\left( \frac{1}{\zeta _{0}}\right) ^{\upsilon p}\cdot
\left( \frac{\varepsilon _{j}}{l}\right) ^{\frac{p-1}{\delta }\upsilon
p}\left( \frac{l}{\varepsilon _{j}}\right) ^{p-1}\cdot \frac{1}{\upsilon
p-\delta }e^{\left( \upsilon p-\delta \right) \cdot 2\pi \frac{p-1}{\delta }%
S\left( \varepsilon _{j}\right) } \\
&\leq &\frac{1}{\upsilon p-\delta }C\delta _{j}^{p}\left( \frac{1}{\zeta _{0}%
}\right) ^{\upsilon p}\cdot \left( \frac{\varepsilon _{j}}{l}\right) ^{\frac{%
p-1}{\delta }\upsilon p-\left( p-1\right) -\frac{\left( \upsilon p-\delta
\right) (p-1)}{\delta }} \\
&=&\frac{1}{\upsilon p-\delta }C\delta _{j}^{p}\left( \frac{1}{\zeta _{0}}%
\right) ^{\upsilon p}\cdot 1\rightarrow 0.
\end{eqnarray*}
\textbf{2. }We consider the regions
\begin{eqnarray*}
\Phi _{\zeta _{0}}\left( \varepsilon _{j}\right) &:&=[-\left( \tau \left(
\varepsilon _{j}\right) -h\left( \zeta _{0}\right) \right) ,\tau \left(
\varepsilon _{j}\right) -h\left( \zeta _{0}\right) ]\times S^{1}. \\
\Phi _{\zeta _{0}}^{+}\left( \varepsilon _{j}\right) &:&=[-\left( \tau
\left( \varepsilon _{j}\right) -h\left( \zeta _{0}\right) +1\right) ,\tau
\left( \varepsilon _{j}\right) -h\left( \zeta _{0}\right) +1]\times S^{1}.
\end{eqnarray*}%
>From the interior Schauder estimate we have%
\begin{eqnarray}
&&\left\Vert {\Greekmath 0272} \xi _{j}\right\Vert _{C^{0}\left( \Phi _{\zeta
_{0}}\left( \varepsilon _{j}\right) \right) } \notag \\
&\leq &\left\Vert \xi _{j}\right\Vert _{C^{1,\alpha }\left( \Phi _{\zeta
_{0}}\left( \varepsilon _{j}\right) \right) } \notag \\
&\leq &C\left( \left\Vert J^{\varepsilon _{j}}(\chi (\varepsilon _{j}\tau ))%
\frac{\partial \xi _{j}}{\partial t}+\varepsilon _{j}A_{j}\left( \chi
(\varepsilon _{j}\tau )\right) \cdot \xi _{j}\right\Vert _{C^{\alpha }\left(
\Phi _{\zeta _{0}}^{+}\left( \varepsilon _{j}\right) \right) }+\left\Vert
\xi _{j}\right\Vert _{C^{0}\left( \Phi _{\zeta _{0}}^{+}\left( \varepsilon
_{j}\right) \right) }\right) \notag \\
&\leq &C\left( \left\Vert B_{j}\left( \chi (\varepsilon _{j}\tau ),\xi
_{j}\right) \right\Vert _{C^{0}\left( \Phi _{\zeta _{0}}^{+}\left(
\varepsilon _{j}\right) \right) }+\left\Vert E_{\varepsilon _{j}}\left( \tau
,t\right) \right\Vert _{C^{0}\left( \Phi _{\zeta _{0}}^{+}\left( \varepsilon
_{j}\right) \right) }+\left\Vert \xi _{j}\right\Vert _{C^{0}\left( \Phi
_{\zeta _{0}}^{+}\left( \varepsilon _{j}\right) \right) }\right) \notag \\
&\leq &C\left( \left\Vert {\Greekmath 0272} \xi _{j}\right\Vert _{C^{0}\left( \Phi
_{\zeta _{0}}^{+}\left( \varepsilon _{j}\right) \right) }\left\Vert \xi
_{j}\right\Vert _{C^{0}\left( \Phi _{\zeta _{0}}^{+}\left( \varepsilon
_{j}\right) \right) }+\left\Vert \xi _{j}\right\Vert _{C^{0}\left( \Phi
_{\zeta _{0}}^{+}\left( \varepsilon _{j}\right) \right) }^{2}\right. \notag
\\
&&\left. +\left\Vert \xi _{j}\right\Vert _{C^{0}\left( \Phi _{\zeta
_{0}}^{+}\left( \varepsilon _{j}\right) \right) }\right) , \label{grad-xi}
\end{eqnarray}%
where the last inequality is because $B$ is quadratic, and $\left\Vert
E_{\varepsilon _{j}}\left( \tau ,t\right) \right\Vert _{C^{0}\left( \Phi
_{\zeta _{0}}^{+}\left( \varepsilon _{j}\right) \right) }\leq C\left\Vert
\xi _{j}\right\Vert _{C^{0}\left( \Phi _{\zeta _{0}}^{+}\left( \varepsilon
_{j}\right) \right) }$ from $\left( \ref{E-error}\right) $. Using the $%
C^{\infty }$ uniform convergence of $\xi _{j}$ outside $\Phi _{\zeta
_{0}}\left( \varepsilon _{j}\right) $, we have
\begin{equation*}
\left\Vert {\Greekmath 0272} \xi _{j}\right\Vert _{C^{0}\left( \Phi _{\zeta
_{0}}^{+}\left( \varepsilon _{j}\right) \right) }\leq \left\Vert {\Greekmath 0272} \xi
_{j}\right\Vert _{C^{0}\left( \Phi _{\zeta _{0}}\left( \varepsilon
_{j}\right) \right) }+\delta _{j}\RIfM@\expandafter\text@\else\expandafter\mbox\fi{.}
\end{equation*}%
Plugging in $\left( \ref{grad-xi}\right) $ and noting $\left\Vert \xi
_{j}\right\Vert _{C^{0}\left( \Phi _{\zeta _{0}}^{+}\left( \varepsilon
_{j}\right) \right) }\leq \delta _{j}$, we have
\begin{equation*}
\ \left( 1-C\delta _{j}\right) \left\Vert {\Greekmath 0272} \xi _{j}\right\Vert
_{C^{0}\left( \Phi _{\zeta _{0}}\left( \varepsilon _{j}\right) \right) }\leq
C\left( \delta _{j}^{2}+\delta _{j}^{2}+\delta _{j}\right) ,
\end{equation*}%
so for $\delta _{j}<\min \left\{ \frac{1}{2C},1\right\} $ we have%
\begin{equation*}
\left\Vert {\Greekmath 0272} \xi _{j}\right\Vert _{C^{0}\left( \Phi _{\zeta _{0}}\left(
\varepsilon _{j}\right) \right) }\leq 2C\left( 3\delta _{j}\right) =6C\delta
_{j}.
\end{equation*}
\textbf{3. }Then we study the equation of $\xi _{j}$ on the region%
\begin{equation*}
\Theta \left( \varepsilon _{j}\right) :=[-R(\varepsilon _{j}),R(\varepsilon
_{j})]\times S^{1}.
\end{equation*}%
For the higher mode $\widetilde{\xi _{j}}$, by $\left( \ref%
{high-mode-exp-decay}\right) $ we have
\begin{equation}
\left\vert {\Greekmath 0272} \widetilde{\xi _{j}}\left( \tau ,t\right) \right\vert \leq
\left\vert {\Greekmath 0272} \xi _{j}\left( \tau ,t\right) \right\vert \leq C\delta
_{j}\left( \frac{1}{\zeta _{0}}\right) ^{\upsilon }\left( \frac{\varepsilon
_{j}}{l}\right) ^{\frac{p-1}{\delta }\upsilon }e^{2\pi \upsilon \left( \tau
-R\left( \varepsilon _{j}\right) \right) }\leq C\delta _{j}\left( \frac{1}{%
\zeta _{0}}\right) ^{\upsilon }\varepsilon _{j}. \label{eq:Schauderxibar}
\end{equation}%
where in the last inequality we have used that $\frac{p-1}{\delta }\upsilon
>1$ and $\left\vert \tau \right\vert \leq R\left( \varepsilon _{j}\right) $.
We notice that on $\Theta \left( \varepsilon \right) =\left[ -R\left(
\varepsilon \right) ,R\left( \varepsilon \right) \right] \times S^{1}$, the
weighting function $\varepsilon ^{1-p}$ dominates the power weight $%
\left\Vert \cdot \right\Vert _{W_{\rho _{\varepsilon }}^{1,p}}$, up to
constant factor $\left( 2l\right) ^{\delta }$, because%
\begin{equation*}
\rho _{\varepsilon }\left( \tau \right) =\varepsilon ^{1-p+\delta }(1+|\tau
|)^{\delta }\leq \varepsilon ^{1-p+\delta }(2l/\varepsilon )^{\delta
}=\left( 2l\right) ^{\delta }\varepsilon ^{1-p}.
\end{equation*}%
So for higher mode $\widetilde{\xi _{j}}$ we obtain
\begin{eqnarray*}
\left\Vert \widetilde{\xi _{j}}\right\Vert _{W_{\beta _{\delta ,\varepsilon
_{j}}}^{1,p}\left( \Theta \left( \varepsilon _{j}\right) \right) }^{p} &\leq
&\int_{-R(\varepsilon _{j})}^{R(\varepsilon _{j})}\int_{0}^{1}\left(
\left\vert \widetilde{\xi _{j}}\right\vert ^{p}+|{\Greekmath 0272} \widetilde{\xi _{j}}%
|^{p}\right) \left( 2l\right) ^{\delta }\varepsilon _{j}^{1-p}\,dt\,d\tau \\
&\leq &\left( 2l\right) ^{\delta }\int_{-R(\varepsilon _{j})}^{R\left(
\varepsilon _{j}\right) }2\left( C\delta _{j}\left( \frac{1}{\zeta _{0}}%
\right) ^{\upsilon }\varepsilon _{j}\right) ^{p}\varepsilon _{j}^{1-p}\,d\tau
\\
&=&2C^{p}\left( 2l\right) ^{\delta }\left( \frac{1}{\zeta _{0}}\right)
^{\upsilon p}\delta _{j}^{p}\rightarrow 0.
\end{eqnarray*}%
For the $0$-mode $\left( \xi _{j}\right) _{0}$, noticing that for $%
\left\vert \tau \right\vert \leq l/\varepsilon $ the error term is $0$, from
the equation of $\left( \xi \right) _{0}$ we have%
\begin{eqnarray}
\left\vert {\Greekmath 0272} \left( \xi _{j}\right) _{0}\left( \tau \right) \right\vert
&=&\left\vert \frac{\partial }{\partial \tau }\left( \xi _{j}\right)
_{0}\right\vert =\left\vert \varepsilon _{j}A\left( \chi \left( \varepsilon
_{j}\tau \right) \right) \left( \xi _{j}\right) _{0}+\left( B\left( \xi
_{j}\right) \right) _{0}\right\vert \notag \\
&\leq &C\left( \varepsilon _{j}\delta _{j}+\delta _{j}^{2}\right) \RIfM@\expandafter\text@\else\expandafter\mbox\fi{(}%
\because B\left( \xi _{j}\right) \RIfM@\expandafter\text@\else\expandafter\mbox\fi{ is quadratic} \RIfM@\expandafter\text@\else\expandafter\mbox\fi{)} \notag \\
&\leq &2C\varepsilon _{j}\delta _{j}\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ ($\because $ }\delta _{j}\leq
\varepsilon _{j}\RIfM@\expandafter\text@\else\expandafter\mbox\fi{).} \label{xi-0-control}
\end{eqnarray}%
Therefore
\begin{eqnarray*}
\left\Vert \left( \xi _{j}\right) _{0}\right\Vert _{W_{\varepsilon
_{j}}^{1,p}\left( \Theta \left( \varepsilon _{j}\right) \right) }^{p}
&=&\int_{-R(\varepsilon _{j})}^{R(\varepsilon _{j})}\left( \varepsilon
_{j}\left\vert \left( \xi _{j}\right) _{0}\right\vert ^{p}+\varepsilon
_{j}^{1-p}\,|{\Greekmath 0272} \left( \xi _{j}\right) _{0}|^{p}\right) d\tau \\
&\leq &\int_{-R(\varepsilon _{j})}^{R(\varepsilon _{j})}\left( \varepsilon
_{j}\delta _{j}^{p}+\varepsilon _{j}^{1-p}\,\left( 2C\varepsilon _{j}\delta
_{j}\right) ^{p}\right) d\tau \\
&\leq &C\left( l\delta _{j}^{p}+R\left( \varepsilon _{j}\right) \varepsilon
_{j}\delta _{j}^{p}\right) =Cl\delta _{j}^{p}\rightarrow 0.
\end{eqnarray*}%
where the second inequality is by $\left( \ref{xi-0-control}\right) $.
By Sobolev embeding $\left\vert \left( \xi _{j}\left( \pm l/\varepsilon
_{j}\right) \right) _{0}\right\vert \leq C\left\Vert \left( \xi _{j}\right)
_{0}\right\Vert _{W_{\varepsilon _{j}}^{1,p}\left( \Theta \left( \varepsilon
_{j}\right) \right) }{}\leq C\delta _{j}.$
Combining these we have
\begin{eqnarray*}
\left. \left\Vert \xi _{j}\right\Vert _{\varepsilon _{j}}\right\vert
_{\Theta \left( \varepsilon _{j}\right) \cup \Omega _{\pm \zeta _{0}}\left(
\varepsilon _{j}\right) } &=&\left\Vert \widetilde{\xi _{j}}\right\Vert
_{W_{\beta _{\delta ,\varepsilon _{j}}}^{1,p}\left( \Theta \left(
\varepsilon _{j}\right) \cup \Omega _{\pm \zeta _{0}}\left( \varepsilon
_{j}\right) \right) }+\left\Vert \left( \xi _{j}\right) _{0}\right\Vert
_{W_{\varepsilon _{j}}^{1,p}\left( \Theta \left( \varepsilon _{j}\right)
\cup \Omega _{\pm \zeta _{0}}\left( \varepsilon _{j}\right) \right) }{} \\
&&+\left\vert \left( \xi _{j}\left( \pm l/\varepsilon _{j}\right) \right)
_{0}\right\vert \\
&\leq &C\delta _{j}\rightarrow 0,
\end{eqnarray*}%
where the constant $C$ is uniform for all $0<\varepsilon _{j}\leq
\varepsilon _{0}$ and $l\geq l_{0}$.
\textbf{4. }Outside the region $\Theta \left( \varepsilon _{j}\right) \cup
\Omega _{\pm \zeta _{0}}\left( \varepsilon _{j}\right) $, namely for $%
\left\vert \tau \right\vert >\tau \left( \varepsilon _{j}\right) -h\left(
\zeta _{0}\right) $, by the $C^{\infty }$ uniform convergence of $u_{j}$ to $%
u_{\pm }^{\varepsilon _{j}}$ it is easy to see
\begin{equation*}
\left. \left\Vert \xi _{j}\right\Vert _{\varepsilon _{j}}\right\vert
_{\Sigma _{\varepsilon _{j}}\backslash \left( \Theta \left( \varepsilon
_{j}\right) \cup \Omega _{\pm \zeta _{0}}\left( \varepsilon _{j}\right)
\right) }=\left. \left\Vert \xi _{j}\right\Vert _{\varepsilon
_{j}}\right\vert _{\Sigma _{\pm }\backslash U\pm \left( \zeta _{0}\right)
}\rightarrow 0.
\end{equation*}%
This finishes the proof
\begin{equation*}
\Vert \xi _{j}\Vert _{\varepsilon _{j}}\rightarrow 0
\end{equation*}%
and so the proof of Proposition \ref{prop:surjective}.
\section{Variants of Adiabatic gluing}
\label{sec:variants}
In this section, we discuss various cases to which similar adiabatic gluing
construction can be applied. Since the necessary analysis will be small
modifications of the current constructions, we will be brief in our
discussion.
\subsection{Pearl complex in Hamiltonian case}
\label{subsec:ham-pearl}
Let $f:M\rightarrow {\mathbb{R}}$ be a Morse function. The the Floer
equation for the Hamiltonian $\varepsilon f$ is%
\begin{equation} \label{Floer_e}
\frac{\partial u}{\partial \tau }+J(u)\left(\frac{\partial u}{\partial t}%
-\varepsilon X_{f}(u)\right)=0
\end{equation}%
for $u:{\mathbb{R\times }}S^{1}\rightarrow M$ with asymptote $u\left( \pm
\infty ,t\right) =z_{\pm }\left( t\right) $, where $z_{\pm }\left( t\right) $
are Hamiltonian $1$-periodic orbits of $\varepsilon f$. In the papers \cite%
{oh:dmj}, \cite{oh:adiabatic}, the first named author studied the adiabatic
degeneration of the moduli space of solutions satisfying the above equation
as $\varepsilon \rightarrow 0$. \cite{MT} studied similar adiabatic
degeneration for twisted holomorphic sections in Hamiltonian $S^{1}$%
-manifolds. The limiting moduli space as $\varepsilon =0$ consists of
sphere-flow-sphere configurations which Biran and Cornea call
\textquotedblleft pearl complexs\textquotedblright and is defined as the
following
\begin{defn}
The configuration
\begin{equation*}
u:=\left( p,\chi _{-\infty },u_{1,}\chi _{1},u_{2},\chi _{2},\cdots
,u_{k},\chi _{\infty },q\right)
\end{equation*}
is called a pearl configuration if $u_{i}:S^{2}\cong {\mathbb{R}}\times
S^{1}\rightarrow M$ are $J$-holomorphic spheres with marked points $%
u_{i}\left( o_{\pm }\right) $ where $o_{\pm }=\left\{ \pm \infty \right\}
\times S^{1}$, and each $\chi _{i}:\left[ -l_{i},l_{i}\right] \rightarrow M$
is a gradient segment of the Morse function $f$ connecting $u_{i}\left(
o_{+}\right) $ to $u_{i+1}\left( o_{-}\right) $, $\chi _{-\infty }$
connecting the critical point $p$ to $u_{1}\left( o_{-}\right) $ and $\chi
_{\infty }$ connecting $u_{k}\left( o_{+}\right) $ to the critical point $q$.
\end{defn}
We define the moduli space
\begin{eqnarray*}
\mathcal{M}_{2}\left( M,J;A_{i}\right) &=&\Big\{\left( u_{i},o_{\pm }\right)
|u_{i}:S^{2}\cong {\mathbb{R}}\times S^{1}\rightarrow M,o_{\pm }\in S^{2}, \\
&{}&\quad \overline{\partial }_{J}u_{i}=0,\left[ u_{i}\right] =A_{i}\in
H_{2}\left( M,\mathbb{Z}\right) \Big\}\Big/{\mathbb{R}}\times S^{1}
\end{eqnarray*}%
where the last ${\mathbb{R}}\times S^{1}$ is the automorphism group ${%
\mathbb{R}}$-translation and $S^{1}$ rotation, and the evaluation maps%
\begin{equation*}
ev_{\pm }^{i}:\mathcal{M}_{2}\left( M,J;A_{i}\right) \rightarrow M,\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ \ }%
u_{i}\rightarrow u_{i}\left( o_{\pm }\right) .
\end{equation*}%
Consider the map
\begin{eqnarray}
{}id\times \left( \Pi _{i=0}^{k-1}\phi _{f}^{2l_{i}}\circ ev_{+}^{i}\times
ev_{-}^{i+1}\right) \times id &:& \label{eq:pearl} \\
W^{u}\left( p\right) \times \Pi _{i=1}^{k}\mathcal{M}_{2}\left(
M,J;A_{i}\right) \times W^{s}\left( p\right) &\rightarrow &\Pi
_{i=1}^{k+1}\left( M\times M\right) , \notag \\
\left( x,u_{1},\cdots u_{k},y\right) &\rightarrow &\left( x,\Pi
_{i=1}^{k-1}\left( \phi _{f}^{2l_{i}}u_{i}\left( o_{+}\right) ,u_{i+1}\left(
o_{-}\right) \right) ,y\right) , \notag
\end{eqnarray}%
where $\phi _{f}^{2l}$ is the time-2$l$ flow of the Morse function $f$, $%
W^{u}\left( p\right) $ and $W^{s}\left( q\right) $are unstable and stable
manifolds of $p$ and $q$ respectively.
\begin{defn}
The moduli space of pearl configuration with flow length vector $\vec{l}%
:=\left( l_{1},l_{2},\cdots l_{k}\right) $ connecting $p$ to $q$ with the $J$%
-holomorphic spheres $u_{i}\left( i=1,2,\cdots k\right) $ in homology class $%
\vec{A}=\left( A_{1},A_{2},\cdots A_{k}\right) $ is defined to be
\begin{equation*}
\mathcal{M}_{pearl}^{\vec{l}}\left( p,q;f;\vec{A}\right) =\left( id\times
\left( \Pi _{i=0}^{k-1}\phi _{f}^{2l_{i}}\circ ev_{+}^{i}\times
ev_{-}^{i+1}\right) \times id\right) ^{-1}\left( \Pi _{i=1}^{k+1}\triangle
\right) ,
\end{equation*}%
where $\triangle \subset M\times M$ is the diagonal.
\end{defn}
We can give the obvious $W^{1,p}$ Banach manifold $\mathcal{B}_{pearl}^{\vec{%
l}}\left( p,q\right) $ to host $\mathcal{M}_{pearl}^{\vec{l}}\left( p,q;f;%
\vec{A}\right) $, and a natural section $e$ of the Banach bundle $\mathcal{L}%
_{pearl}^{\vec{l}}\left( p,q\right) \rightarrow $ $\mathcal{B}_{pearl}^{\vec{%
l}}\left( p,q\right) ,$
\begin{equation*}
e:\mathcal{B}_{pearl}^{\vec{l}}\left( p,q\right) \rightarrow \mathcal{L}%
_{pearl}^{\vec{l}}\left( p,q\right) ,
\end{equation*}%
such that
\begin{equation*}
\mathcal{M}_{pearl}^{\vec{l}}\left( p,q;f;\vec{A}\right) =e^{-1}\left(
0\right) .
\end{equation*}%
We let the linearization of $\ e$ at $u\in \mathcal{M}_{pearl}^{\vec{l}%
}\left( p,q;f;\vec{A}\right) $ to be $E\left( u\right) $.
We assume the pearl configuration $u:=\left( p,\chi _{-\infty },u_{1,}\chi
_{1},u_{2},\chi _{2},\cdots ,u_{k},\chi _{\infty },q\right) $ satisfies the
\textquotedblleft sphere-flow-sphere\textquotedblright\ transversality
defined as the following, which was also defined in \cite{biran-cornea} for
the case with Lagrangian boundary condition.
\begin{defn}
\label{sfs-trans}The pearl configuration $u=\left( p,\chi _{-\infty
},u_{1,}\chi _{1},u_{2},\chi _{2},\cdots ,u_{k},\chi _{\infty },q\right) $
satisfies the \textquotedblleft sphere-flow-sphere\textquotedblright\
transversality if the map
\begin{equation*}
id\times \left( \Pi _{i=0}^{k-1}\phi _{f}^{2l_{i}}\circ ev_{+}^{i}\times
ev_{-}^{i+1}\right) \times id
\end{equation*}
in $\left( \ref{eq:pearl}\right) $ is transversal to the diagonal $\Pi
_{i=1}^{k+1}\triangle \subset \Pi _{i=1}^{k+1}\left( M\times M\right) $.
\end{defn}
The \textquotedblleft sphere-flow-sphere\textquotedblright\ transversality
is known to be achievable for generic $J$ and $f$ (see \cite{biran-cornea}
for example). Small modifications of Proposition \ref{prop:dfdindex},
Corollary \ref{achieve-dfd-trans} and Proposition \ref{family-dfd-trans}
leads to the corresponding
\begin{prop}
For any $u\in \mathcal{M}_{pearl}^{\vec{l}}\left( p,q;f;\vec{A}\right) $,
the operator $E(u)$ is a Fredholm operator and
\begin{equation}
\operatorname{Index}E(u)=\mu _{f}\left( p\right) -\mu _{f}(q)+\Sigma
_{i=1}^{k}2c_{1}(A_{i}).
\end{equation}
where $\mu _{f}$ is the Morse index for critical points of $f$.
\end{prop}
\begin{cor}
Suppose that each $u_{i}\in {\mathcal{M}}_{2}(M,J;A_{i})$ is Fredholm
regular, and each gradient segment $\chi _{\pm \infty }$ belongs to a full
gradient trajectory that is Fredholm regular, then $u\in \mathcal{M}%
_{pearl}^{\vec{l}}\left( p,q;f;\vec{A}\right) $ is Fredholm regular (in the
sense that $E(u)$ is surjective) if and only if the configuration $u$
satisfies the \textquotedblleft sphere-flow-sphere\textquotedblright\
transversality in definition \ref{sfs-trans}.
\end{cor}
\begin{prop}
\label{family-sfs-trans}Suppose that each $u_{i}\in {\mathcal{M}}%
_{2}(M,J;A_{i})$ is Fredholm regular. Then there exists a dense subset of $%
f\in C^{\infty }\left( M\right) $ such that any element $u$ in
\begin{equation*}
\mathcal{M}_{pearl}^{para}\left( p,q;f;\vec{A}\right) =\bigcup_{l_{i}>0}%
\mathcal{M}_{pearl}^{\vec{l}}\left( p,q;f;\vec{A}\right)
\end{equation*}%
is Fredholm regular, in the sense that $E(u)$ is surjective. Therefore $%
\mathcal{M}_{pearl}^{para}\left( p,q;f;\vec{A}\right) $ is a smooth manifold
with dimension equal to the index of $E(u)$:
\begin{equation*}
\operatorname{dim}\mathcal{M}_{pearl}^{para}\left( p,q;f;\vec{A}\right) =\operatorname{Index}%
E(u)=\mu _{f}\left( p\right) -\mu _{f}(q)+\Sigma _{i=1}^{k}2c_{1}(A_{i})+1.
\end{equation*}
\end{prop}
\ The gluing problem from \textquotedblleft pearl
configuration\textquotedblright\ to nearby Floer trajectories was mentioned
and left as a future work in \cite{oh:newton}, \cite{MT}. Now we can study
the gluing from the pearl configuration $u=\left( p,\chi _{-\infty
},u_{1,}\chi _{1},u_{2},\chi _{2},\cdots ,u_{k},\chi _{\infty },q\right) $
to nearby Floer trajectories satisfying $\left( \ref{Floer_e}\right) $,
using the techniques from this paper. We assume $J$ and $f$ are generic such
that $E(u)$ in Proposition \ref{family-sfs-trans} is surjective. The outcome
is
\begin{thm}
For any pearl configuration $u=\left( p,\chi _{-\infty },u_{1,}\chi
_{1},u_{2},\chi _{2},\cdots ,u_{k},\chi _{\infty },q\right)$ in $\mathcal{M}%
_{pearl}^{para}\left( p,q;f;\vec{A}\right) $ whose $E(u)$ is surjective,
then for sufficiently small $\varepsilon >0$, nearby $u$ we have the
solutions $u_{\vec{\theta},\vec{d}}^{\varepsilon }\left( \tau ,t\right) $ of
Floer equation $\left( \ref{Floer_e}\right) $ parameterized by the gluing
parameters $\varepsilon >0$, $\vec{\theta}=\left( \theta _{1},\theta
_{2},\cdots ,\theta _{k}\right) \in \left( S^{1}\right) ^{k}\,$and $\vec{d}%
=\left( d_{1},d_{2},\cdots ,d_{k}\right) \in \left( {\mathbb{R}}\right) ^{k}$%
, with the asymptotes $u_{\vec{\theta},\vec{d}}^{\varepsilon }\left( -\infty
,t\right) =p$, and $u_{\vec{\theta},\vec{d}}^{\varepsilon }\left( +\infty
,t\right) =q$.
\end{thm}
\begin{proof}
\bigskip (Outline) Locally the gluing is the same, in the sense that it is
the gluing of gradient segments (with noncritical joint points) with $J$%
-holomorphic curves. The construction of approximate solution is the same,
except that for any $J$-holomorphic sphere $u_{i}$ we have the $S^{1}\times
\mathbb{R}$ family of $J$-holomorphic spheres $u_{i}\left( \tau
+d_{i},\theta +\theta _{i}\right) $ in pregluing. The $\overline{\partial }$%
-error estimate is the same, with the pair $\left( K_{\varepsilon
},J_{\varepsilon }\right) $ in the piece $\left( 3\right) $ in section \ref%
{section:d-bar-error} replaced by $\left( \varepsilon f,J_{0}\right) $ hence
$\overline{\partial }_{\left( K_{\varepsilon },J_{\varepsilon }\right) }$
replaced by $\overline{\partial }_{\left( \varepsilon f,J_{0}\right) }$. The
weight $\beta _{\delta ,\varepsilon }\left( \tau \right) $ on $\chi _{\pm
\infty }$ is not of a polynomial weight but of an exponential weight when $%
\chi _{\pm \infty }\left( \tau \right) $ approaches the critical points $p,q$%
, but by remark \ref{exp-weight-dbar-error} this doe not destroy $\overline{%
\partial }_{J,\varepsilon f}$-error estimate. The construction and estimates
of the right inverse are the same. The quadratic estimate remains the same,
where the domain Riemann surface $\Sigma _{\varepsilon }\simeq {\mathbb{R}}%
\times S^{1}$ is equipped with standard measure on disjoint cylinders $\left[
-l_{i}/\varepsilon -\frac{p-1}{\delta }S\left( \varepsilon \right)
,l_{i}/\varepsilon +\frac{p-1}{\delta }S\left( \varepsilon \right) \right]
\times S^{1}$ and $(\pm \infty ,0]\times S^{1}$, glued with standard middle
annulus of $S^{2}$ (with compact measure).
\end{proof}
Notice that we do not have the surjectivity part in this theorem, because a
sequence of solutions $u^{\varepsilon }$ of Floer equation $\left( \ref%
{Floer_e}\right) $ may develop multiple covering $J$-holomorphic spheres or
multiple covering of gradient segments in the limit, which lacks Fredholm
regularity; The bubbling sphere may also occur at any interior point on the
gradient segment $\chi $, or worse, the joint points. For that situation we
have not fully developed the gluing analysis.
\subsection{Pearl complex in Lagrangian case}
Let $\left( M,\omega \right) $ be a compact symplectic manifold with
compatible almost complex structure $J$ and $L$ be a compact Lagrangian
submanifold. For a Morse function $f:L\rightarrow {\mathbb{R}}$, we extend $%
f $ \ to a Morse function $f:M\rightarrow {\mathbb{R}}$ such that in a
Weinstein neighborhood of $L\,$\ which is symplectomorphic to $T^{\ast }M$, $%
f$ is constant along each fiber. Let $\phi _{\varepsilon f}^{t}$ be the time-%
$t$ Hamiltonian flow of $f$. Then $L_{\varepsilon f}:=\phi _{\varepsilon
f}^{1}L$ is a Lagrangian submanifold Hamiltonian isotopic to $L$. For
generic $f$, $L$ and $L_{\varepsilon f}$ transversally intersect. Similar to
the Hamiltonian case, we study the $J$-holomorphic stripe $u:{\mathbb{R}}%
\times \left[ 0,1\right] \rightarrow M$ satisfying%
\begin{eqnarray}
\frac{\partial u}{\partial \tau }+J(u)\frac{\partial u}{\partial t} &=&0,%
\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ }u\left( {\mathbb{R}},0\right) \in L\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ and }u\left( {\mathbb{R}}%
,1\right) \in L_{\varepsilon f} \notag \\
u\left( -\infty ,\cdot \right) &=&p, \, u\left( +\infty ,\cdot \right) =q
\label{Lagr-Floer-eq-e}
\end{eqnarray}%
where $p,q$ are intersections of $L$ and $L_{\varepsilon f}$. The limiting
moduli space as $\varepsilon \rightarrow 0$ consists of \textquotedblleft
pearl complexs\textquotedblright\ (or disk-flow-disk configurations) defined
as the following
\begin{defn}
The configuration
\begin{equation*}
u:=\left( p,\chi _{-\infty },u_{1,}\chi _{1},u_{2},\chi _{2},\cdots
,u_{k},\chi _{\infty },q\right)
\end{equation*}%
is called a pearl configuration if $u_{i}:D^{2}\backslash \left\{ \pm
1\right\} \cong {\mathbb{R}}\times [0,1]\rightarrow M$ are $J$-holomorphic
discs with lower and upper boundaries ending on $L$ and $L_{\varepsilon f}$
respectively, with marked points $u_{i}\left( o_{\pm }\right) $ where $%
o_{\pm }=\{\pm 1\}\in D^2=\left\{ \pm \infty \right\} \times \left[ 0,1%
\right] \in \mathbb{R}\times [0,1]$, and each $\chi _{i}:\left[ -l_{i},l_{i}%
\right] \rightarrow L$ is a gradient segment of the Morse function $f$
connecting $u_{i}\left( o_{+}\right) $ to $u_{i+1}\left( o_{-}\right) $, $%
\chi _{-\infty }$ connecting the critical point $p$ to $u_{1}\left(
o_{-}\right) $ and $\chi _{\infty }$ connecting $u_{k}\left( o_{+}\right) $
to the critical point $q$.
\end{defn}
Suppose that each $u_{i}$ is Fredholm regular. The disk-flow-disk
transversality has been defined in \cite{biran-cornea} and proven to be
achievable for generic $\left( J,f\right) $. Parallel to the previous
section, we can define the pearl complex moduli space $\mathcal{M}%
_{pearl}^{para}\left( p,q;L,f;\vec{A}\right) $, establish the index formula
for $u$ in such moduli space, and show $E\left( u\right) $ is surjective for
generic $\left( J,f\right) $. These are more or less standard so we omit
details (See \cite{biran-cornea} for precise definitions of pearl complex
moduli space and index formula). The following is the gluing theorem.
\begin{thm}
For any pearl configuration $u=\left( p,\chi _{-\infty },u_{1,}\chi
_{1},u_{2},\chi _{2},\cdots ,u_{k},\chi _{\infty },q\right)$ in $\mathcal{M}%
_{pearl}^{para}\left( p,q;L,f;\vec{A}\right) $ whose $E(u)$ is surjective,
then for sufficiently small $\varepsilon >0$, nearby $u$ we have $J$%
-holomorphic stripes $u_{\vec{d}}^{\varepsilon }\left( \tau ,t\right) $ of
equation $\left( \ref{Lagr-Floer-eq-e}\right) $ parameterized by the gluing
parameters $\varepsilon >0$ and $\vec{d}=\left( d_{1},d_{2},\cdots
,d_{k}\right) \in \left( {\mathbb{R}}\right) ^{k}$, with the asymptotes $u_{%
\vec{d}}^{\varepsilon }\left( -\infty ,t\right) =p$, and $u_{\vec{d}%
}^{\varepsilon }\left( +\infty ,t\right) =q$.
\end{thm}
\begin{proof}
(Outline) The equation $\left( \ref{Lagr-Floer-eq-e}\right) $ can be changed
into the following form which is very similar to the Hamiltonian case: For
any $u:{\mathbb{R}}\times \left[ 0,1\right] \rightarrow M$, let $%
u^{\varepsilon }\left( \tau ,t\right) :=$ $\phi _{\varepsilon f}^{-t}\circ
u\left( \tau ,t\right) $. Then $u$ satisfies $\left( \ref{Lagr-Floer-eq-e}%
\right) $\ if and only if $u^{\varepsilon }$ satisfies
\begin{eqnarray}
\frac{\partial u^{\varepsilon }}{\partial \tau }+J_{t}^{\varepsilon f}\left(
u^{\varepsilon }\right) \frac{\partial u^{\varepsilon }}{\partial t}%
+\varepsilon {\Greekmath 0272} f\left( u^{\varepsilon }\right) &=&0,\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ }%
u^{\varepsilon }\left( {\mathbb{R}},0\right) \RIfM@\expandafter\text@\else\expandafter\mbox\fi{ and }u^{\varepsilon
}\left( {\mathbb{R}},1\right) \subset L, \notag \\
u^{\varepsilon }\left( -\infty ,\cdot \right) &=&p \quad \RIfM@\expandafter\text@\else\expandafter\mbox\fi{ and } \quad
u^{\varepsilon }\left( +\infty ,\cdot \right) =q. \label{LH-Floer-eq-e}
\end{eqnarray}%
where $\left\{ J_{t}^{\varepsilon f}\right\} _{0\leq t\leq 1}:=\left\{
\left( \phi _{\varepsilon f}^{t}\right) _{\ast }J\left( \phi _{\varepsilon
f}^{-t}\right) _{\ast }\right\} _{0\leq t\leq 1}$is a $1$-parameter family
of compatible almost complex structures.
>From the pearl configuration $u$ we can construct the approximate solution
for equation $\left( \ref{LH-Floer-eq-e}\right) $ from $J$-holomorphic discs
$u_{i}$ and gradient segments $\chi _{i}\left( i=1,2,\cdots ,k\right) $ and $%
\chi _{\pm \infty }$ as in the Hamiltonian case, but there is one
difference: The almost complex structure $J_{t}^{\varepsilon f}$ is $t$%
-dependent. \
The $t$-dependent $J_{t}^{\varepsilon f}$ seems to destroy the decomposition
of $0$-mode and higher modes of variation vector fields on $\chi $ as in the
Hamiltonian case, but actually here we have even better situation: the
linearized operator of $\overline{\partial }_{J_{t}^{\varepsilon
f},\varepsilon f}$ is canonically related to linearized operator of $\frac{d%
}{d\tau }-\operatorname{grad}\left( \varepsilon f\right) $, and the right inverse
bound for $D_{\chi _{\varepsilon }}\overline{\partial }_{J_{t}^{\varepsilon
f},\varepsilon f}$ can be controlled by the right inverse bound of $\frac{D}{%
d\tau }-{\Greekmath 0272} \operatorname{grad}\left( \varepsilon f\right) $ up to a uniform
constant (as in the Proposition 6.1 of \cite{FO} and Proposition 4.6 in \cite%
{oh:imrn}). Therefore we do not need to decompose the $0$-mode and higher
modes of variation vector fields on $\chi $ in the Banach manifold setting,
and the weighting function $\beta _{\varepsilon ,\delta }\left( \tau \right)
$ on the portion over $\chi _{i}\left( \varepsilon \tau \right) $ $\left(
-l_{i}/\varepsilon \leq \tau \leq l_{i}/\varepsilon \right) $ is just the $%
\varepsilon $-adiabatic weight, namely
\begin{equation*}
\left\Vert \xi \right\Vert _{W_{\varepsilon }^{1,p}[-l_{i}/\varepsilon
,l_{i}/\varepsilon ]\times \left[ 0,1\right] }^{p}=\int_{0}^{1}\int_{-l/%
\varepsilon }^{l/\varepsilon }\left( \varepsilon ^{2}\left\vert \xi
_{0}\right\vert ^{p}+\varepsilon ^{2-p}\left\vert {\Greekmath 0272} \xi _{0}\right\vert
^{p}\right) d\tau dt
\end{equation*}%
as in \cite{FO}. The weighting function $\beta _{\varepsilon ,\delta }\left(
\tau \right) $ on the portion over $u_{i}^{\varepsilon }\left( \tau \right) $
remains the same, namely $\beta _{\varepsilon ,\delta }\left( \tau \right)
=e^{\delta \left\vert \tau \right\vert }$ for large $\left\vert \tau
\right\vert $ of the ends $u_{i}^{\varepsilon }\left( \tau \right) $. The
construction of combined right inverse and uniform quadratic estimates are
similar to the Hamiltonian case.
\end{proof}
The limit of moduli spaces of $J$-holomorphic $\left( k+1\right) $-gons
ending on $\left( L_{\varepsilon f_{0}},\cdots ,L_{\varepsilon f_{k}}\right)
$\ \ as $\varepsilon \rightarrow 0$ consists of \textquotedblleft cluster
complexes\textquotedblright , which are $J$-holomorphic discs $u_{i}$ ending
on $L$ connected with \emph{gradient flow trees} $\chi _{j}$ on $L$. In \cite%
{FO}, the gluing from gradient flow trees $\chi _{j}$ to \textquotedblleft
thin\textquotedblright\ $J$-holomorphic polygons was known. The gluing of
\textquotedblleft thin\textquotedblright\ $J$-holomorphic polygon with
\textquotedblleft thick\textquotedblright\ $J$-holomorphic discs $u_{i}$
locally is the same as the case of pearl complex as above.
\bigskip Again, we do not have the surjectivity part in the gluing theorem.
The reasons are similar to Hamiltonian case. The multicovering is the
essential difficulty, and we also need the joint point to be immersed to
prove surjectivity. However, for pearl complex moduli spaces of virtual
dimension $0,1$ for monotone Lagrangian submanifolds, in Proposition 3.13 of
\cite{biran-cornea} it has been proved that the $J$-holomorphic discs in
pearl complex are Fredholm regular, simple (non-multicovering) and with
absolutely distinct, provided $\left( J,f\right) $ is generic. Then, for
simple $J$-holomorphic discs, relative version (with Lagrangian boundary
condition) of Theorem 2.6 in \cite{oh-zhu2} implies that the condition of
existence of non-immersed point on some $J$-holomorphic disk of the pearl
complex cut down the dimension of pearl complex moduli space by $%
\mathop{\kern0pt{\rm
dim}}\nolimits M-2$, provided $\left( J,f\right) $ is generic. So with
non-immersed condition the moduli space will have negative dimension if $%
\mathop{\kern0pt{\rm dim}}\nolimits M\geq 4$. When $%
\mathop{\kern0pt{\rm
dim}}\nolimits M=2$, \ $M$ is a Riemann surface, and the $J$-holomorphic
discs are just discs on surfaces $M$ with embedded boundary curve $L$. If
the $J$-holomorphic disc is simple, by Riemann mapping theorem it is
immersed. Hence we have
\begin{cor}
Let $(M,\omega )$ be monotone and $L\subset (M,\omega )$ a monotone
Lagrangian submanifold. Then for generic $\left( J,f\right) ,$ all $J$%
-holomorphic discs in pearl complex moduli spaces of virtual dimension $0,1$%
, are simple, Fredholm regular and immersed.
\end{cor}
Now with immersion condition, similar analysis as Hamiltonian case can
establish surjectivity so we have
\begin{thm}
Let $(M,\omega )$ be a monotone symplectic manifold and $L\subset (M,\omega
) $ be a monotone Lagrangian submanifold. Fix a Darboux neighborhood $U$ of $%
L$ and identify $U $ with a neighborhood of the zero section in $T^{\ast }L$%
. Consider $k+1$ Hamiltonian deformations of $L$ by autonomous Hamiltonian
functions $F_{0},\ldots ,F_{k}:M\rightarrow {\mathbb{R}}$ such that
\begin{equation*}
F_{i}=\chi f_{i}\circ \pi
\end{equation*}%
where $f_0,f_1,\ldots f_k: L\rightarrow {\mathbb{R}}$ are generic Morse
functions, $\chi $ is a cut-off function such that $\chi=1$ on $U$ and
supported nearby U. Now consider
\begin{equation*}
L_{i,\varepsilon }=\operatorname{Graph}(\varepsilon df_{i})\subset U\subset M,\quad
i=0,\ldots ,k.
\end{equation*}%
Assume transversality of $L_{i}$'s of the type given in \cite{FO}. Consider
the intersections $p_{i}\in L_{i}\cap L_{i+1}$ such that
\begin{equation*}
\mathop{\kern0pt{\rm dim}}\nolimits{\mathcal{M}}(L_{0},\ldots
,L_{k};p_{0},\ldots ,p_{k})=0,\,1.
\end{equation*}%
Then when $\varepsilon $ is sufficiently small, the moduli space ${\mathcal{M%
}}(L_{0,\varepsilon },\ldots ,L_{k,\varepsilon };p_{0},\ldots ,p_{k})$ is
diffeomorphic to the moduli space of pearl complex defined in \cite%
{biran-cornea}.
\end{thm}
\section{Discussion and an example\label{discussion}}
\subsection{The immersion condition: an example of adiabatic limit in $%
\mathbb{CP}^{n}$}
Consider the standard $\mathbb{CP}^{n}$ equipped with Fubini-Study metric $\
g_{FS}$. Given a Morse function $f:\mathbb{CP}^{n}\rightarrow \mathbb{R}$,
we consider the Floer equation with the Hamiltonian $\varepsilon f$:%
\begin{equation*}
\frac{\partial u}{\partial \tau }+J_{0}\frac{\partial u}{\partial t}%
=-\varepsilon \,{\Greekmath 0272} f\,\left( u\right) .
\end{equation*}%
Let $U_{0}\simeq \mathbb{C}^{n}$ be an affine chart, where
\begin{eqnarray*}
U_{0} &=&\left\{ \left[ z_{0}:z_{1}:\cdots :z_{n}\right] |z_{i}\in \mathbb{C}%
,\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ }z_{0}\neq 0\in \mathbb{C}\right\} \\
&=&\left\{ \left[ 1:z_{1}:\cdots :z_{n}\right] |z_{i}\in \mathbb{C}\right\} ,%
\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ \ }\left( i=1,2,\cdots ,n\right) .
\end{eqnarray*}%
Let
\begin{equation*}
\chi :\left[ -l,l\right] \mathbb{\rightarrow }U_{0}\simeq \left( \mathbb{C}%
^{n},g_{FS}\right)
\end{equation*}%
be a segment of a Morse trajectory of $f$ \ with end points $\chi \left( \pm
l\right) =p_{\pm }\in \mathbb{C}^{n}$, and assume the full Morse trajectory
extending $\chi $ to two critical endpoints is contained in the unit ball $%
B_{1}=\left\{ \left\vert z\right\vert \leq 1\right\} \subset $ $U_{0}$. Let $%
u_{\pm }:\left( \mathbb{R\times }S^{1},o_{\pm }\right) \rightarrow
U_{0}\simeq \left( \mathbb{C}^{n},g_{FS}\right) $ be two holomorphic curves,
\begin{equation*}
u_{\pm }\left( \tau ,t\right) =u\left( z\right) =A_{\pm }z^{\pm 1}+p_{\pm },
\end{equation*}%
where $A_{\pm }$ are two complex linearly independent vectors in $\mathbb{C}%
^{n}$, and
\begin{equation*}
z=e^{2\pi \left( \tau +it\right) },\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ \ \ }o_{\pm }=\left\{ -\pm \infty
\right\} \times S^{1}.
\end{equation*}%
We thus have the \textquotedblleft disk-flow-disk\textquotedblright\ element
$\left( u_{-},\chi ,u_{+}\right) $ in $\mathbb{CP}^{n}$ with immersed joint
points
\begin{equation*}
p_{\pm }=\chi \left( \pm l\right) =u_{\pm }\left( o_{\pm }\right) .
\end{equation*}
\bigskip We can construct a family of maps $u_{\varepsilon }:\mathbb{R\times
}S^{1}\rightarrow \mathbb{C}^{n+1}$ approximately satisfies the above Floer
equation when $\varepsilon $ is small. Let
\begin{equation*}
u_{\varepsilon }\left( \tau ,t\right) =\varepsilon ^{\alpha }e^{-2\pi \frac{l%
}{\varepsilon }}\left( A_{+}z+A_{-}z^{-1}\right) +\chi \left( \varepsilon
\tau \right)
\end{equation*}%
with a fixed $\alpha >0$. Then%
\begin{equation*}
\frac{\partial u_{\varepsilon }}{\partial \tau }+J_{0}\frac{\partial
u_{\varepsilon }}{\partial t}=\varepsilon {\Greekmath 0272} f\left( \chi \left(
\varepsilon \tau \right) \right) \approx \varepsilon {\Greekmath 0272} f\left(
u_{\varepsilon }\left( \tau ,t\right) \right) ,
\end{equation*}%
where $\chi \left( \varepsilon \tau \right) \approx u_{\varepsilon }\left(
\tau ,t\right) $ for $\left\vert \tau \right\vert \leq l/\varepsilon $ is
justified by the following Proposition, and for $\left\vert \tau \right\vert
>l/\varepsilon $, the following Proposition also proves $u_{\varepsilon
}\approx \varepsilon ^{\alpha }e^{-2\pi \frac{l}{\varepsilon }}\left(
A_{+}z+A_{-}z^{-1}\right) $, hence
\begin{equation*}
\frac{\partial u_{\varepsilon }}{\partial \tau }+J_{0}\frac{\partial
u_{\varepsilon }}{\partial t}\approx \left( \frac{\partial }{\partial \tau }%
+J_{0}\frac{\partial }{\partial t}\right) \left[ \varepsilon ^{\alpha
}e^{-2\pi \frac{l}{\varepsilon }}\left( A_{+}z+A_{-}z^{-1}\right) \right]
=0\approx \varepsilon {\Greekmath 0272} f\left( u_{\varepsilon }\left( \tau ,t\right)
\right) .
\end{equation*}
Although $u_{\varepsilon }$ is only an approximate solution, it is very
explicit and illustrates the mechanism of adiabatic convergence. It also
indicates the convergence rates and $L^{2}$-energy distribution of true
solutions on different regions of $\mathbb{R\times }S^{1}$.
\begin{prop}
The adiabatic limit of $u_{\varepsilon }\left( \tau ,t\right) $ as $%
\varepsilon \rightarrow 0$ is the \textquotedblleft
disk-flow-disk\textquotedblright\ configuration $\left( u_{-},\chi
,u_{+}\right) $.
\end{prop}
\begin{proof}
We recall some useful inequalities of the Fubini-Study metric $g_{FS}$ on
the affine chart $\mathbb{C}^{n}$. Let $g_{st}$ be the standard Euclidean
metric on $\mathbb{C}^{n}$ and the Euclidean norm be $\left\vert \cdot
\right\vert $. Note that $g_{FS}\left( z\right) \leq \frac{1}{1+\left\vert
z\right\vert ^{2}}g_{st}\left( z\right) \leq g_{st}\left( z\right) $ for any
$z\in \mathbb{C}^{n}$, so for any vectors $p,q\in \mathbb{C}^{n}$, we have%
\begin{eqnarray*}
\RIfM@\expandafter\text@\else\expandafter\mbox\fi{dist}_{g_{FS}}\left( p,q\right) &\leq &\left\vert p-q\right\vert , \\
\RIfM@\expandafter\text@\else\expandafter\mbox\fi{dist}_{g_{FS}}\left( p,q\right) &\leq &2\frac{\left\vert
p-q\right\vert }{\left\vert p\right\vert }\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ }\ \RIfM@\expandafter\text@\else\expandafter\mbox\fi{if }\left\vert
p-q\right\vert <\frac{\left\vert p\right\vert }{2},
\end{eqnarray*}%
where the second inequality is because both $p$ and $q$ are outside the
Euclidean ball of radius $\left\vert p\right\vert /2$, and the Fubini-Study
metric satisfies $g_{FS}\leq \frac{1}{\left\vert z\right\vert ^{2}}g_{st}$.
Let $R\left( \varepsilon \right) =l/\varepsilon $. We first check that $%
d_{H}\left( u_{\varepsilon }\left( \tau ,t\right) \left( \left[ -R\left(
\varepsilon \right) ,R\left( \varepsilon \right) \right] \times S^{1}\right)
,\chi \left( \left[ -l,l\right] \right) \right) \rightarrow 0$.
(i) For $\left\vert \tau \right\vert \leq R\left( \varepsilon \right)
=l/\varepsilon $: we have
\begin{eqnarray*}
\RIfM@\expandafter\text@\else\expandafter\mbox\fi{dist}_{g_{FS}}\left( u_{\varepsilon }\left( \tau ,t\right) ,\chi
\left( \varepsilon \tau \right) \right) &\leq &\left\vert u_{\varepsilon
}\left( \tau ,t\right) -\chi \left( \varepsilon \tau \right) \right\vert \\
&=&\left\vert \varepsilon ^{\alpha }e^{-2\pi \frac{l}{\varepsilon }}\left(
A_{+}z+A_{-}z^{-1}\right) \right\vert \\
&\leq &\varepsilon ^{\alpha }e^{-2\pi \frac{l}{\varepsilon }}\cdot \left(
\left\vert A_{+}\right\vert e^{2\pi \frac{l}{\varepsilon }}+\left\vert
A_{+}\right\vert e^{2\pi \frac{l}{\varepsilon }}\right) \\
&=&\varepsilon ^{\alpha }\left( \left\vert A_{-}\right\vert +\left\vert
A_{+}\right\vert \right) \rightarrow 0
\end{eqnarray*}%
uniformly as $\varepsilon \rightarrow 0$.
Let
\begin{equation*}
b\left( \varepsilon \right) =-\frac{1}{2\pi }\ln \varepsilon ,\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ \ }%
S\left( \varepsilon \right) =\alpha b\left( \varepsilon \right) ,
\end{equation*}
and the shift of $u_{\pm }$ be
\begin{equation*}
u_{\pm }^{\varepsilon }\left( \tau ,t\right) :=u_{\pm }\left( \tau -\pm
l/\varepsilon -\alpha b\left( \varepsilon \right) ,t\right) =\varepsilon
^{\alpha }e^{-2\pi \frac{l}{\varepsilon }}A_{\pm }z^{\pm 1}+p_{\pm }.
\end{equation*}
(ii) For $\left\vert \tau \right\vert \geq R\left( \varepsilon \right)
=l/\varepsilon $: we consider two cases where $l/\varepsilon \leq \left\vert
\tau \right\vert \leq l/\varepsilon +2\alpha b\left( \varepsilon \right) $
or $\left\vert \tau \right\vert \geq l/\varepsilon +2\alpha b\left(
\varepsilon \right) $.
If $l/\varepsilon \leq \tau \leq l/\varepsilon +2\alpha b\left( \varepsilon
\right) $, then
\begin{eqnarray*}
\RIfM@\expandafter\text@\else\expandafter\mbox\fi{dist}_{g_{FS}}\left( u_{\varepsilon }\left( \tau ,t\right)
,u_{+}^{\varepsilon }\left( \tau ,t\right) \right) &\leq &\left\vert
u_{\varepsilon }\left( \tau ,t\right) -u_{+}^{\varepsilon }\left( \tau
,t\right) \right\vert \\
&\leq &\left\vert \chi _{\varepsilon }\left( \tau \right) -p_{+}\right\vert
+\left\vert \varepsilon ^{\alpha }e^{-2\pi \frac{l}{\varepsilon }%
}A_{-}z^{-1}\right\vert \\
&\leq &\left\vert \varepsilon {\Greekmath 0272} f\right\vert _{C^{o}}\cdot \left\vert
\tau -l/\varepsilon \right\vert +\varepsilon ^{a}\left\vert A_{-}\right\vert
\\
&\leq &2\left\vert {\Greekmath 0272} f\right\vert _{C^{o}}\alpha \cdot \varepsilon
b\left( \varepsilon \right) +\varepsilon ^{a}\left\vert A_{-}\right\vert
\rightarrow 0
\end{eqnarray*}%
uniformly as $\varepsilon \rightarrow 0$.
If $\tau \geq l/\varepsilon +2\alpha b\left( \varepsilon \right) $, we have
\begin{eqnarray*}
\left\vert u_{+}^{\varepsilon }\left( \tau ,t\right) \right\vert
&=&\left\vert \varepsilon ^{\alpha }e^{-2\pi \frac{l}{\varepsilon }%
}A_{+}z\right\vert =\left\vert A_{+}\right\vert e^{2\pi \left( \tau
-l/\varepsilon -\alpha b\left( \varepsilon \right) \right) } \\
&\geq &\left\vert A_{+}\right\vert e^{2\pi b\left( \varepsilon \right)
\alpha }=\left\vert A_{+}\right\vert \varepsilon ^{-\alpha }\rightarrow
\infty , \\
\left\vert u_{\varepsilon }\left( \tau ,t\right) -u_{+}^{\varepsilon }\left(
\tau ,t\right) \right\vert &=&\left\vert \varepsilon ^{\alpha }e^{-2\pi
\frac{l}{\varepsilon }}A_{-}z^{-1}+\chi \left( \varepsilon \tau \right)
\right\vert \leq \left\vert \varepsilon ^{\alpha }e^{-2\pi \frac{l}{%
\varepsilon }}A_{-}z^{-1}\right\vert +\left\vert \chi \left( \varepsilon
\tau \right) \right\vert \\
&\leq &\varepsilon ^{a}\left\vert A_{-}\right\vert e^{-2\pi \left(
l/\varepsilon +\tau \right) }+1\rightarrow 1<\frac{\left\vert
u_{+}^{\varepsilon }\left( \tau ,t\right) \right\vert }{2},
\end{eqnarray*}%
when $\varepsilon $ is small, hence
\begin{eqnarray*}
\RIfM@\expandafter\text@\else\expandafter\mbox\fi{dist}_{g_{FS}}\left( u_{\varepsilon }\left( \tau ,t\right)
,u_{+}^{\varepsilon }\left( \tau ,t\right) \right) &\leq &2\frac{\left\vert
u_{\varepsilon }\left( \tau ,t\right) -u_{+}^{\varepsilon }\left( \tau
,t\right) \right\vert }{\left\vert u_{\varepsilon }\left( \tau ,t\right)
\right\vert } \\
&\leq &2\frac{\left\vert \varepsilon ^{\alpha }e^{-2\pi \frac{l}{\varepsilon
}}A_{-}z^{-1}+\chi \left( \varepsilon \tau \right) \right\vert +\left\vert
p_{+}\right\vert }{\left\vert \varepsilon ^{\alpha }e^{-2\pi \frac{l}{%
\varepsilon }}A_{+}z\right\vert -\left\vert \varepsilon ^{\alpha }e^{-2\pi
\frac{l}{\varepsilon }}A_{-}z^{-1}+\chi \left( \varepsilon \tau \right)
\right\vert } \\
&\leq &2\frac{\left( \varepsilon ^{\alpha }\left\vert A_{-}\right\vert
e^{-2\pi \left( l/\varepsilon +\tau \right) }+1\right) +1}{\left\vert
A_{+}\right\vert \varepsilon ^{-\alpha }-\left( \varepsilon ^{a}\left\vert
A_{-}\right\vert e^{-2\pi \left( l/\varepsilon +\tau \right) }+1\right) } \\
&\leq &2\frac{3}{\left\vert A_{+}\right\vert \varepsilon ^{-\alpha }-2}%
\rightarrow 0
\end{eqnarray*}%
uniformly as $\varepsilon \rightarrow 0$. Combining these we have for $\tau
\geq R\left( \varepsilon \right) $,%
\begin{equation*}
\RIfM@\expandafter\text@\else\expandafter\mbox\fi{dist}_{g_{FS}}\left( u_{\varepsilon }\left( \tau ,t\right)
,u_{+}^{\varepsilon }\left( \tau ,t\right) \right) \rightarrow 0,\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ }
\end{equation*}
or equivalently%
\begin{equation*}
\RIfM@\expandafter\text@\else\expandafter\mbox\fi{dist}_{g_{FS}}\left( u_{\varepsilon }\left( \tau +R\left( \varepsilon
\right) +\alpha b\left( \varepsilon \right) ,t\right) ,u_{+}\left( \tau
,t\right) \right) \rightarrow 0
\end{equation*}%
uniformly as $\varepsilon \rightarrow 0$ on $[-\alpha b\left( \varepsilon
\right) ,+\infty )\times S^{1}$, especially for\ any $[K,+\infty )\times
S^{1}$ for any fixed $K\in \mathbb{R}$.
The case when $\tau \leq -l/\varepsilon $ is similar; we have
\begin{equation*}
\RIfM@\expandafter\text@\else\expandafter\mbox\fi{dist}_{g_{FS}}\left( u_{\varepsilon }\left( \tau -R\left( \varepsilon
\right) -\alpha b\left( \varepsilon \right) ,t\right) ,u_{-}\left( \tau
,t\right) \right) \rightarrow 0
\end{equation*}
uniformly as $\varepsilon \rightarrow 0$ on $(-\infty ,\alpha b\left(
\varepsilon \right) ]\times S^{1}$, especially for\ any $(-\infty ,K]\times
S^{1}$ for any fixed $K\in \mathbb{R}$.
Next we compute the energy $E\left( u_{\varepsilon }\right) $ on $\Theta
_{\varepsilon }:=\left[ -R\left( \varepsilon \right) ,R\left( \varepsilon
\right) \right] \times S^{1}$. We have
\begin{eqnarray*}
\frac{\partial }{\partial \tau }u_{\varepsilon }\left( \tau ,t\right)
&=&2\pi \varepsilon ^{\alpha }e^{-2\pi \frac{l}{\varepsilon }}\left(
A_{+}e^{2\pi \left( \tau +it\right) }-A_{-}e^{-2\pi \left( \tau +it\right)
}\right) +\varepsilon {\Greekmath 0272} f\left( \chi \left( \varepsilon \tau \right)
\right) \\
\frac{\partial }{\partial t}u_{\varepsilon }\left( \tau ,t\right) &=&2\pi
i\varepsilon ^{\alpha }e^{-2\pi \frac{l}{\varepsilon }}\left( A_{+}e^{2\pi
\left( \tau +it\right) }-A_{-}e^{-2\pi \left( \tau +it\right) }\right) \RIfM@\expandafter\text@\else\expandafter\mbox\fi{
\ \ \ \ \ \ \ \ \ } \\
\left\vert du_{\varepsilon }\left( \tau ,t\right) \right\vert _{g_{st}}
&\leq &C\varepsilon ^{\alpha }e^{-2\pi \frac{l}{\varepsilon }}\cdot e^{2\pi
\left\vert \tau \right\vert }+C\varepsilon .
\end{eqnarray*}%
For $\left\vert \tau \right\vert \leq R\left( \varepsilon \right) $, from
the above third inequality we have%
\begin{eqnarray*}
\int_{\left[ -R\left( \varepsilon \right) ,R\left( \varepsilon \right) %
\right] \times S^{1}}\left\vert du_{\varepsilon }\right\vert
_{g_{FS}}^{2}d\tau dt &\leq &\int_{\left[ -R\left( \varepsilon \right)
,R\left( \varepsilon \right) \right] \times S^{1}}\left\vert du_{\varepsilon
}\right\vert _{g_{st}}^{2}d\tau dt \\
&\leq &2C^{2}\int_{\left[ -R\left( \varepsilon \right) ,R\left( \varepsilon
\right) \right] \times S^{1}}\left( \varepsilon ^{2\alpha }e^{-4\pi \left(
\frac{l}{\varepsilon }-\left\vert \tau \right\vert \right) }+\varepsilon
^{2}\right) d\tau dt \\
&=&2C^{2}\left[ \varepsilon ^{2\alpha }\frac{1-e^{-4\pi \frac{l}{\varepsilon
}}}{2\pi }+2l\varepsilon \right] \\
&\leq &\widetilde{C}\left( \varepsilon ^{2\alpha }+\varepsilon \right)
\rightarrow 0.
\end{eqnarray*}
\end{proof}
\bigskip If the joint points of $u_{\pm }$ are not immersed, in the next
example we will see extra family of approximate solutions of the Floer
equation beyond our pre-gluing construction. Let
\begin{equation*}
u_{\varepsilon }\left( \tau ,t\right) =\varepsilon ^{\alpha }\left[ e^{-2\pi
k\frac{l}{\varepsilon }}A_{+}z^{k}+e^{-2\pi m\frac{l}{\varepsilon }%
}A_{-}z^{-m}\right] +\beta \left( \varepsilon \right) P\left( z\right) +\chi
\left( \varepsilon \tau \right) ,
\end{equation*}%
where $k,m>0\ $are integers,\ and at least one of them $>1$, $P\left(
z\right) $ is any Laurent polynomial of intermediate degree between $z^{k%
\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ }}$ and $z^{-m}$ with $\mathbb{C}^{n}$ vector-valued coefficients,
and $\beta \left( \varepsilon \right) $ is a fast vanishing real constant
when $\varepsilon \rightarrow 0$.\ Then for fixed $l>0$, similarly we can
show $u_{\varepsilon }$ has the adiabatic limit $\left( u_{-},\chi
,u_{+}\right) $, where $u_{\pm }\left( z\right) $ are the holomorphic
spheres in $\mathbb{CP}^{n}$ that in the affine chart $U_{0}\simeq \mathbb{C}%
^{n}$,
\begin{equation*}
u_{+}\left( z\right) =A_{+}z^{k}+p_{+},\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ \ \ \ \ \ \ \ }u_{-}\left(
z\right) =A_{-}z^{-m}+p_{-}
\end{equation*}%
for $z=e^{2\pi \left( \tau +it\right) }$. But other than the approximate
solutions
\begin{equation*}
u_{\varepsilon }\left( \tau ,t\right) =\varepsilon ^{\alpha }\left[ e^{-2\pi
k\frac{l}{\varepsilon }}A_{+}z^{k}+e^{-2\pi m\frac{l}{\varepsilon }%
}A_{-}z^{-m}\right] +\chi \left( \varepsilon \tau \right)
\end{equation*}%
of the Floer equation, now we have extra family of approximate solutions
from the term $\beta \left( \varepsilon \right) P\left( z\right) $. This
gives evidence that we can not prove surjectivity if there is non-immersed
joint point.
The adiabatic degeneration with Lagrangian boundary condition can be
constructed similarly, with the Lagrangian $L=$ $\mathbb{RP}^{n}$ and $\chi
\left( \tau \right) $ inside $\mathbb{RP}^{n}$, and $u_{\varepsilon }:%
\mathbb{R\times }\left[ 0,\frac{1}{2}\right] \rightarrow U_{0}\simeq \mathbb{%
C}^{n}$,
\begin{equation*}
u_{\varepsilon }\left( \tau ,t\right) =\varepsilon ^{\alpha }\left[ e^{-2\pi
k\frac{l}{\varepsilon }}A_{+}z^{k}+e^{-2\pi m\frac{l}{\varepsilon }%
}A_{-}z^{-m}\right] +\beta \left( \varepsilon \right) P\left( z\right) +\chi
\left( \varepsilon \tau \right)
\end{equation*}%
be the approximate solution of Floer equation $\frac{\partial u}{\partial
\tau }+J_{0}\frac{\partial u}{\partial t}+\varepsilon {\Greekmath 0272} f\left(
u\right) =0$ with Lagrangian boundary condition $u\left( \mathbb{R\times }%
\left\{ 0,\frac{1}{2}\right\} \right) \subset L$.
\bigskip
|
\section{Introduction}
\section{Introduction}
\medskip
The quantum filtering theory provides a foundation of statistical inference inspired in e.g. quantum optical systems. These systems are described by continuous-time quantum stochastic differential equations. These stochastic master equations include the measurement back-action on the quantum-state.
The quantum filtering theory has been developed by Davies in the 1960s~\cite{davies1969quantum,davies1976quantum} and in its modern form by Belavkin in the 1980s~\cite{belavkin1995quantum,belavkin2004towards,belavkin1992quantum}.
\medskip
To these stochastic master equations are attached so-called quantum filters providing, from the real-time measurements, estimations of the quantum states. Robustness and convergence of such estimation process has been investigated in many papers. For example, sufficient convergence conditions, related to observability issues, are given in \cite{van-handel:proba2009} and \cite{van2006filtering}. As far as we know, general and verifiable necessary and sufficient convergence conditions do not exist yet. For links between quantum filtering and observers design on cones see~\cite{bonnabel-et-al:cdc2011a}. In this paper, we generalize a stability result for pure states (see, e.g., \cite{diosi2006coupled}) to arbitrary mixed quantum states. More precisely, we prove that the fidelity between the quantum state (that could be a mixed state) and its associated quantum-filter state is a sub-martingale: this means that in average, the estimated state tends to be closer to the system state. This does not imply its asymptotic convergence for large times. To prove such convergence, more specific analysis depending on the precise structure of the Hamiltonian, Lindbladian and measurement operators defining the system model is required. This paper can also be seen as an extension to continuous-time evolution of~\cite{rouchon2010fidelity}.
\medskip
This paper is organized as follows. In section~\ref{sec:second}, we introduce the non linear stochastic master equations driven by Wiener processes and providing the evolutions of the quantum state and of the quantum-filter state and we state the main result (Theorem~\ref{thm:main-tool}). Section~\ref{sec:one} is devoted to the proof of this result: firstly we consider an approximation via stochastic master equations driven by Poisson processes (diffusion approximation); secondly, we prove the sub-martingale property via a time discretization. In final section, we suggest some possible extensions of this work.
\section{Main result}\label{sec:second}
We will consider quantum systems of finite dimensions $1 <N<\infty.$ The state space of such a system is given by the set of density matrices
\begin{equation*}
{\mathcal D}:=\{\rho\in\CC^{N\times N}|\quad\rho=\rho^\dag,\quad\tr{\rho}=1,\quad\rho\geq 0\}.
\end{equation*}
Formally the evolution of the real state $\rho\in \mathcal D$ is described by the following stochastic master equation (cf. ~\cite{belavkin1992quantum,bouten2004stochastic,van2005feedback})
\begin{equation}\label{eq:master1}
d\rho_t=-\tfrac{i}{\hbar}[H,\rho_t]\;dt+\LL(\rho_t)\, dt+\,\Lambda(\rho_t)\, \;dW_t\;,
\end{equation}
where
\begin{itemize}
\item the notation $[A,B]$ refers to $AB-BA;$
\item $H=H^\dag$ is a Hermitian operator which describes the action of external forces on the system\,;
\iffalse
\item $L$ is an arbitrary matrix which determines the coupling to the external field\,;
\fi
\item $dW_t$ is the Wiener process which is the following innovation
\begin{equation}\label{eq:observation}
dW_t=dy_t-\,\tr{(L+L^\dag)\,\rho_t}\,dt\;,
\end{equation}
where $y_t$ is a continuous semi-martingale with quadratic variation $\langle y,y\rangle_t=t$ (which is the observation process obtained from the system) and $L$ is an arbitrary matrix which determines the measurement process (typically the coupling to the probe field for quantum optic systems)\;;
\item the super-operator $\LL$ is the Lindblad operator,
\begin{equation*}
\LL(\rho):=-\half\{L^\dag L,\rho\}+L \rho L^\dag,
\end{equation*}
where the notation $\{A,B\}$ refers to $AB+BA;$
\item the super-operator $\Lambda$ is defined by
\begin{equation*}
\Lambda(\rho):= L\rho+\rho L^\dag-\tr{(L+L^\dag)\rho}\rho.
\end{equation*}
\end{itemize}
All the developments remain valid when $H$ and $L$ are deterministic time-varying matrices. For clarity sake, we do not recall below such possible time dependence.
\medskip
The evolution of the quantum filter of state $\widehat\rho_t\in{\mathcal D}$ is described by the following stochastic master equation which depends on the time-continuous measurement $y_t$ depending on the true quantum state~$\rho_t$ via~\eqref{eq:observation} (see, e.g.,~\cite{barchielli1990direct}):
\begin{align}\label{eq:master2}
d\widehat{\rho}_t=-\tfrac{i}{\hbar}[H,\widehat{\rho}_t]dt&+\LL(\widehat{\rho}_t)\, dt
+\Lambda(\widehat{\rho}_t)\big(dy_t-\,\tr{(L+L^\dag)\widehat{\rho_t}}dt\big).
\end{align}
Replacing $dy_t$ by its value given in~\eqref{eq:observation}, we obtain
\begin{align*}
d\widehat{\rho}_t=&-\tfrac{i}{\hbar}[H,\widehat{\rho}_t]dt+ \LL(\widehat{\rho}_t)dt+\,\Lambda(\widehat{\rho}_t) \,dW_t\nonumber\\
&+\Lambda(\widehat{\rho}_t)\Big(\tr{(L+L^\dag)\rho_t}-\tr{(L+L^\dag)\widehat{\rho}_t}\Big)dt.
\end{align*}
A usual measurement of the difference between two quantum states $\rho$ and $\sigma$, is given by the fidelity, a real number between $0$ and $1$. More precisely, the fidelity between $\rho$ and $\sigma$ in $\mathcal D$ is given by (see~\cite[chapter 9]{nielsen1999quantum} for more details)
\begin{equation}\label{eq:fid}
F(\rho,\sigma)=\trr{\sqrt{\sqrt{\rho}\sigma\sqrt{\rho}}}.
\end{equation}
Here $F(\rho,\sigma)=1$ means $\rho=\sigma$, and $F(\rho,\sigma)=0$ means that the support of $\rho$ and $\sigma$ are orthogonal. $F(\rho,\sigma)$ coincides with their inner product $\tr{\rho\sigma}$ when at least one of the states $\rho$ or $\sigma$ is pure (i.e., orthogonal projector of rank one). It is well known that the stochastic master equations~\eqref{eq:master1} and~\eqref{eq:master2} leave the domain $\mathcal D$ positively invariant.
This results form the fact that, using Ito rules, we have
\begin{multline}\label{eq:rhoJump}
\rho_{t+dt}=\tfrac{\left(\II-\tfrac{iH}{\hbar}dt-\half L^\dag L dt+Ldy_{t}\right)~\rho_t~\left(\II-\tfrac{iH}{\hbar}dt-\half L^\dag L dt+Ldy_{t}\right)^\dag}{\tr{\left(\II-\tfrac{iH}{\hbar}dt-\half L^\dag L dt+Ldy_{t}\right)~\rho_t~\left(\II-\tfrac{iH}{\hbar}dt-\half L^\dag L dt+Ldy_{t}\right)^\dag}}
\end{multline}
and
\begin{multline}\label{eq:rhohatJump}
\widehat{\rho}_{t+dt}=\tfrac{\left(\II-\tfrac{iH}{\hbar}dt-\half L^\dag L dt+Ldy_{t}\right)~\widehat{\rho}_t~\left(\II-\tfrac{iH}{\hbar}dt-\half L^\dag L dt+Ldy_{t}\right)^\dag}{\tr{\left(\II-\tfrac{iH}{\hbar}dt-\half L^\dag L dt+Ldy_{t}\right)~\widehat{\rho}_t~\left(\II-\tfrac{iH}{\hbar}dt-\half L^\dag L dt+Ldy_{t}\right)^\dag}}
\end{multline}
where $dy_{t}= \tr{(L+L^\dag)\,\rho_{t}}\,dt + dW_{t}$.
These alternative formulations imply then directly that, as soon as, $\rho_0$ and $\widehat\rho_0$ belong to $\mathcal D$, $\rho_t$ and $\widehat\rho_t$ remain in $\mathcal D$ for all $t \geq 0$. Therefore the expression of fidelity given by~\eqref{eq:fid} is well defined.
\medskip
\noindent We are now in position to state the main result of this paper.
\vspace{.3cm}
\begin{thm}\label{thm:main-tool} \rm Consider the Markov processes $(\rho_t,\widehat{\rho}_t)$ satisfying the stochastic master Equations~\eqref{eq:master1} and~\eqref{eq:master2} respectively with $\rho_0$, $\widehat\rho_0$ in $\mathcal D$. Then the fidelity $F(\rho_t,\widehat{\rho}_t)$, defined in Equation~\eqref{eq:fid}, is a submartingale, i.e.
$\EE{F(\rho_t,\widehat{\rho}_t)|(\rho_s,\widehat{\rho}_s)}\geq F(\rho_s,\widehat{\rho}_s),$ for all $t\geq s.$
\end{thm}
\medskip
We recall that the above theorem generalize the results of~\cite{diosi2006coupled} to arbitrary purity of the real states and quantum filter. If $\rho_0$ is pure, then $\rho_t$ remains pure for all $t>0$. In this case, $F(\rho_t,\widehat\rho_t)$ coincides with $\tr{\rho_t\widehat{\rho}_t}$. It is proved in ~\cite{diosi2006coupled} that this Frobenius inner product is a sub-martingale for any initial value of $\widehat{\rho}_t$:
$\frac{d}{dt}\,\EE{\tr{\rho_t\widehat{\rho}_t}}\geq 0.$ The main idea of the proof in~\cite{diosi2006coupled} consists in using It\^{o}'s formula to reduce the theorem to showing that $\EE{\tr{d\rho_t\widehat{\rho}_t+\rho_t d\widehat{\rho}_t+d\rho_t d\widehat{\rho}_t}}\geq 0$, and then using the shift invariance of the operator $L$ in the dynamics~\eqref{eq:master1} and~\eqref{eq:master2} and choosing an appropriate value.
In the absence of any information on the purity of the real states and the quantum filter, the fidelity is given by \eqref{eq:fid}, and the application of It\^{o}'s formula for the above expression becomes much more involved. In particular, the calculation of the cross derivatives was so complicated that it became hopeless to proceed this way. As the proof presented in the next section shows, we had to choose an undirect way to approach the theorem which allowed us to avoid the heavy calculations based on second order derivative of $F$.
\section{Proof of Theorem~\ref{thm:main-tool}}\label{sec:one}
We proceed in two steps.
\begin{itemize}
\item In the first step, we describe briefly how we obtain the stochastic master equations~\eqref{eq:master1} and~\eqref{eq:master2} as the limits of the stochastic master equations with Poisson processes using the diffusive limits inspired from the physical homodyne detection model~\cite{barchielli2009quantum,wiseman1993interpretation}.
\item In the second step, we show that the fidelity between the real state and the quantum filter which are the solutions of stochastic master equations with Poisson processes is a submartingale.
\end{itemize}
\medskip
\bf Step $1$. \rm Take $\alpha>0$ a large real number and consider the evolution of the quantum state $\rho^\alpha_t$ described by the following stochastic master equation derived from homodyne detection scheme (see section $6.4$ of~\cite{breuer2002theory} or~\cite{barchielli2009quantum},~\cite{wiseman1993interpretation}) for more physical details):
\begin{align}\label{eq:master3}
d\rho_t^{\alpha}=&-\tfrac{i}{\hbar}[H,\rho_t^{\alpha}]dt-\tfrac{1}{4}\Lambda_{\alpha}(\rho_t^{\alpha}) dt+\Upsilon_{\alpha}(\rho_t^{\alpha}) dN_1\\\nonumber
&-\tfrac{1}{4}\Lambda_{-\alpha}(\rho_t^{\alpha}) dt+\Upsilon_{-\alpha}(\rho_t^{\alpha}) dN_2\,,
\end{align}
where the super-operators $\Upsilon_\alpha$ is defined as follows
\begin{equation*}
\Upsilon_\alpha(\rho):= \frac{(L+\alpha)\rho (L^\dag+\alpha )}{\tr{(L+\alpha)\rho (L^\dag+\alpha )}}-\rho,
\end{equation*}
and the super-operator $\Lambda_\alpha$ is defined by
\begin{equation*}
\Lambda_\alpha(\rho):=(L+\alpha)\rho+\rho (L^\dag+\alpha )-\tr{(L+L^\dag+2\alpha) \rho}\rho.
\end{equation*}
The super-operators $\Lambda_{-\alpha}$ and $\Upsilon_{-\alpha}$ are just obtained with replacing $\alpha$ by $-\alpha$ in the expressions given in above.
The two processes $dN_1$ and $dN_2$ are defined by
$$dN_1:=N_{t+dt}^1-N_t^1 \quad \textrm{and} \quad dN_2:=N_{t+dt}^2-N_t^2 $$
where $N^1$ and $N^2$ are two Poisson processes. \noindent $dN_1$ and $dN_2$ take value $1$ by probabilities {\small$\half\tr{(L^\dag+\alpha )(L+\alpha)\rho_t^\alpha}dt$} and {\small$\half\tr{(L^\dag-\alpha )(L-\alpha)\rho_t^\alpha}dt$}, respectively, and take value $0$ by the complementary probabilities.
\medskip
Similarly, the following stochastic master equation describes the infinitesimal evolution of associated quantum filter of state $\widehat\rho^\alpha_t$ (see~\cite{barchielli1990direct}):
\begin{align}\label{eq:master4}
d\widehat{\rho_t}^{\alpha}=&-\tfrac{i}{\hbar}[H,\widehat{\rho_t}^{\alpha}]dt-\tfrac{1}{4}\Lambda_\alpha(\widehat{\rho_t}^{\alpha}) dt+\Upsilon_\alpha(\widehat{\rho_t}^{\alpha}) dN_1\\\nonumber
&-\tfrac{1}{4}\Lambda_{-\alpha}(\widehat{\rho_t}^{\alpha}) dt+\Upsilon_{-\alpha}(\widehat{\rho_t}^{\alpha}) dN_2.
\end{align}
The following diffusive limit is obtained by the central limit theorem when $\alpha$ tends to $+\infty$ for the semi-martingale processes applied to $dN_q$, $q=1,2$, (see ~\cite{lane1984central} or~\cite{jacod1987limit} for more details)
\begin{equation}\label{eq:limit}
dN_q\stackrel{\text{\tiny law}}{\longrightarrow} \langle\tfrac{dN_q}{dt}\rangle \,dt+\sqrt{\langle\tfrac{dN_q}{dt}\rangle}\,dW_q\,,
\end{equation}
\normalsize
where the notation $\langle A\rangle$ refers to the mean value of A. Here \\
$\langle dN_1\rangle=\half\tr{(L^\dag+\alpha )(L+\alpha)\rho_t^\alpha}\,dt$ and $\langle dN_2\rangle=\half\tr{(L^\dag-\alpha )(L-\alpha)\rho_t^\alpha}\,dt$ and $dW_1$ and $dW_2$ are two independent Wiener processes and the convergence in~\eqref{eq:limit} is in law.
\medskip
The stochastic master Equations~\eqref{eq:master1} and~\eqref{eq:master2} are obtained by replacing the processes $dN_q$ for $q\in\{1,2\}$ by their limits given in~\eqref{eq:limit} in the master equations~\eqref{eq:master3} and~\eqref{eq:master4} and taking the limit when $\alpha$ goes to $+\infty$ and keeping only the lowest ordered terms in $\alpha^{-1}.$ Such a result is usually called diffusion approximation (see e.g~\cite{costantini1991diffusion}).
\medskip
Notice that {$dW$} appearing in the stochastic master equations~\eqref{eq:master1} and~\eqref{eq:master2} is given in terms of its independent constituents by
\begin{equation*}
dW=\sqrt{\half}\,\big(dW_1+dW_2\big),
\end{equation*}
and is thus itself a standard Wiener process.
The following theorem from~\cite{pellegrini2009diffusion} justifies the diffusion approximation described above.
\medskip
\begin{thm}[Pellegrini-Petruccione~\cite{pellegrini2009diffusion}]\label{lem:loi}
\rm The solutions of the stochastic master Equations~\eqref{eq:master3} and~\eqref{eq:master4} converge in law, when $\alpha\rightarrow+\infty$, to the solutions of the stochastic master Equations~\eqref{eq:master1} and~\eqref{eq:master2}, respectively.
\end{thm}
\medskip
\bf Step $2$. \rm We now prove that the fidelity between two arbitrary solutions of the stochastic master Equations~\eqref{eq:master3} and~\eqref{eq:master4} is a submartingale.
\medskip
\begin{prop}\label{prop:second}
Consider the Markov process $(\rho^{\alpha},\widehat{\rho}^{\alpha})$ which satisfy the stochastic master Equations~\eqref{eq:master3} and~\eqref{eq:master4}. Then the
fidelity defined in Equation~\eqref{eq:fid} is a
submartingale, i.e.,
\rm for all $t\geq s$, we have
\begin{equation*}
\EE{F(\rho_t^{\alpha},\widehat{\rho}_t^{\alpha})|(\rho_s^{\alpha},\widehat{\rho}_s^{\alpha})}\geq F(\rho_s^{\alpha},\widehat{\rho}_s^{\alpha}).
\end{equation*}
\end{prop}
\medskip
\begin{proof}
We consider approximations of the time-continuous Markov processes~\eqref{eq:master3} and~\eqref{eq:master4} by discrete-time Markov processes $\xi_k$ and $\widehat{\xi_k}$:
\begin{equation}\label{eq:markov}
\xi_{k+1}=\tfrac{M_{\mu_k}\xi_k M_{\mu_k}^\dag}{\tr{M_{\mu_k}\xi_k M_{\mu_k}^\dag}}\quad\text{and}\quad \widehat{\xi}_{k+1}=\tfrac{M_{\mu_k}\widehat{\xi}_k M_{\mu_k}^\dag}{\tr{M_{\mu_k}\widehat{\xi}_k M_{\mu_k}^\dag}},
\end{equation}
where
\begin{itemize}
\item $k\in\{0,\cdots,n\}$ for a fixed large $n;$
\item initial condition $\xi_0=\rho_s^\alpha$ and $\widehat\xi_0=\widehat\rho_s^\alpha$;
\item $\mu_k$ is a random variable taking values $\mu\in\{0,1,2\}$ with probability $P_{\mu,k}=\tr{M_{\mu}\xi_k M_{\mu}^\dag};$
\item The operators $M_0,$ $M_1$ and $M_2$ are defined as follows
\begin{align*}
M_0:=1&-\tfrac{1}{4}(L^\dag+\alpha )(L+\alpha)\epsilon_n
-\tfrac{1}{4}(L^\dag-\alpha )(L-\alpha)\epsilon_n-\tfrac{i}{\hbar}H\epsilon_n;
\end{align*}
\begin{equation*}
M_1:=(L+\alpha)\sqrt{\half\epsilon_n};
\end{equation*}
and
\begin{equation*}
M_2:=(L-\alpha)\sqrt{\half\epsilon_n};
\end{equation*}
\end{itemize}
with $\epsilon_n=\tfrac{t-s}{n}.$
\medskip
In the following lemma, we show that $\xi_n$ and $\widehat{\xi}_n$ correspond to the Euler-Maruyama time discretization. Since~\eqref{eq:master3} and~\eqref{eq:master4} depend smoothly on $\rho^\alpha_t$ and $\widehat\rho^\alpha_t$, $\xi_n$ and $\widehat{\xi}_n$ converge in law towards $\rho^\alpha_t$ and $\widehat\rho^\alpha_t$ when $n\mapsto +\infty$.
\medskip
\begin{lem}\label{lem:third}
The processes $\xi_k$ and $\widehat{\xi}_k$ correspond up to second order terms in $\epsilon_n$, to the Euler-Maruyama discretization scheme of~\eqref{eq:master3} and~\eqref{eq:master4} on $[s,t]$.
\end{lem}
\iffalse
\begin{lem}\label{lem:third}
\rm The process $\rho_s^\alpha$ and $\widehat{\rho}_s^\alpha$ are given by the following point-wise convergence $\rho_s^\alpha=\lim_{N\longrightarrow\infty}\xi_N$ and
$\widehat{\rho}_s^\alpha=\lim_{N\longrightarrow\infty}\widehat{\xi}_N$.
\end{lem}
\fi
\medskip
\begin{proof} we regard the three following possible cases which arrive in according to the different values of $\mu_k$ . In each case, we show that $\xi_k$ and $\widehat{\xi}_k$ for $k\in\{0,\cdots,n\}$ are the numerical solutions of the dynamics ~\eqref{eq:master3} and~\eqref{eq:master4} respectively, with the following partition $s\leq s+\epsilon_n\leq\cdots\leq s+(n-1)\epsilon_n\leq t,$ where the uniform step length $\epsilon_n$ is $\tfrac{t-s}{n}.$
\medskip
\bf Case 1. \rm We first consider the case where $\mu_k=0$ which arrives with probability $P_{0,k}=\tr{M_0\xi_k M_0^\dag}.$ Note that
\begin{align*}
M_0\xi_k M_0^\dag=& \,\xi_k-\tfrac{1}{4}\{(L^\dag+\alpha )(L+\alpha),\xi_k\}\,\epsilon_n\\
&-\tfrac{1}{4}\{(L^\dag-\alpha )(L-\alpha),\xi_k\}\,\epsilon_n\\
&-\tfrac{i}{\hbar}[H,\xi_k]\, \epsilon_n+\OO(\epsilon_n^2).
\end{align*}
Therefore
\begin{align*}
\tr{M_0\xi_k M_0^\dag}&=1-\half\tr{(L^\dag+\alpha )(L+\alpha)\xi_k}\,\epsilon_n\\
&-\half\tr{(L^\dag-\alpha )(L-\alpha)\xi_k}\,\epsilon_n+\OO((\epsilon_n)^2)
\end{align*}
and
\begin{align*}
\big(\tr{M_0\xi_k M_0^\dag}\big)^{-1}&\approx 1+\half\tr{(L^\dag+\alpha )(L+\alpha)\xi_k}\,\epsilon_n+\\
&\half\tr{(L^\dag-\alpha )(L-\alpha)\xi_k}\,\epsilon_n+\OO((\epsilon_n)^2).
\end{align*}
Therefore, we find the following dynamics
\begin{align*}
\xi_{k+1}&\approx\xi_k-\tfrac{1}{4}\{(L^\dag+\alpha )(L+\alpha),\xi_k\}\,\epsilon_n\\
&-\tfrac{1}{4}\{(L^\dag-\alpha )(L-\alpha),\xi_k\}\,\epsilon_n\\
&+\half\tr{(L^\dag+\alpha )(L+\alpha)\xi_k}\,\xi_k\,\epsilon_n\\
&+\half\tr{(L^\dag-\alpha )(L-\alpha)\xi_k}\,\xi_k\,\epsilon_n+\OO(\epsilon_n^2).
\end{align*}
This can also be written as follows
\begin{align}\label{eq:m1}
\xi_{k+1}-\xi_k\approx&-\tfrac{1}{4}\Lambda_\alpha(\xi_k)\,\epsilon_n-\tfrac{1}{4}\Lambda_{-\alpha}(\xi_k)\,\epsilon_n+\OO(\epsilon_n^2).
\end{align}
Obviously, this dynamics in the first order of $\epsilon_n$ is equivalent to the dynamics of the numerical solution of the stochastic master Equation~\eqref{eq:master3} with the partition $s\leq s+\epsilon_n\leq\cdots\leq s+(n-1)\epsilon_n\leq t,$ when
\begin{equation*}\label{eq:case1}
N_{s+(k+1)\epsilon_n}^1-N_{s+k\epsilon_n}^1=0\quad \textrm{and}\quad N_{s+(k+1)\epsilon_n}^2-N_{s+k\epsilon_n}^2=0,
\end{equation*}
which arrives with probability
\begin{align*}
\big(1-\half\tr{(L+\alpha)(L^\dag+\alpha )\,\xi_k}\,\epsilon_n\big)\cdots\\
\cdots\big(1-\half\tr{(L-\alpha)(L^\dag-\alpha )\,\xi_k}\,\epsilon_n\big).
\end{align*}
This probability, in the first order of $\epsilon_n$ is equal to $\tr{M_0\xi_k M_0^\dag}.$
\medskip
\bf Case 2. \rm The second case corresponds to $\mu_k=1$ which arrives with probability $\tr{M_1\xi_k M_1^\dag}.$ We find the following dynamics
\begin{equation*}
\xi_{k+1}=\tfrac{(L+\alpha)\xi_k (L^\dag+\alpha )}{\tr{(L+\alpha)\xi_k (L^\dag+\alpha )}}=\Upsilon[L+\alpha]\,\xi_k+\xi_k.
\end{equation*}
We observe that the numerical solution of the stochastic master Equation~\eqref{eq:master3} follows also the same dynamics when
$$N_{s+(k+1)\epsilon_n}^1-N_{s+k\epsilon_n}^1=1\quad\textrm{and}\quad N_{s+(k+1)\epsilon_n}^2-N_{s+k\epsilon_n}^2=0,$$ which arrives with probability
\begin{align*}
\big(\half\tr{(L+\alpha)(L^\dag+\alpha )\,\xi_k}\,\epsilon_n\big)\big(1-\half\tr{(L-\alpha)(L^\dag-\alpha )\,\xi_k}\,\epsilon_n\big).
\end{align*}
This is equal to $\tr{M_1\xi_k M_1^\dag},$ in the first order of $\epsilon_n.$
\medskip
\bf Case 3. \rm Now we consider the last case $\mu_k=2$ which arrives with probability $\tr{M_2\xi_k M_2^\dag}.$ Therefore, we have
\begin{equation*}
\xi_{k+1}=\tfrac{(L-\alpha)\xi_k (L^\dag-\alpha )}{\tr{(L-\alpha)\xi_k (L^\dag-\alpha )}}=\Upsilon_{-\alpha}(\xi_k)+\xi_k.
\end{equation*}
Which can also be written by the stochastic master equation~\eqref{eq:master3} with taking $\xi_k$ as the numerical solution and
$$N_{s+(k+1)\epsilon_n}^1-N_{s+k\epsilon_n}^1=0 \quad\textrm{and}\quad N_{s+(k+1)\epsilon_n}^2-N_{s+k\epsilon_n}^2=1,$$ which arrives with probability
\begin{align*}
\big(1-\half\tr{(L+\alpha)(L^\dag+\alpha )\,\xi_k}\,\epsilon_n\big)\big(\half\tr{(L-\alpha)(L^\dag-\alpha )\,\xi_k}\,\epsilon_N\big).
\end{align*}
Where in the first order of $\epsilon_n,$ this probability is equal to $\tr{M_2\xi_k M_2^\dag}.$
\medskip
Remark that, if we neglect the terms in the order of $\epsilon_n^2,$ The probability of $N_{s+(k+1)\epsilon_n}^1-N_{s+k\epsilon_n}^1=1$ and $N_{s+(k+1)\epsilon_n}^2-N_{s+k\epsilon_n}^2=1$ is negligible. Now it is clear that $\xi_k$ and similarly $\widehat{\xi}_k$ are respectively the numerical solutions of the stochastic master Equations~\eqref{eq:master3} and~\eqref{eq:master4} obtained by Euler-Maruyama method. As the right hand side of the stochastic master Equations~\eqref{eq:master3} and~\eqref{eq:master4} are smooth with respect to $\rho$ and $\widehat{\rho}$, we can use the result of~\cite[Theorem 1]{higham2005numerical} to conclude the convergence in law of $\xi_n$ and $\widehat{\xi}_n$ to $\rho^\alpha_t$ and $\widehat\rho^\alpha_t$ for large $n$.
\end{proof}
Now we notice that
\begin{equation*}
M_0^\dag M_0+M_1^\dag M_1+M_2^\dag M_2=\II+\OO(\epsilon_n^2):=A,
\end{equation*}
Take $\widetilde{M_r}:=(\sqrt{A})^{-1}M_r$ for $r=0,1,2$ which satisfy necessarily
\begin{equation}\label{eq:qnd}
\widetilde{M_0}^\dag\widetilde{M_0}+\widetilde{M_1}^\dag\widetilde{M_1}+\widetilde{M_2}^\dag\widetilde{M_2}=\II.
\end{equation}
Now we define the following Markov processes $\chi_k$ and $\widehat{\chi}_k$ by
\begin{equation}\label{eq:dyn1}
\chi_{k+1}=\tfrac{\widetilde{M_{\mu_k}}\chi_k\widetilde{M_{\mu_k}}^\dag}{\tr{\widetilde{M_{\mu_k}}\chi_k\widetilde{M_{\mu_k}}^\dag}}
\end{equation}
and
\begin{equation}\label{eq:dyn2}
\widehat{\chi}_{k+1}=\tfrac{\widetilde{M_{\mu_k}}\widehat{\chi}_k\widetilde{M_{\mu_k}}^\dag}{\tr{\widetilde{M_{\mu_k}}\widehat{\chi}_k\widetilde{M_{\mu_k}}^\dag}}\,,
\end{equation}
where
\begin{itemize}
\item $k\in\{0,\cdots,n\}$ for a fixed large $n;$
\item $\chi_0=\rho_s^\alpha$ and $\widehat\chi_0=\widehat\rho_s^\alpha$;
\item $\mu_k$ is a random variable taking values $\mu\in\{0,1,2\}$ with probability $P_{\mu,k}=\tr{\widetilde{M}_{\mu}\chi_k \widetilde{M}_{\mu}^\dag}.$
\end{itemize}
Clearly $\chi_k$ and $\widehat{\chi}_k$ can also be seen as the numerical solutions of the stochastic master Equations~\eqref{eq:master3} and~\eqref{eq:master4}, since $(\sqrt{A})^{-1}=\II-\OO(\epsilon_n^2),$ therefore in the first order of $\epsilon_n,$ the solutions $\xi_k$ and $\widehat{\xi}_k$ are equal to $\chi_k$ and $\widehat{\chi}_k,$ respectively. But, the advantage of using $\chi_k$ and $\widehat{\chi}_k$ instead of $\xi_k$ and $\widehat{\xi}_k$ is that the operators $\widetilde{M_r}$ are Kraus operators since they satisfy Equality~\eqref{eq:qnd}. Thus we can apply Theorem~$1$ in~\cite{rouchon2010fidelity}, which proves that $F(\chi_k,\widehat{\chi}_k)$ is a sub-martingale.
\medskip
\begin{thm}[\cite{rouchon2010fidelity}]\label{thm:discret}
\rm Consider the Markov chain $(\chi_k,\widehat{\chi}_k)$ satisfying~\eqref{eq:dyn1} and~\eqref{eq:dyn2}. Then $F(\chi_k,\widehat{\chi}_k)$ is a sub-martingale:
$
\EE{F(\chi_{k+1},\widehat{\chi}_{k+1})|(\chi_k,\widehat{\chi}_k)}\geq F(\chi_k,\widehat{\chi}_k).
$
\end{thm}
\medskip
Thus we have
$$
\EE{F(\chi_{n},\widehat{\chi}_{n})~|~ \chi_0,\widehat{\chi}_0 }\geq {F(\chi_0,\widehat{\chi}_0)}=F(\rho^\alpha_s,\widehat{\rho}^\alpha_s)
$$
Therefore by Lemma~\ref{lem:third}, we have necessarily
\begin{equation*}
\EE{F(\rho_t^{\alpha},\widehat{\rho}_t^\alpha)|\rho_s^{\alpha},\widehat{\rho}_s^\alpha)}\geq F(\rho_s^\alpha,\widehat{\rho}_s^{\alpha}),
\end{equation*}
for all $t\geq s,$ since we have (convergence in law) $\rho_t^\alpha=\lim_{n\longrightarrow\infty}\chi_n,$ $\widehat{\rho}_t^\alpha=\lim_{n\longrightarrow\infty}\widehat{\chi}_n,$ $\chi_0=\rho_s^\alpha$ and $\widehat{\chi}_0=\widehat{\rho}_s^\alpha.$
\end{proof}
\medskip
We now apply Theorem~\ref{lem:loi} and we use the fact that the function $F$ is bounded by one and continuous with respect to $\rho$ and $\widehat{\rho}$:
\begin{equation*}
\EE{F(\rho_t,\widehat{\rho}_t)|(\rho_s,\widehat{\rho}_s)}\geq F(\rho_s,\widehat{\rho}_s),
\end{equation*}
for all $t\geq s,$ which ends the proof of Theorem~\ref{thm:main-tool}.
\section{Numerical Test}
In this section, we test the result of Theorem~\ref{thm:main-tool} through numerical simulations. Considering the two-level system of~\cite{vanHandel-et-al-2005}, we take the following Hamiltonian and measurement operators:
\begin{equation*}
H=\sigma_y=\begin{pmatrix}
0&-i \\
i&0\end{pmatrix}\qquad L=\sigma_z=\begin{pmatrix}
1&0 \\
0&-1\end{pmatrix}.
\end{equation*}
The simulations of figure~\ref{fig:fidelity} illustrates the fidelity for 500 random trajectories starting at
\begin{equation*}
\rho_0=\begin{pmatrix}
\tfrac{1}{2}&\tfrac{1}{4} \\
\tfrac{1}{4}&\tfrac{1}{2}\end{pmatrix} \quad\textrm{and}\quad \widehat{\rho_0}=\begin{pmatrix}
\tfrac{1}{3}&0 \\
0&\tfrac{2}{3}\end{pmatrix}.
\end{equation*}
In particular, we note that both initial states are mixed ones. As it can be seen the average fidelity is monotonically increasing. Here, the fidelity converges to one indicating the convergence of the filter towards the physical state. An interesting direction here is to characterize the situations where this convergence is ensured.
Here in order to simulate the Equations~\eqref{eq:master1} and~\eqref{eq:master2}, we have considered the alternative formulations~\eqref{eq:rhoJump} and~\eqref{eq:rhohatJump} and the resulting discretization scheme ($k\in\NN$ and time step $0<dt\ll 1$)
$$
\rho_{(k+1)dt}=\tfrac{\mathcal M_k\rho_{(kdt)} \mathcal M_k^\dag}{\tr{\mathcal M_k\rho_{(kdt)} \mathcal M_k^\dag}}, \quad
\widehat{\rho}_{(k+1)dt}=\tfrac{{\mathcal M}_k\widehat{\rho}_{(kdt)}{\mathcal M}_k^\dag}{\tr{{\mathcal M}_k\widehat{\rho}_{(kdt)}{\mathcal M}_k^\dag}},
$$
where $\mathcal M_k =\II-\tfrac{iH}{\hbar}dt-\half L^\dag L dt+Ldy_{(k dt)}$ and
$ dy_{(kdt)}= \tr{(L+L^\dag)\,\rho_{(kdt)}}\,dt + dW_{(kdt)}$. For each $k$, the Wiener increment $dW_{(kdt)}$ is a centered Gaussian random variable of standard deviation $\sqrt{dt}$. The major interest of such discretization is to guaranty that, if $\rho_0,\widehat\rho_0\in\mathcal D,$ then $\rho_k$ and $\widehat\rho_k$ also remain in $\mathcal D$ for any $k\geq 0$.
\begin{figure}
\centerline{\includegraphics[width=0.7\textwidth]{fidelity}}
\caption{The average fidelity between the Markov processes $\rho$ and $\widehat{\rho},$ over $500$ realizations,
time $t$ from $0$ to $T=3$ with discretization time step $dt=10^{-4}$. }\label{fig:fidelity}
\end{figure}
\section{Concluding remarks}\label{sec:third}
The fact that the fidelity between the real quantum state and the quantum-filter state increases in average remains valid for more general stochastic master equations where other Lindbald terms are added to $\LL(\rho)$ appearing in~\eqref{eq:master1}. In this case the dynamics~\eqref{eq:master1} and~\eqref{eq:master2} become
\begin{multline*}
d\rho_t=-\tfrac{i}{\hbar}[H,\rho_t]\;dt+\sum_{\nu=1}^{m'}\LL'_\nu(\rho_t)\, dt
+\sum_{\mu=1}^m\LL_\mu(\rho_t)\, dt+\sum_{\mu=1}^{m}\Lambda_\mu(\rho_t)dW_t^\mu
\end{multline*}
\and
\begin{multline*}
d\widehat{\rho}_t=-\tfrac{i}{\hbar}[H,\widehat{\rho}_t]\,dt+\sum_{\nu=1}^{m'}\LL'_\nu(\widehat\rho_t)\, dt
+ \sum_{\mu=1}^m\LL_\mu(\widehat{\rho}_t)\, dt\\+\sum_{\mu=1}^m\Lambda_\mu(\widehat{\rho}_t)\bigg(dy_t^\mu-\,\tr{(L_\mu+L^\dag_\mu)\widehat{\rho_t}}dt\bigg).
\end{multline*}
where $dW_t^\mu$ are independent Wiener processes,
\begin{equation*}
\LL_\mu(\rho):=-\half\{L_\mu^\dag L_\mu,\rho\}+L_\mu \rho L_\mu^\dag,
\end{equation*}
\begin{equation*}
\LL'_\nu(\rho):=-\half\{{L'_\nu}^\dag L'_\nu,\rho\}+L'_\nu \rho {L'_\nu}^\dag,
\end{equation*}
and $\Lambda_\mu(\rho):= L_\mu\rho+\rho L_\mu^\dag-\tr{(L_\mu+L_\mu^\dag)\rho}\rho.$
Here $m,m^\prime \geq 1$, and $(L^\prime_\nu)_{1\leq \nu\leq m^\prime}$ and $(L_\mu)_{1\leq \mu\leq m}$ are arbitrary operators. The special case considered here corresponds to $m=1$ and $m'=1$ with $L_1=L$ and $L'_1=0.$
The formulations analogue to~\eqref{eq:rhoJump} and~\eqref{eq:rhohatJump} read then
$$
\rho_{t+dt} = \frac{(\II-dM_t) \rho_t (\II-dM_t^\dag) + \sum_{\nu=1}^{m'} L'_\nu \rho_t {L'_\nu}^\dag dt }{\tr{(\II-dM_t) \rho_t (\II-dM_t^\dag) + \sum_{\nu=1}^{m'} L'_\nu \rho_t {L'_\nu}^\dag dt}}
$$
and
$$
\widehat\rho_{t+dt} = \frac{(\II-dM_t)\widehat\rho_t (\II-dM_t^\dag) + \sum_{\nu=1}^{m'} L'_\nu \widehat\rho_t {L'_\nu}^\dag dt }{\tr{(\II-dM_t) \widehat\rho_t (\II-dM_t^\dag) + \sum_{\nu=1}^{m'} L'_\nu \widehat\rho_t {L'_\nu}^\dag dt}}
$$
where, denoting $dy_t^\mu = \tr{(L_\mu+L_\mu^\dag)\rho_t} dt + dW_t^\mu$,
$$
dM_t=\tfrac{iH}{\hbar}dt+\half\sum_{\nu=1}^{m'} {L'_\nu}^\dag L'_\nu dt
+
\half\sum_{\mu=1}^{m} {L_\mu}^\dag L_\mu dt - \sum_{\mu=1}^{m} L_\mu dy^\mu_{t}
.
$$
For this general case, the proof of Theorem~\ref{thm:main-tool} should follow the same lines: first step still relies on Theorem~\ref{lem:loi}; second step relies now on~\cite[Theorem $2$]{rouchon2010fidelity}.
\bf Acknowledgements. \rm The authors thank C. Pellegrini, L. Zambotti and M. Gubinelli for stimulating discussions and useful suggestions during the preparation of the paper.
|
\section{Introduction}
\indent
As velocity and momentum in the Dirac theory~\cite{dirac2-1928}\
are related in a different manner than in non--relativistic
mechanics, it leads~\cite{Dirac58}--\cite{Merzbacher98} to the
fact that the average speed of spin--${1/2}$ particles must be
less than $c$, while the instantaneous speed is always $\pm c$ in
this theory. Indeed, by using an arbitrary state of a
spin--${1/2}$ free particle in which it is localized in space, one
can show~\cite{huang1952} that the expectation value of the
position, even in the absence of external field, includes a
time--dependent part which originates from an interference between
the positive and negative energy states. Hence, the motion of free
particle has a peculiar oscillatory component with a very high
frequency of $2mc^2/\hbar\simeq 1.6\times {10}^{21} {\rm s}^{-1}$\
for electrons with the rest mass m -- which secures $\pm c$ as the
eigenvalues of instantaneous speed. That is, a free particle obeys
a rapid oscillatory motion with the speed $c$ around the center of
mass while moving like a relativistic particle with velocity ${\bf
p}/m$. This angular frequency is referred to $\omega_{_{\rm
zbw}}\equiv 2mc^2/\hbar$ after Schr\"odinger who called this
motion \emph{Zitterbewegung}\ ({\bf
Zbw})~\cite{schrodinger30-31,barut-etal81a} -- trembling or
quivering motion -- which he theoretically contrived it after
Dirac had employed his equation for free relativistic electron in
vacuum. He interpreted this as a fluctuation in the position of
electron. The amplitude of the Zbw has also been indicated to be
of the order of the Compton wavelength,
$\lambdabar_c=\hbar/(mc)\simeq 3.86\times {10}^{-13} {\rm m}$ for
electrons. Now, the well--known phenomenon of the Zbw, caused by
the interference between the positive and negative energy
components of a wave packet, has been described in many textbooks
as well~\cite{Dirac58}--\cite{Merzbacher98}.
It is also well--known~\cite{Dirac58}--\cite{Merzbacher98} that as
the position operator in the Dirac theory is not a one--particle
observable, the path of an electron can only be indicated up to
the order of its Compton wavelength. Indeed, the best localized
wave packet exclusively consisting of plane wave components with
positive or negative energy has characteristic linear dimension of
about $\lambdabar_c$~\cite{newton-etal49}. However, the Zbw for
free electrons is not an observable motion and has never been
directly detected~\cite{rusin-etal2006} (nevertheless, see
discussion in Sec.~$3$). Actually, with certain initial
conditions, one can show~\cite{huang1952} that all components of
$<\!\mbox{\boldmath$r$}\!>$ vanish upon integration, which
indicates that the total contribution of all waves vanishes (see
Sec.~$2$). Hence, in these conditions, the center of the wave
packet as a whole remains at rest, as experimental evidence rules
out the possibility that electron is an extended
body\rlap.\footnote{Indeed, scattering experiments have revealed
that the size of electron is less than $10^{-18} {\rm
m}$~\cite{bender84}.}\
Though, the consequence of the Zbw is that it is
impossible to localize electron better than to a certain finite
volume.
The above argument leaves open the possibilities that electron may
be regarded as a point charge which generates spin and magnetic
moment by an intrinsic local motion~\cite{hestenes93}. Actually,
in the literature, people have tried to analyze some picture about
the detailed motion of the Zbw and to interpret the structural
feature of electron. The motivation has mainly come from the close
connection between the Zbw and the intrinsic magnetic moment of
electron. That is, using the current density expression of a free
electron, one can
infer~\cite{gordon27frenkel341,gordon27frenkel342} that the
classical expression for the magnetic moment has its root in the
Zbw (see also Sec.~$2$). Hence, a temptation that the magnetic
moment and the spin of electron are generated by some sort of the
local circular motion of electron, i.e. mass and charge,
\cite{huang1952,hestenes93,barut-etal81b,hestenes90,hestenes2009}.
Indeed, many classical treatments of a free Dirac electron have
been formed based on this picture for the Zbw, see
Refs.~\cite{MHWB1}--\cite{SalesiRecami2000} and references
therein. A significant achievement toward understanding the Zbw
has been made~\cite{PFetalT1}--\cite{PFetalT3} when it was shown
that in the absence of external fields the Dirac Hamiltonian can
be transformed into a form in which the positive and negative
electron energies are decoupled. In this new representation,
electron is not a point--like particle and has the size of about
$\lambdabar_c$. The interpretation of these two pictures has been
considered in
Refs.~\cite{huang1952,barut-etal81a,newton-etal49},~\cite{FetalBetalL1}--\cite{Krekora-etal2004},
and the compatibility of the Zbw interpretation with the details
of the Dirac theory has also been
demonstrated~\cite{hestenes90,hestenes2009,BarutZanghi84}.
Huang~\cite{huang1952} obtained this picture from the Dirac
equation and claimed that it is not an unreasonable one. Actually,
using an explicit form for the wave packet with specified initial
condition to construct the wave packet mainly from positive or
negative waves, he showed that the contribution of each plane wave
to the motion of the wave packet is an orbit whose projection on
the plane perpendicular to the direction of the spin is a circle
of radius $\lambdabar_c/2$. Then, using the method of the Gordon
decomposition~\cite{gordon27frenkel341,gordon27frenkel342}, he
also justified that the intrinsic spin of electron may be
considered as an ``orbital angular momentum'' due to the Zbw. That
is, the current produced by the Zbw actually causes the intrinsic
magnetic moment of the particle\rlap,\footnote{The same issue, but
in the presence of an external magnetic field, is studied at the
end of Sec.~$2$.}\
and hence, the total
magnetic moment of electron is represented by both the orbital and
the intrinsic angular momenta with the correct gyromagnetic $g$
factor.
A classical analogous model for spinning electron, which has been
claimed to be a natural classical limit of the Dirac
theory~\cite{SalesiRecami2000}, has been
presented~\cite{BarutZanghi84} by Barut and Zanghi ({\bf BZ}) in a
particular five dimensional space--time (which we refer to as BZ
approach). They introduced a classical Lagrangian (see
relation~(\ref{Lag})) in a manifestly covariant form using a pair
of spinorial variables (representing internal degrees of freedom)
besides the usual pair of dynamical variables $(x^{\mu},
p^{\mu})$, as functions of the proper time parameter measured in
the center of mass frame (see
Refs.~\cite{SalesiRecami2000,RecamiSalesi98} and references
therein). The general solution of the motion for free electrons
includes an oscillatory part as well, which denotes the presence
of the Zbw for the spinning particles (see Sec.~$3$). That is, the
Zbw occurs because the motion of particle does~not coincide with
the motion of the center of mass~\cite{Salesi2002}, and a spinning
particle appears as extended object~\cite{RecamiSalesi98}. It has
also been found that free polarized particles have circular
motions in the plane perpendicular to the direction of the spin,
with the Zbw frequency and half the Compton wavelength radius in
the center of mass frame.
After quantization, the BZ approach has also been indicated to
describe the Dirac electron~\cite{Barut-pavsic}. In fact, this has
been originated from the classical Hamiltonian corresponding to a
classical Lagrangian devised by BZ~\cite{BarutZanghi84} (here
shown with relation~(\ref{Lag})). Then, it has been converted to a
quantum version with reinterpretation of the spinorial variables
as Dirac field spinors~\cite{RecamiSalesi1995,Pavsic-elal1998}.
Also, by recasting the BZ approach into the language of Clifford
formalism\rlap,\footnote{In which they have also claimed that this
algebra allows various independent classical approaches of the
spin to be unified.}\
quantization of the theory has been
accomplished~\cite{Pavsic-elal1998}.
The significant correspondence between the Dirac theory and BZ
approach is intriguing and invites one to discover up to where
this compatibility lasts, and actually whether there is~not any
conflict between them. To pursue such a query, we purpose to study
the particle's Zbw in the presence of an external uniform static
magnetic field. Our stimulated hope in exploring this procedure is
that theoretical understanding of the Zbw may shed light on the
nature of spin of elementary particles.
In the standard textbooks~\cite{Dirac58}--\cite{Merzbacher98}, the
Zbw has only been examined in details for free spinning particles
without considering any interactions. Though, it has been
indicated that electron in the hydrogen atom, i.e. in the presence
of the Coulomb potential, also displays the Zbw which gives rise
to the so called Darwin term~\cite{Dirac58}--\cite{Merzbacher98}.
Interestingly enough, the phenomenon of Zbw for electrons has been
shown~\cite{rusin-etal2006},~\cite{ZCetalFetalZJetalZZSetalK1}--\cite{BrusheimXu2008}
to occur in non--relativistic case and in solids in the absence of
external fields. Even, an analogous quantum--relativistic effect
in a single trapped ion has been
studied~\cite{LamataLeonSchatzSolano}. In these conditions, it has
much lower angular frequency and much larger amplitude, and it may
lead easier to experimental detection (also see Sec.~$3$). Also,
in
Refs.~\cite{SalesiRecami2000,BarutZanghi84,Villavicencio-etal2000}
and intuitively in Ref.~\cite{holten92,holten93}, the phenomenon
of Zbw for electrons in the presence of an external magnetic field
has been established. Hence, the Zbw is~not a special
characteristic of plane wave solution, and does occur whenever
positive and negative energy states are
mixed~\cite{hestenes90,Villavicencio-etal2000}.
Actually, by interpreting the interference between positive and
negative energy states, the Zbw may be considered to arise in
electron--positron pair creation and
annihilation~\cite{thirring58} -- which manifests the Zbw as a
purely relativistic phenomenon. On the other hand, the Zbw is
considered to be the origin of the spin only in the
non--relativistic domain. By reformulating the Dirac theory, it
has been claimed~\cite{hestenes90,hestenes2009}\footnote{Also, see
Ref.~\cite{Pavsic-elal1998}.}\
that the Zbw is a
phenomenon which manifests itself in every application of quantum
mechanics, i.e. either in the non--relativistic domain and/or in
the presence of an arbitrary electromagnetic interaction. Indeed,
it has been claimed~\cite{hestenes93,hestenes90} that the most
features of quantum mechanics are manifestations of the Zbw, e.g.
the complex phase factor in the electron wave function as physical
representation of rotation can be associated directly with the
Zbw. However, we should mention the
assertion~\cite{Krekora-etal2004} that quantum field theory
prohibits the occurrence of the Zbw for an electron, and in
principle, there is no difficulty to localize an electron narrower
than the Compton wavelength if it is involved with a positron.
Regarding the question of an electron in vacuum in the presence of
an external magnetic field, the references~\cite{SalesiRecami2000}
and~\cite{Villavicencio-etal2000} are of much interest to us. This
is in fact the topic that we purpose to deal with by making
recourse to the Huang quantum approach~\cite{huang1952} where the
next section is devoted to. There, we contrast the results with
those two investigations, particularly with the more accurate
solutions that we will indicate by using
Ref.~\cite{SalesiRecami2000} in where the BZ classical approach
has been employed to study the issue. Actually, in Sec.~$2$, the
expectation value approach has been utilized to find changes of
$\omega_{_{\rm zbw}}$ in an external uniform static magnetic
field. In Sec.~$3$, we review the classical approach of
Ref.~\cite{SalesiRecami2000} and infer some improvements. It will
be shown that the results derived via the expectation value
quantum approach are different from the corresponding ones
obtained via the classical approach from
Ref.~\cite{SalesiRecami2000}. Finally, a brief discussion and
conclusion is given in the last section.
\section{Frequency Shift via Expectation Value Approach}
\indent
Huang~\cite{huang1952} used solutions to the Dirac equation and
the expectation value of velocity to study the issue of free
electron's Zbw. In this work, we evaluate the electron's Zbw in
the presence of an external uniform static magnetic field
utilizing his method. Thus, we first very briefly review some
results for the free particle and then, we will extend the method
to our case.
To examine the average position of an electron, which is localized
in space, a general Dirac wave packet, expanded in terms of
momentum eigenfunctions, as
\begin{equation}\label{Wav}
\!\!\!\!\!\Psi({\mbox{\boldmath$r$}},t)=h^{-3/2}\int{\Bigl[C_+({\mathbf{p}})\exp(-i\omega
t) +C_{-}({\mathbf{p}})\exp(i\omega
t)\Bigr]\exp\Bigl(i\frac{{\mathbf{p}}\cdot{\mbox{\boldmath$r$}}}{\hbar}\Bigr)
d^{3}p}
\end{equation}
can be used. This wave packet includes both negative and positive
energies, and $C_{+}({\mathbf{p}})$\
$\left(C_{-}({\mathbf{p}})\right)$ is a linear combination of the
spin--up and spin--down amplitudes of the free particle Dirac
waves with momentum $\textbf{p}$ and positive (negative) energy.
Using this wave packet and calculating
$<\!\dot{\mbox{\boldmath$r$}}\!>$, and consequently
$<\!\mbox{\boldmath$r$}\!>$ with a suitable initial condition
having a well--defined initial spin direction, it has been
shown~\cite{huang1952} that each plane wave in the above Fourier
decomposition contributes a circular motion in the plane
perpendicular to the direction of the electron spin with radius
$\lambdabar_c/2$ and frequency $\omega_{_{\rm zbw}}$. This leads
to the internal spin and hence, the magnetic moment of the Dirac
particle.
Now, the Dirac equation for an electron in an external
electromagnetic field is
\begin{equation}\label{Ham}
i\hbar\frac{\partial\Psi}{\partial
t}=\left[c\alpha_i(p^i-\frac{e}{c}A^i) +\beta mc^2+eA_0\right]\Psi
,
\end{equation}
where the matrices $\mbox{\boldmath$\alpha$}$ and $\beta$ are
defined as\footnote{In terms of the Dirac (gamma) matrices,
$\gamma_a$, they are $\beta\equiv\gamma^0$ and
$\alpha_i\equiv\gamma^0\gamma_i$ for $i=1, 2, 3$.}
\begin{eqnarray}\label{Pau}
\alpha_i=\left(
\begin{array}{cc}
0 & \sigma_i \\
\sigma_i & 0 \\
\end{array}
\right)\
\qquad\ \textrm{and}\qquad\ \beta=\left(
\begin{array}{cc}
I & 0 \\
0 & -I \\
\end{array}
\right),
\end{eqnarray}
with $\sigma_i$ and $I$ as the $2\times2$ Pauli matrices and the
unit matrix, respectively. We proceed with special case of an
external uniform static magnetic field without a scalar Coulomb
field (i.e. $A_0=0$). Then, using the standard abbreviation
$\pi_i\equiv p_i-eA_i/c$ in the Gaussian units -- i.e. defining
the gauge--invariant momentum operator correspondence
$\pi_\mu\rightarrow -i\hbar D_{_{\mu}}$ in terms of the covariant
derivative\footnote{In a general case, the rate of change of this
generalized operator, via the Heisenberg picture, gives the
Lorentz force, i.e. $\dot{\mbox{\boldmath$\pi$}}=i[H,
{\mbox{\boldmath$\pi$}}]/\hbar-(e/c)\partial {\mathbf{A}}/\partial
t=e({\bf E}+{\mbox{\boldmath$\alpha$}}\times{\bf B})$.}\
-- the Dirac equation (\ref{Ham}) reads $i\hbar\partial\Psi/\partial
t=\left(c\,\alpha _i\pi^i +\beta mc^2\right)\Psi$. Making a
general wave solution {\it ansatz\/}\footnote{We should emphasis
that due to the presence of an external magnetic field the wave
packet $\Psi$ is~not an eigenfunction of neither the operator
$\pi_i$ nor the operator $p_i$.}
\begin{equation}\label{Anz0}
\Psi(\mbox{\boldmath$r$}, t)=\left(
\begin{array}{c}
\phi(\mbox{\boldmath$r$}) \\
\chi(\mbox{\boldmath$r$})\\
\end{array}
\right)\exp -i\left(\frac{E_{_{\rm B}} t}{\hbar}\right),
\end{equation}
with $E_{_{\rm B}}$ as the particle energy in the presence of an
external magnetic field, hence the Dirac equation yields
\begin{eqnarray}\label{Equ}
\left(E_{_{\rm B}}+mc^2\right)\chi=c\sigma_i \pi^i\phi\qquad\
\textrm{and}\qquad\
\left(E_{_{\rm B}}-mc^2\right)\phi=c\sigma_i\pi^i\chi\, .
\end{eqnarray}
A uniform magnetic field is derivable from a vector potential $
{\mathbf{A}}=({\mathbf{B}}\times{\mbox{\boldmath$r$}})/2$, which
for $\mathbf{B}$ along the $z$--direction, becomes $
{\mathbf{A}}=-(yB\hat{x}-xB\hat{y})/2$. Such a field, without loss
of generality, can also be achieved by choosing a vector potential
as $A_x=0=A_z$ and $A_y=xB$. This way, one can make {\it
ansatz}~\cite{ItzyksonZuber}
\begin{equation}\label{Anz}
\phi(\mbox{\boldmath$r$})=h(x)\exp [i(p_y y+p_z z)/\hbar]
\end{equation}
to eliminate $\chi$ in equations~(\ref{Equ}) and then finds that
$h(\xi)\propto H_n(\xi)\exp( {-\xi^2/2})$, where $H_n$ represents
the Hermit polynomials. Noting that positive energy corresponds to
electron with negative charge ($-|e|$) while negative energy
describes positron with positive charge ($|e|$), and taking the
sign of $B$ such that $-|e|B>0$, one has
$\xi=\sqrt{-|e|B/c}\left(x\pm cp_y/|e|B\right)$ with the plus
(minus) sign for positive (negative) energy.
Thus, the energy spectrum of a particle with a magnetic moment
\mbox{\boldmath$\mu$} in a strong external magnetic field along
the $z$--direction, employing the cylindrical coordinates and
neglecting back--reaction of the motion of charge on the field,
takes the form~\cite{ItzyksonZuber,holten92}
\begin{equation}\label{Eng}
E_{_{\rm ln}}^2=m^2c^4+p_z^2c^2+ceB(n-l+1-2s_{z}),
\end{equation}
where $p_z=\hbar k$, $s_{z}=\pm\hbar/2$, $n$ is a non--negative
integer and $L_z=\hbar l$, while $l$ is restricted to values
$l=-n, -n+2, \cdots, n-2, n$.
The term $n-l+1$ represents contribution from the orbital part of
the magnetic moment, whereas we purpose to deal with effects of an
external magnetic field only on the internal spin. Thus, from the
spectrum~(\ref{Eng}), one can infer~\cite{holten93}
\begin{equation}\label{Eng1}
E_{_{\rm B}}^2=m^2c^4+p_z^2c^2-2ceBs_{z}\, .
\end{equation}
Defining $\sigma_{\pm}=\pm1$ as the eigenvalues of the operator
$\sigma_z$ with positive (negative) sign representing the up
(down) spin state, and reminding that positive (negative) energy
relates to electron (positron), relation~(\ref{Eng1}), in the weak
field approximation and non--relativistic limit (justified below
Eq.~(\ref{vawf})), reads
\begin{equation}\label{Omg}
\frac{E_{_{\rm B\pm}}}{\hbar}\simeq \pm\left(\frac{
mc^2}{\hbar}\pm\frac{\mid\! e\!\mid B \sigma_{\pm}}{2m
c}\right)\equiv \pm\left(\omega\pm\Omega\sigma_{\pm}\right),
\end{equation}
where $\omega\equiv\omega{_{\rm zbw}}/2$,\ $\Omega\equiv\,\mid\!
e\!\mid\! B/2m c=\omega_c/2$,\ $\omega_c$ is the usual classical
cyclotron frequency. The weak field approximation means
$\Omega\ll\omega$ or $\omega_c\ll\omega_{_{\rm zbw}}$, which is a
reasonable approximation even with the highest magnetic field
achievable in the current laboratories. One should also note that,
having the external magnetic field along the $z$--direction with
the proposed vector potential, the only altered generalized
momentum is $\pi_y$, while $\pi_z=p_z$ and $\pi_x=p_x$. In
addition, in order to justify use of Eq.~(\ref{Eng}) and also
simplify later on calculations, we ignore the non--commutativity
of $\pi_x$ and $\pi_y$ (justified below Eq.~(\ref{commut})).
As we are interested to find effects of an external magnetic field
on the $\omega_{_{\rm zbw}}$, following Huang's approach, it is
more appropriate and instructive to solve these equations by the
initial spinoral eigenstates. That is, using
$\phi^\uparrow=(h^{\uparrow}_+, 0)$ and $\phi^\downarrow=(0,
h^{\downarrow}_+)$ as the initial positive energy eigenstates for
the spin--up and spin--down, and $\chi^\uparrow=(h^{\uparrow}_-,
0)$ and $\chi^\downarrow=(0, h^{\downarrow}_-)$ as the initial
negative energy eigenstates for the spin--up and spin--down,
respectively. Then, by relation~(\ref{Omg}) and {\it
ansatz}~(\ref{Anz}), a general solution of Eqs.~(\ref{Equ})
becomes
\begin{eqnarray}\label{Ans}
\Psi({\mbox{\boldmath$r$}},t)=h^{-3/2}\int \Biggl\{{\cal A}\left(
\begin{array}{c}
\phi^\uparrow \\
k_{+}^\uparrow\mbox{\boldmath$\sigma$}\cdot\mbox{\boldmath$\pi$}\phi^\uparrow \\
\end{array}
\right)\exp\Bigl[-i(\omega+\Omega)t\Bigr]
+{\cal B}\left(
\begin{array}{c}
\phi^\downarrow \\
k_{+}^\downarrow\mbox{\boldmath$\sigma$}\cdot\mbox{\boldmath$\pi$}\phi^\downarrow \\
\end{array}
\right)\exp\Bigl[-i(\omega-\Omega)t\Bigr]\cr
+\,{\cal C}\left(
\begin{array}{c}
\!\!k_{-}^\uparrow\mbox{\boldmath$\sigma$}\cdot\mbox{\boldmath$\pi$}\chi^\uparrow\!\! \\
\chi^\uparrow \\
\end{array}
\right)\exp\Bigl[i(\omega-\Omega)t\Bigr]
+{\cal D}\left(
\begin{array}{c}
\!\!k_{-}^\downarrow\mbox{\boldmath$\sigma$}\cdot\mbox{\boldmath$\pi$}\chi^\downarrow\!\! \\
\chi^\downarrow\\
\end{array}
\right)\exp\Bigl[i(\omega+\Omega)t\Bigr]\,\Biggr\}\exp [i(p_y y+p_z z)/\hbar] \,d^3\pi,
\end{eqnarray}
where ${\cal A}$, ${\cal B}$, ${\cal C}$ and ${\cal D}$ are
complex functions of $\mbox{\boldmath$\pi$}$ and $k$'s. The
notation $k_{\pm}$ for positive and negative energies in the
absence of external fields are $k_{\pm}=\pm c/(E+mc^2)$ with
$E=(p^2c^2+ m^2c^4)^{1/2}$~\cite{huang1952}, while in our
situation, with the aid of relation~(\ref{Omg}), it reads
\begin{equation}\label{Norm1}
k_{-}^\uparrow= -k_{+}^\downarrow=\frac{-
c}{2mc^2-\hbar\Omega}\equiv- K_{1}
\qquad\quad {\rm and}\qquad\quad
k_{-}^\downarrow=-k_{+}^\uparrow =\frac{- c}{2mc^2+\hbar\Omega}\equiv-K_{2}\, .
\end{equation}
More explicitly, comparing solution~(\ref{Ans}) with the one in
the absence of magnetic field -- namely the plane wave
packet~(\ref{Wav}) -- reveals that one has
\begin{eqnarray}\label{UJ1}
\cases{
\exp(i\omega{t})\rightarrow{\exp[i(\omega-\Omega)t]}\ &\textrm{for negative energy}\cr
\cr
\exp(-i\omega{t})\rightarrow{\exp[-i(\omega+\Omega)t]}\quad &\textrm{for positive energy}\cr}
\end{eqnarray}
for the spin--up states, and
\begin{eqnarray}\label{DJ1}
\cases{
\exp(i\omega{t})\rightarrow{\exp[i(\omega+\Omega)t]}\ &\textrm{for negative energy}\cr
\cr
\exp(-i\omega{t})\rightarrow{\exp[-i(\omega-\Omega)t]}\quad &\textrm{for positive energy}\cr}
\end{eqnarray}
for the spin--down states. Let us write the wave
packet~(\ref{Ans}) in the form
\begin{eqnarray}\label{Ansc}
\Psi({\mbox{\boldmath$r$}},t)\!\!\!\!&=&\!\!\!h^{-3/2}\int\Biggl\{
C_{+}^\uparrow(\mbox{\boldmath$\pi$})h^\uparrow_+\exp\Bigl[ -i
(\omega+\Omega) t\Bigr]+
C_{+}^\downarrow(\mbox{\boldmath$\pi$})h^\downarrow_+\exp\Bigl[ -i
(\omega-\Omega) t\Bigr]\cr
&+&\!\!\!\!C_{-}^\uparrow(\mbox{\boldmath$\pi$})h^\uparrow_-\exp\Bigl[ i (\omega-\Omega) t\Bigr]
+\!C_{-}^\downarrow(\mbox{\boldmath$\pi$})h^\downarrow_-\exp\Bigl[ i (\omega+\Omega) t\Bigr]\,\Biggr\}
\exp [i(p_y y+p_z z)/\hbar] \,d^3\pi\, ,
{\qquad}
\end{eqnarray}
where
\begin{equation}\label{Spi}
C_{+}^\uparrow(\mbox{\boldmath$\pi$})\!={\cal A}\!\left(
\begin{array}{c}
1 \\
0 \\
\!\!\!K_{2}\pi_z \!\!\! \\
\!\!\!K_{2}\pi_+ \!\!\!\\
\end{array}
\right)\!,\ \
C_{+}^\downarrow(\mbox{\boldmath$\pi$})\!={\cal B}\!\left(
\begin{array}{c}
0 \\
1 \\
\!\!\!K_{1}\pi_- \!\!\! \\
\!\!\!-K_{1}\pi_z \!\!\!\\
\end{array}
\right)\!,\ \
C_{-}^\uparrow(\mbox{\boldmath$\pi$})\!={\cal C}\!\left(
\begin{array}{c}
\!\!\!-K_{1}\pi_z \!\!\! \\
\!\!\!-K_{1}\pi_+ \!\!\!\\
1 \\
0\\
\end{array}
\right)\!,\ \
C_{-}^\downarrow(\mbox{\boldmath$\pi$})\!={\cal D}\!\left(
\begin{array}{c}
\!\!\!-K_{2}\pi_- \!\!\! \\
\!\!\!K_{2}\pi_z \!\!\! \\
0 \\
1\\
\end{array}
\right)\!,
\end{equation}
where $\pi_{\pm}\equiv\pi_x\pm i\pi_y$. Implicitly, the
normalization condition of the wave packet~(\ref{Ansc}) is
provided if one requires that
\begin{eqnarray}\label{Norm6}
\int\left[\left|C^{\uparrow\ast}_+(\mbox{\boldmath$\pi$})\right|^2
+\left|C^{\uparrow\ast}_-(\mbox{\boldmath$\pi$})\right|^2
+\left|C^{\downarrow\ast}_+(\mbox{\boldmath$\pi$})\right|^2
+\left|C^{\downarrow\ast}_-(\mbox{\boldmath$\pi$})\right|^2\right]d^3\pi=1\,.
\end{eqnarray}
In order to write the exact forms of
$C^{\uparrow\downarrow}_\pm(\mbox{\boldmath$\pi$})$, one needs to
have an explicit expression for $\Psi$, which leads to evaluation
of $<\!\dot{\mbox{\boldmath$r$}}\!>$ = c
$<\!\mbox{\boldmath$\alpha$}\!>$. For this purpose, we take the
initial state of the wave packet as a localized spin--up electron
in the $z$--direction while its center is at rest in the origin,
namely
\begin{equation}\label{Int}
\Psi(\mbox{\boldmath$r$},0)=\left(
\begin{array}{c}
1 \\
0 \\
0 \\
0\\
\end{array}
\right)f\left(\frac{r}{r_{_o}}\right),
\end{equation}
where
$f(r/r_{_o})=\left[2/(\bar{\pi}r_{_o}^2)\right]^{3/4}\exp\left[-(r/r_{_o})^2\right]$
is a normalized Gaussian function with $r\equiv|r|$ and ${r_{_o}}$
as a constant that indicates approximate spatial extension of the
wave packet. The notation $\bar{\pi}$ has been introduced to
indicate the usual pi number in order not to be confused with the
generalized momentum $\mbox{\boldmath$\pi$}$. The Fourier
transformation of $f(r/r_{_o})$ is
\begin{equation}\label{Gus2}
f(\frac{\pi}{\pi_{_{o}}})=(\frac{2}{\bar{\pi}\pi_{_o}^2})^{3/4}\exp\Bigl[-(\frac{\pi}{\pi_{_o}})^2\Bigr],
\end{equation}
which, taking $\mbox{\boldmath$\pi$}$ in the spherical coordinates
with
$\mbox{\boldmath$\pi$}=\mbox{\boldmath$\pi$}(\pi,\theta,\varphi)$,
satisfies the normalization condition $\int
f^2(\pi/\pi_{_{o}})\,d^3\pi=1$, and where the constant
$\pi_{_{o}}=2\hbar/r_{_{o}}$ gives the width of wave packet in the
momentum space~\cite{huang1952}.
As we have applied the non--relativistic approximation, to
simplify the model further, we drop terms of the order
$\epsilon^2\equiv(\pi/2mc)^2$ or higher. Thus, in the weak field
approximation, $\Omega\ll\omega$, one has
$(\pi/2mc)(\hbar\Omega/2mc^2)\approx {\cal O}(\epsilon^2)$ or
higher, and then from~(\ref{Norm1}) one can write
\begin{equation}\label{Norm4}
\left(
\begin{array}{c}
K_1 \\
K_2 \\
\end{array}
\right)\pi
\simeq\frac{\pi}{2mc}\left(
\begin{array}{c}
1+\frac{\hbar\Omega}{2mc^2} \\
\\
1-\frac{\hbar\Omega}{2mc^2} \\
\end{array}
\right)\simeq\frac{\pi}{2mc}\left(
\begin{array}{c}
1 \\
1 \\
\end{array}
\right)\equiv K\pi\left(
\begin{array}{c}
1 \\
1 \\
\end{array}
\right)\!\, .
\end{equation}
In addition, during the calculation one encounters
non--commutative terms due to the existence of the magnetic field.
For our choice of vector potential, when $\pi_i$'s are operators
(i.e. not integration variables), one gets
\begin{eqnarray}\label{commut}
\pi_+\pi_- -\pi_-\pi_+=2i(\pi_y\pi_x-\pi_x\pi_y)=\frac{2 e
B\hbar}{c}\, .
\end{eqnarray}
Therefore, up to the order of approximation we are dealing with,
the non--commutativity of $\pi_x$ and $\pi_y$ can be
ignored\rlap,\footnote{Note that, for a typical field of one
Tesla, the right hand side of relation~(\ref{commut}) is an order
of about $10^{-63}$.}\
as assumed before, and we use $\pi_+\pi_- \simeq\pi_-\pi_+$ in the
following. Also, regarding the fact that
$h^{\uparrow\downarrow}_{\pm}(\xi)\propto H_n(\xi)\exp(
{-\xi^2/2})$, one gets $(\pi_x
h^{\uparrow\downarrow}_{\pm}-h^{\uparrow\downarrow}_{\pm}\pi_x)$
to be of the order $\hbar\sqrt{eB/c}$, where it can also be
ignored in the proposed approximation. Hence, using rough
coefficients~(\ref{Coff}) for the initial state (\ref{Int}), the
wave packet~(\ref{Ansc}) reads
\begin{eqnarray}\label{vawf}
\Psi({\mbox{\boldmath$r$}},t)\!\!\!&\simeq&\!\!\!h^{-3/2}\int\Biggl\{\left(
\begin{array}{c}
1 \\
0 \\
K\pi_z \\
K\pi_+ \\
\end{array}
\right)\exp\Bigl[-i(\omega+\Omega)t\Bigr]+\left(
\begin{array}{c}
0 \\
0 \\
-K\pi_z \\
0\\
\end{array}
\right)\exp\Bigl[i(\omega-\Omega) t\Bigr]\cr
& +&\!\!\!
\left(
\begin{array}{c}
0 \\
0 \\
0 \\
-K\pi_+ \\
\end{array}
\right)\exp\Bigl[i(\omega+\Omega) t\Bigr]\Biggr\}f(\pi/\pi_{_{o}})
\exp [i(p_y y+p_z z)/\hbar] \,d^3\pi\, .\qquad
\end{eqnarray}
Note that, each of the above rough solutions satisfy the Dirac
equation (\ref{Ham}) when the aforementioned approximations are
taken into account. Actually, by the exact
coefficients~(\ref{Coff}), they do fulfill the Dirac equation
exactly, and our approximations do~not alter the results for
altered Zbw, rather they simplify calculations.
Since the wave function must not spread rapidly in time, one can
justify the non--relativistic approximation applied
in~(\ref{Norm4}) if $r_{_{o}}$ is taken to be very large compared
to the Compton wavelength~\cite{huang1952} or equally
$\pi_{_{o}}/mc\ll1$. Therefore, all momenta which are greater than
$\pi_{_{o}}$ are diminished in the integrand~(\ref{vawf}), and
hence, the main role is played by momenta that satisfy
$\pi/mc\ll1$. This justifies the non--relativistic approximation.
Now, making use of the expectation value formula for velocity
vector,
\begin{equation}\label{Exp}
<\dot{\mbox{\boldmath$r$}}>=\int\Psi^{\ast}({\mbox{\boldmath$r$}},t)(c\mbox{\boldmath$\alpha$})
\Psi({\mbox{\boldmath$r$}},t)d^{3}r
\end{equation}
with the wave packet~(\ref{vawf}), it is straightforward to show
that for an initially localized spin--up electron
\begin{eqnarray}\label{Expv}
\!\!\!
<{\dot{\mbox{\boldmath$r$}}}>=\!\!{\mathbf{V}}\!\!\!+2c\int\Biggl\{l_{1}(\mbox{\boldmath$\pi$})
\cos\Bigl[(\omega_{_{\rm zbw}}\!\!\!+\omega_c)t+\varphi(\mbox{\boldmath$\pi$})\Bigr]
\!\!+l_{2}
(\mbox{\boldmath$\pi$})\cos\,\omega_{_{\rm zbw}} t\Biggr\}\,d^{3}\pi\, ,\!\!\!
\end{eqnarray}
where
\begin{equation}\label{Lean}
{\mathbf{V}}=
c\int\left(C^{\uparrow\ast}_+\mbox{\boldmath$\alpha$}
C_{+}^\uparrow +C^{\uparrow\ast}_-\mbox{\boldmath$\alpha$}
C_{-}^\uparrow+C^{\downarrow\ast}_-
\mbox{\boldmath$\alpha$}C_{-}^\downarrow\right)d^{3}\pi
\end{equation}
is the linear velocity of the average motion of electron,
$l_{1}(\mbox{\boldmath$\pi$})=\left|C^{\uparrow\ast}_+\mbox{\boldmath$\alpha$}
C_{-}^\downarrow\right|$,\ \,
$l_{2}(\mbox{\boldmath$\pi$})=\left|C^{\uparrow\ast}_+\mbox{\boldmath$\alpha$}
C_{-}^\uparrow\right|$
and the phase term turns out to be
\begin{equation}\label{Phas}
\varphi(\mbox{\boldmath$\pi$})=\tan^{-1}\frac{{\rm
Im}\left(C^{\uparrow\ast}_+\mbox{\boldmath$\alpha$}C_{-}^\downarrow\right)}{{\rm
Re}\left(C^{\uparrow\ast}_+\mbox{\boldmath$\alpha$}C_{-}^\downarrow\right)
}\ .
\end{equation}
Integration of Eq.~(\ref{Expv}) in the weak field approximation
reads
\begin{eqnarray}\label{Expv1}
<\!{\mbox{\boldmath$r$}}\!>\simeq {\mbox{\boldmath$r$}}_{_{\rm i}}\!\!+\!\!{\mathbf{V}}t+\!\!\lambdabar_c\!\!\int\!\Biggl\{l(\mbox{
\boldmath$\pi$})
\sin\Bigl[(\omega_{_{\rm zbw}}\!\!\!+\omega_{_{c}})t+\!\varphi(\mbox{\boldmath$\pi$})\Bigr]
\!\!+l_{2}(\mbox{\boldmath$\pi$})\sin\,\omega_{_{\rm zbw}}t\Biggr\}\,d^{3}
\pi,
\end{eqnarray}
where $l(\mbox{\boldmath$\pi$})\simeq
l_{1}(\mbox{\boldmath$\pi$})\left(1-\omega_c/\omega_{_{\rm
zbw}}\right)$
and we assume ${\mbox{\boldmath$r$}}_{_{\rm i}}\equiv{\mbox{\boldmath$r$}}_{\rm initial}=0$.
This relation, compared to the results with no external magnetic
field, shows that the oscillatory part -- which involves
interference between the positive and negative energy states of
electron -- is now split into two terms. As it will be discussed
bellow, within our approximation, the first term of the
oscillatory part signifies the Zbw circular motion in the
$xy$--plane with a shifted frequency, while the second term is
responsible for Zbw in the $z$--direction with an intact
frequency, as expected.
For the wave packet~(\ref{vawf}), the linear velocity term in
$<\!{\mbox{\boldmath$r$}}\!>$ vanishes. Whereas, up to the first
order in $\pi/mc$, we have
\begin{eqnarray}\label{Kfac2}
&l_{1x}=l_{1y}\simeq
-f^2(\frac{\pi}{\pi_{_{o}}})\frac{(\pi_x^2+\pi_y^2)^\frac{1}{2}}
{2mc}\, ,\qquad l_{2x}=l_{2y}= l_{1z}\simeq0\, ,\qquad
l_{2z}\simeq -f^2(\frac{\pi}{\pi_{_{o}}})\frac{\pi_z}{2mc}\, ,\cr
{}\cr
&\varphi_{x}(\mbox{\boldmath$\pi$})\simeq\tan^{-1}\left(\frac{\pi_y}{\pi_x}\right),\qquad
\varphi_{y}(\mbox{\boldmath$\pi$})\simeq\tan^{-1}\left(-\frac{\pi_x}{\pi_y}\right)\qquad
{\rm and}\qquad\varphi_{z}(\mbox{\boldmath$\pi$})=0.
\end{eqnarray}
In the spherical coordinates, one gets
\begin{eqnarray}\label{Kf}
l_{1x}=l_{1y}\simeq{-f^2(\frac{\pi}{\pi_{_{o}}})\frac{\pi\sin\theta}{2mc}}\,
,\qquad l_{2x}=l_{2y}= l_{1z}\simeq0\, ,\qquad l_{2z}\simeq
-f^2(\frac{\pi}{\pi_{_{o}}})\frac{\pi\cos\theta}{2mc}\, ,\cr
{}\cr
\varphi_{x}(\mbox{\boldmath$\pi$})\simeq\tan^{-1}\left(\tan\varphi\right)=\varphi\,
,\qquad
\varphi_{y}(\mbox{\boldmath$\pi$})\simeq\tan^{-1}\left(-\cot\varphi\right)=
\varphi+\frac{\bar{\pi}}{2}\qquad {\rm and}\qquad
\varphi_{z}(\mbox{\boldmath$\pi$})=0.
\end{eqnarray}
Substituting these terms in relation~(\ref{Expv1}), it becomes
\begin{eqnarray}\label{Spc1}
&<x>\simeq
I^\uparrow\frac{\lambdabar_c}{2}\int_{0}^{2\bar{\pi}}\sin\Bigl[\left(\omega_{\rm
zbw}+\omega_c\right)t+\varphi\Bigr]\, d\varphi\, ,\qquad
<y>\simeq
I^\uparrow\frac{\lambdabar_c}{2}\int_{0}^{2\bar{\pi}}\cos\Bigl[\left(\omega_{\rm
zbw}+\omega_c\right)t+\varphi\Bigr]\, d\varphi\cr {}\cr
&
{\rm and}\qquad <z>\simeq J\lambdabar_c\sin\left(\omega_{\rm zbw}
t\right),
\end{eqnarray}
where
\begin{equation}\label{Cospc}
I^\uparrow\equiv-2\!\left(1-\!\frac{\omega_c}{\omega_{_{\rm
zbw}}}\right)\!\int_{0}^{\infty}\!\int_{0}^{\bar{\pi}}\frac{f^2(\frac{\pi}{\pi_{_{o}}})}{2mc}
\pi^3\sin^2\theta\,d\theta
\,d\pi=-(8\bar{\pi})^{-\frac{1}{2}}\frac{\lambdabar_c}{r_{_o}}\left(1-\!\frac{\omega_c}{\omega_{_{\rm
zbw}}}\right)
\end{equation}
and
\begin{eqnarray}\label{Cospc1}
J\equiv-\bar{\pi}\int_{0}^{\infty}\int_{0}^{\bar{\pi}}\frac{f^2(\frac{\pi}{\pi_{_{o}}})}{2mc}\pi^3
\sin2\theta\,d\theta\,d\pi=0.
\end{eqnarray}
As $\varphi$ is the azimuthal angle in the spherical momentum
space, the nature of the integral in the $xy$--plane over
$\varphi$ restricts us to a fixed value, e.g. $\varphi_{_{o}}$, in
order to figure out the role of a specific Fourier decomposition
in the Zbw circular motion. Of course, if one accomplishes the
integrations over the full domain, then all oscillatory motions
will disappear, and the center of the wave packet as a whole
remains at rest, as expected. Hence, by considering those Fourier
components in the $xy$--plane within $\varphi$ and
$\varphi+d\varphi$, one gets
\begin{equation}\label{Spc11}
<x>_{\varphi_{_{o}}}\simeq
\frac{\lambdabar_c}{2}\sin\Bigl[\left(\omega_{_{\rm
zbw}}+\omega_{c}\right)t+\varphi_{_{o}}\Bigr]\qquad {\rm
and}\qquad
<y>_{\varphi_{_{o}}}\simeq
-\frac{\lambdabar_c}{2}\cos\Bigl[\left(\omega_{_{\rm
zbw}}+\omega_{c}\right)t+\varphi_{_{o}}\Bigr].
\end{equation}
The notation $I^\uparrow$ -- shown in Eq.~(\ref{Cospc}) to be
proportional to $\lambdabar_c/r_{_o}$ -- only determines the
weight of Zbw relative to the degree of localization of electron.
Eqs.~(\ref{Spc11}) express a circular motion in the $xy$--plane,
therefor one can conclude that the Zbw frequency of rotation for a
spin--up electron in an external uniform static magnetic field
changes by
\begin{equation}\label{Shiff1}
\omega^\uparrow_{_{\rm zbw}}\rightarrow\omega^\uparrow_{_{\rm
zbw}}+\omega_c\, ,
\end{equation}
while the frequency for $<\! z\!>$ remains unchanged, as expected.
The same procedure for the spin--down state yields
\begin{eqnarray}\label{Spc2}
&<x>\simeq
I^\downarrow\frac{\lambdabar_c}{2}\int_{0}^{2\bar{\pi}}\sin\Bigl[\left(\omega_{_{\rm
zbw}}-\omega_{c}\right)t-\varphi\Bigr]\, d\varphi \,,\qquad\quad
<y>\simeq
I^\downarrow\frac{\lambdabar_c}{2}\int_{0}^{2\bar{\pi}}\cos\Bigl[\left(\omega_{_{\rm
zbw}}-\omega_{c}\right)t-\varphi\Bigr]\, d\varphi\cr
{}\cr
&{\rm and}\qquad <z>\simeq J\lambdabar_c\sin\left(\omega_{_{\rm
zbw}} t\right),
\end{eqnarray}
with
\begin{equation}\label{Cospc2}
I^\downarrow\equiv-2\left(1+\frac{\omega_c}{\omega_{_{\rm
zbw}}}\right)\!\int_{0}^{
\infty}\!\int_{0}^{\bar{\pi}}\frac{f^2(\frac{\pi}{\pi_{_{o}}})}{2mc}
\pi^3\sin^2\theta\,d\theta d\pi
=-(8\bar{\pi})^{-\frac{1}{2}}\frac{\lambdabar_c}{r_{_o}}\left(1+\frac{\omega_c}{\omega_{_{\rm
zbw}}}\right).
\end{equation}
Thus, for a spin--down electron, the frequency changes by
\begin{equation}\label{Shiff2}
\omega^\downarrow_{_{\rm zbw}}\rightarrow\omega^\downarrow_{_{\rm
zbw}}-\omega_c\, .
\end{equation}
Obviously, the new Zbw frequencies reduce to the field free case
in the absence of external magnetic field.
Motion of electron in an external uniform static magnetic field,
employing the Heisenberg picture, has been discussed in
Ref.~\cite{Villavicencio-etal2000}. They have also shown that the
Zbw frequency, in the weak field approximation, changes to
\begin{equation}\label{UJ3}
\omega_{_{\rm zbw}}\pm\omega_c\, ,
\end{equation}
without explicitly distinguishing the spin--up from spin--down
states. However, we have proved that one can assign a distinct
modification to the Zbw frequency of each of these states given in
relations~(\ref{Shiff1}) and~(\ref{Shiff2}). Though, one may still
claim that $\omega_{_{\rm zbw}}\pm\omega_c$\ term can intuitively
be understood to stand for the spin--up/down states, respectively.
But, such an interpretation has explicitly been revealed within
our results. Besides, our outcome can be compared with the
classical results via the BZ approach~\cite{SalesiRecami2000},
that will be accomplished in the next section.
But before that, in order to gain further insight into the
situation raised by the phenomenon of Zbw, one can examine the
effect of an external uniform static magnetic field on some matrix
elements that do~not vanish. For this purpose, let us calculate
the expectation value of the magnetic moment of a spin--up/down
electron, that in the operator form it is~\cite{huang1952}
\begin{equation}\label{expmagmom}
<\!\mbox{\boldmath$\mu$}\!>=\frac{e}{2c}<\!{\mbox{\boldmath$r$}}\times\dot{\mbox{\boldmath$r$}}\!>
=\frac{e}{2}<\!{\mbox{\boldmath$r$}}\times\mbox{\boldmath$\alpha$}\!>\rightarrow
\frac{ie\hbar}{2}<\!\mbox{\boldmath$\nabla_{\pi}$}\times\mbox{\boldmath$\alpha$}\!>.
\end{equation}
Using the wave packet (\ref{vawf}), one can easily show that, with
the initial condition (\ref{Int}), the contribution from each
Fourier component of expression (\ref{expmagmom}) vanishes
individually in the $x$-- and $y$--directions, but survives in the
$z$--direction, as maybe expected. That is, from classical point
of view, the $z$--component of (\ref{expmagmom}), which is
perpendicular to the circle of Zbw motion, is the only
non--vanishing component of the magnetic moment. Straightforward
calculations, using the employed approximations, reveal that one
gets
\begin{equation}\label{expmagmomup}
<\!\mu^\uparrow_x\!>=0,\qquad<\!\mu^\uparrow_y\!>=0\qquad{\rm
and}\qquad <\!\mu^\uparrow_z\!>=-\frac{\mid\!
e\!\mid\lambdabar_c}{2}\biggl[1-\cos(\omega_{_{\rm
zbw}}+\omega_{c})t\biggr]
\end{equation}
for the spin--up states, and
\begin{equation}\label{expmagmomdown}
<\!\mu^\downarrow_x\!>=0,\qquad<\!\mu^\downarrow_y\!>=0\qquad{\rm
and}\qquad <\!\mu^\downarrow_z\!>=\frac{\mid\!
e\!\mid\lambdabar_c}{2}\biggl[1-\cos(\omega_{_{\rm
zbw}}-\omega_{c})t\biggr]
\end{equation}
for the spin--down states. These results indicate that each
Fourier wave contributes a circular motion about the direction of
spin. Hence, one may again interpret that the intrinsic magnetic
moment is a result of the (shifted) Zbw. Incidentally, the time
dependant part of the magnetic moment, in
relations~(\ref{expmagmomup}) and~(\ref{expmagmomdown}), is a
consequence of the fact that we have assumed that the spin of
electron had initially been observed. This requirement leads to
bring in the negative energy parts which do~not vanish by letting
the wave packet spread in space~\cite{huang1952}.
Furthermore, it would be instructive to employ the classical
interpretation of
spin~\cite{huang1952,hestenes93,barut-etal81b,hestenes90,hestenes2009}
via the intrinsic/microscopic angular momentum vector caused from
the current produced by the local Zbw frequency when this
frequency is shifted in the presence of an external magnetic
field. For this purpose, by relations~(\ref{Spc1})
and~(\ref{Spc2}) when considering~(\ref{Cospc})
and~(\ref{Cospc2}), one can claim that the radius of the Zbw
circle is $r_{_{\rm zbw}}=\lambdabar_c(1\pm\varepsilon)/2$, where
$\varepsilon\equiv -\omega_c/\omega_{_{\rm zbw}}$. And thus, the
speed of electron around the center of mass will be
\begin{equation}
v_{_{\rm
zbw}}=\left[\frac{\lambdabar_c}{2}(1\pm\varepsilon)\right]\left[\omega_{_{\rm
zbw}}(1\mp\varepsilon)\right]=c(1-\varepsilon^2)\simeq c\, .
\end{equation}
This result is consistent with the calculation of Zbw velocity via
the Heisenberg picture, where the position operator
${\mbox{\boldmath$r$}}$ satisfies $\dot{\mbox{\boldmath$r$}}=i[H,
{\mbox{\boldmath$r$}}]/\hbar=c{\mbox{\boldmath$\alpha$}}$. Hence,
the intrinsic/microscopic angular momentum is
\begin{equation}
s_{_{\rm zbw}}=r_{_{\rm zbw}}(mv_{_{\rm zbw}})\simeq
\frac{\hbar}{2}(1\pm\varepsilon)\simeq\frac{\hbar}{2(1\mp\varepsilon)}\,
.
\end{equation}
This indicates that the gyromagnetic $g$ factor of electron in the
presence of an external magnetic field changes to
$2(1\mp\varepsilon)$, however the magnetic moment, relations
(\ref{expmagmomup}) and (\ref{expmagmomdown}), shows that the time
average of the gyromagnetic $g$ factor in a period does~not
change.
On the other hand, if the spin of electron must remain unchanged,
i.e. with the fixed value of $\hbar/2$, then by assuming $v_{_{\rm
zbw}}\propto c$, (e.g. $v_{_{\rm zbw}}=\zeta c$), the radius of
the Zbw circle will be
\begin{equation}
r_{_{\rm zbw}}=\frac{v_{_{\rm zbw}}}{\omega_{_{\rm
zbw}}(1\mp\varepsilon)}=\frac{\lambdabar_c}{2}\left(\frac{\zeta}{1\mp\varepsilon}\right).
\end{equation}
Hence, the intrinsic/microscopic angular momentum is
\begin{equation}
s_{_{\rm
zbw}}=\frac{\lambdabar_c}{2}\left(\frac{\zeta}{1\mp\varepsilon}\right)(mv_{_{\rm
zbw}})
=\frac{\hbar}{2}\left(\frac{\zeta^2}{1\mp\varepsilon}\right),
\end{equation}
which if it is supposed to remain $\hbar/2$, one will get
$\zeta\simeq 1\mp\varepsilon/2$. That is, in this view, $v_{_{\rm
zbw}}\simeq c(1\mp\varepsilon/2)$ and $r_{_{\rm
zbw}}\simeq\lambdabar_c(1\pm\varepsilon/2)/2$.
\section{Frequency Shift via Classical Spinning Electron Approach}
\indent
As mentioned in the introduction, BZ
devised~\cite{SalesiRecami2000,BarutZanghi84,RecamiSalesi1995} a
new classical analogous model for the relativistic spinning
electron in the ordinary space--time by introducing a classical
Lagrangian as
\begin{equation}\label{Lag}
L=\frac{1}{2}i\alpha(\dot{\bar{z}}z-\bar{z}\dot{z})+
p_{\mu}(\dot{x}^\mu-\bar{z}\gamma^\mu z)+eA_{\mu}\bar{z}\gamma^\mu
z\, ,
\end{equation}
where $\mu=0, 1, 2, 3$, $A^\mu$ is the electromagnetic potential,
$\alpha$ is a constant with the dimension of action, $z$ and
$\bar{z}\equiv z^\dag\gamma^0$ are ordinary ${\mathcal{C}}^4$
bi--spinors which used as a second pair of conjugate classical
spinorial variables, the speed of light, $c$, is set equal to one
and dot indicates derivation with respect to the proper time,
$\tau$, where $d/d\tau=\dot{x}^\mu\partial_\mu$\,.
The four Euler--Lagrange equations can be obtained with respect to
variables $x^\mu$, $p^\mu$, $z$ and $\bar{z}$. By assuming
$-\alpha=1=\hbar$ and adopting independent dynamical variables
$x^\mu$, $\pi^\mu$, $v^\mu$ and $S^{\mu\nu}$, where $v^\mu$ is the
$4$--velocity and $S^{\mu\nu}\equiv i\bar{z}[\gamma^\mu,
\gamma^\nu]z/4$ is the {\it spin tensor} met in the Dirac theory,
the Euler--Lagrange equations are obtained within the classical
approach to be~\cite{SalesiRecami2000}
\begin{equation}\label{Magn0}
\dot{\pi}^\mu=eF^{\mu\nu}v_\nu\, ,\qquad\dot{x}^\mu=v^\mu\,
,\qquad\dot{v}^\mu=4S^{\mu\nu}\pi_\nu\qquad{\rm
and}\qquad\dot{S}^{\mu\nu}=v^\nu\pi^\mu-v^\mu\pi^\nu .
\end{equation}
For a free electron (i.e. $A^{\mu}=0$), the equation of motion for
$4$--velocity is derived to be~\cite{RecamiSalesi1995}
\begin{equation}\label{Moeq}
\mbox{$\upsilon$}^\mu=\frac{{p}^\mu}{m}
-\frac{\ddot{\mbox{$\upsilon$}}^\mu}{4m^2}\ .
\end{equation}
A general solution for this case has been shown to
be~\cite{barut-etal81a,BarutZanghi84,Barut-pavsic}
\begin{equation}\label{Velo}
\upsilon^{\mu}=\frac{p^\mu}{m}+\left[\upsilon^\mu(0)-\frac{p^\mu}{m}\right]\cos(\omega_{_{\rm
zbw}}\tau)+\frac{\dot{\upsilon}^\mu(0)}{2m}\sin(\omega_{_{\rm
zbw}}\tau)\, .
\end{equation}
The first term of this solution belongs to the rectilinear part of
the velocity, the second and third ones indicate the Zbw motion.
For a spinning electron in an external electromagnetic field,
third equation of Eqs.~(\ref{Magn0}) indicates that the general
solution represents a {\it helical} motion in the ordinary
$3$--space. In fact, the equation of motion from
Eqs.~(\ref{Magn0}) is derived to be~\cite{SalesiRecami2000}
\begin{equation}\label{Magn}
\ddot{\upsilon}^\mu-4m\pi^\mu+4m^2\upsilon^\mu+{4e\upsilon^\mu}F_{\alpha\beta}S^
{\alpha\beta}-4eS^{\mu\nu}F_{\nu\beta}\upsilon^\beta=0\, ,
\end{equation}
where $F^{\mu\nu}$\ is the electromagnetic tensor. Writing the
$x$-- and $y$--components of Eq.~(\ref{Magn}) for an electron with
the $z$--component of spin in an external uniform static magnetic
field along the $z$--direction (in which the only nonzero
components of the electromagnetic tensor are $F^{21}=-F^{12}=B$),
then deriving with respect to $\tau$ while using the first
equation of~(\ref{Magn0}) results in~\cite{SalesiRecami2000}
\begin{eqnarray}\label{Shif22unew2eqs}
\thrdotovervx+4\left(m^2-3\,es_zB\right)\dot{v}_x-4\,meBv_y=0\,,\cr
\thrdotovervy+4\left(m^2-3\,es_zB\right)\dot{v}_y+4\,meBv_x=0\,.
\end{eqnarray}
These equations describe a uniform circular motion in the
$xy$--plane whose angular velocities must satisfy the algebraic
equation~\cite{SalesiRecami2000}
\begin{equation}\label{Freq}
\omega^3-4\left(m^2-3\,es_zB\right)\omega+4\,meB=0\, ,
\end{equation}
or equivalently, with our notations,
\begin{equation}\label{NewFreq}
\omega^3-\omega_{_{\rm zbw}}^2\omega\pm
3\,\varepsilon\omega_{_{\rm zbw}}^2\omega+\varepsilon\omega_{_{\rm
zbw}}^3=0\, ,
\end{equation}
where the upper (lower) sign represents the spin--up (down) state.
Solutions of~(\ref{NewFreq}), by the weak field approximation and
up to the first--order terms in $\varepsilon $, have been asserted
to be~\cite{SalesiRecami2000}\footnote{Again, we have rewritten
their solutions with our notations.}
\begin{equation}\label{Omchr}
\omega_{_{1}}\simeq\omega_c\left(1\pm 3\,\varepsilon\right)\,
,\qquad \omega_{_{2}}\simeq\omega_{_{\rm
zbw}}\left(1\mp\frac{3}{2}\,\varepsilon\right)\qquad\textrm{and}\qquad
\omega_{_{3}}\simeq-\omega_{_{\rm
zbw}}\left(1\mp\frac{3}{2}\,\varepsilon\right).
\end{equation}
Frequency $\omega_{_{1}}$ has been
regarded~\cite{SalesiRecami2000} as the usual rotational frequency
of electron in an external magnetic field, which due to the Zbw,
differs from that of spinless particles. Namely, it shows the
effect of spin on the cyclotron frequency. It has also been
claimed~\cite{SalesiRecami2000} that the small predicted deviation
to the cyclotron frequency can be detected by their suggested
device. This argument nearly explains the purpose of
Ref.~\cite{SalesiRecami2000}. Meanwhile, the other two solutions,
$\omega_{_{2}}$ and $\omega_{_{3}}$, have been described as
modified forms of $\omega_{_{\rm zbw}}$ which, due to the
smallness of $\varepsilon$, have essentially been
considered~\cite{SalesiRecami2000} identical to $\omega_{_{\rm
zbw}}$ without further investigation.
In this work, we purpose to probe these terms more closely.
Actually, we are interested in investigating effects of external
magnetic fields on the Zbw frequency itself. Of course, there is
no doubt that due to the smallness of variations involved in this
issue, one has little hope of detection by the current
experiments, unless innovative device will
emerge\rlap.\footnote{Interested readers in related experimental
techniques may consult Ref.~\cite{Van-DyckSchwinbergDehmelt1986}
(and references therein), where they have measured the electron's
$g$ factor. Also, see
Refs.~\cite{rusin-etal2006,hestenes2009},~\cite{ZCetalFetalZJetalZZSetalK6}--\cite{LamataLeonSchatzSolano}
for more recent theoretical studies for experimentally observing
and/or experimental devices on the phenomenon of Zbw. For example,
an optical lattice scheme, which may permit experimental
observation of the Zbw with ultra--cold neutral atoms, has been
proposed in Ref.~\cite{VaishnavClark2008}, in where they also
claim that the phenomenon of Zbw will occur at experimentally
accessible frequencies and amplitudes.}\
Examples fall short of the level of this effect, even detection of the
$\omega_{_{\rm zbw}}$ itself for free electron are outside of the
current experimental and technological
reach~\cite{rusin-etal2006,VaishnavClark2008,LamataLeonSchatzSolano}.
Unless, e.g., they are indirectly in the response of electron to
external fields and/or a kind of condition (e.g. a resonance
condition) employed to minimize the Zbw frequency. For example,
the Darwin term in the Hamiltonian of an atomic electron comes
from the Zbw motion which makes electron sensitive to the
potential of the nuclei in its average
position~\cite{Dirac58}--\cite{Merzbacher98}. Furthermore, to
grasp an intuitive sense, one may recall that the Zbw frequency of
oscillation is too high and hence the period of the time involved
is too low, whereas the time measurement sensitivity limits are
generally known to be about $10^{-16} {\rm
s}$~\cite{timeprecision}.
Nevertheless, following our purpose while probing the solutions
$\omega_{_{2}}$ and $\omega_{_{3}}$, we have noticed that these
two solutions of~(\ref{Omchr}), are very rough approximations.
Actually, the more accurate solutions $\omega_{_{2}}$ and
$\omega_{_{3}}$, after solving equation (\ref{NewFreq}), are
\begin{eqnarray}\label{Shif2}
\omega_{_{2}}\simeq\omega_{_{\rm zbw}}({{1-2\varepsilon}\atop
{1+\varepsilon}})\ \quad\qquad \textrm{and}\ \quad\qquad
\omega_{_{3}}\simeq-\omega_{_{\rm zbw}}({{1-\varepsilon}\atop
{1+2\varepsilon}}),
\end{eqnarray}
where again the upper (lower) value represents solution for the
spin--up (down) state. These solutions contradict with the
solutions derived from the quantum approaches, i.e. the approach
employed in this work via the expectation value and the one used
in Ref.~\cite{Villavicencio-etal2000} via the Heisenberg picture.
The more accurate solutions (\ref{Shif2}) lead to some
discrepancies. In fact, regarding solution~(\ref{Velo}) and via
the BZ spinning particle model, Salesi and Recami have associated
$(-\omega_{_{\rm zbw}})$ with antiparticles and $\omega_{_{\rm
zbw}}$ with particles~\cite{SalesiRecami1997}. This way, they have
concluded that the only change which happens to the Zbw part of
the motion via $\omega_{_{\rm zbw}}\rightarrow-\omega_{_{\rm
zbw}}$ would be conversion of counterclockwise motion to the
clockwise one. It means that solution~(\ref{Velo}) should treat
electrons and positrons, other than assigning a different
direction of rotation to their Zbw motions, in the same spin
states. In this merit, it has also been
inferred~\cite{SalesiRecami2000} that $\omega_{_{2}}$ and
$\omega_{_{3}}$ are generalized features of $\omega_{_{\rm zbw}}$
and $(-\omega_{_{\rm zbw}})$, respectively.
Now, let us investigate this issue for the
solutions~(\ref{Shif2}), in which they can be rewritten for
electron and positron with the spin--up as\footnote{Note that,
according to our convention $\varepsilon$ has a negative value.}
\begin{eqnarray}\label{Shif22unew}
\omega^\uparrow_{_{\rm
zbw}}&\rightarrow&\omega^\uparrow_{_{2}}\simeq\omega^\uparrow_{_{\rm
zbw}}-2\,\varepsilon\omega^\uparrow_{_{\rm zbw}}\cr {}\cr
(-\omega^\uparrow_{_{\rm
zbw}})&\rightarrow&\omega^\uparrow_{_{3}}\simeq
(-\omega^\uparrow_{_{\rm
zbw}})-\varepsilon(-\omega^\uparrow_{_{\rm zbw}})\, ,
\end{eqnarray}
while with the spin--down as
\begin{eqnarray}\label{Shif22dnew}
\omega^\downarrow_{_{\rm
zbw}}&\rightarrow&\omega^\downarrow_{_{2}}\simeq\omega^\downarrow_{_{\rm
zbw}}+\varepsilon\omega^\downarrow_{_{\rm zbw}}\cr {}\cr
(-\omega^\downarrow_{_{\rm
zbw}})&\rightarrow&\omega^\downarrow_{_{3}}\simeq
(-\omega^\downarrow_{_{\rm
zbw}})+2\,\varepsilon(-\omega^\downarrow_{_{\rm zbw}})\, .
\end{eqnarray}
These relations show that the existence of an external magnetic
field breaks the usual symmetries which hold for free particles
and antiparticles. That is, the spin--up electron experiences a
new Zbw frequency while the one for the spin--up positron is
different, and vice versa for the spin--down states. In addition,
the Zbw frequency of an electron (positron) exposed to an external
magnetic field behaves oddly, and the absolute value of frequency
shift depends on the orientation of spin. Indeed, the Zbw
frequency of the spin--up particle changes twice that of the
spin--down, and vice versa for antiparticles. Therefore, the usual
symmetry of free particles and antiparticles states, which is
valid in the electromagnetic interactions, is broken. More
obvious, it means that the well--known CP symmetry, which is
respected in the quantum electrodynamics\rlap,\footnote{See
standard textbooks on the quantum field theory, e.g.
Ref.~\cite{ItzyksonZuber}. }\
is violated, however, the quantum mechanical symmetries
might not necessarily hold in the classical level.
Incidentally, if one repeats the calculations of Sec.~$2$ with the
same initial state but for positron instead of electron, namely
\begin{equation}\label{Intposit}
\Psi({\mathbf{r}},0)=\left(
\begin{array}{c}
0 \\
0 \\
1 \\
0\\
\end{array}
\right)f\left(\frac{r}{r_{_o}}\right)
\end{equation}
instead of relation~(\ref{Int}), one will get
\begin{equation}\label{Shiff1posit}
\omega^\uparrow_{_{\rm zbw}}\rightarrow\omega^\uparrow_{_{\rm
zbw}}-\omega_c
\end{equation}
and
\begin{equation}\label{Shiff2posit}
\omega^\downarrow_{_{\rm zbw}}\rightarrow\omega^\downarrow_{_{\rm
zbw}}+\omega_c
\end{equation}
instead of shifts~(\ref{Shiff1}) and~(\ref{Shiff2}). Now, if one
can justify -- e.g. based on the initial states~(\ref{Int})
and~(\ref{Intposit}) -- the application of the classical
interpretation of Ref.~\cite{SalesiRecami1997} for the expectation
value approach, then one will easily be able to show that this
approach holds the above discussed
symmetries\rlap.\footnote{However, one cannot make such a
classical energy separation within the relativistic quantum
approaches, see, e.g., the wave packet~(\ref{vawf}).}\
In another words, our results, relations (\ref{Shiff1}), (\ref{Shiff2}),
(\ref{Shiff1posit}) and (\ref{Shiff2posit}), indicate that the
amount of frequency shifts are opposite in the spin--up/down
states and in electron/positron cases, which are consistent with
the CP symmetry.
\section{Discussion and Conclusion}
\indent
Much effort in the literature has been devoted to putting forward
a consistent theory to describe the spin of electron. Among them,
Huang~\cite{huang1952} used the Dirac equation to tackle this
issue and BZ~\cite{BarutZanghi84} introduced a new classical
model, which is believed to be the most satisfactory picture of a
classical spinning electron constituting a natural classical limit
of the Dirac equation~\cite{SalesiRecami2000,Barut-pavsic}.
In this work, we have employed the expectation value quantum
approach~\cite{huang1952} to investigate the effects of an
external magnetic field on the Zbw frequency of electron. We have
found that, in the non--relativistic weak field approximation, the
Zbw frequency in the presence of an external uniform static
magnetic field changes according to relations~(\ref{Shiff1})
and~(\ref{Shiff2}) for electrons, and
relations~(\ref{Shiff1posit}) and~(\ref{Shiff2posit}) for
positrons. Actually, the individual Fourier components of the
position expectation value of an electron with the certain initial
conditions do~not vanish, and indeed lead to an interpretation in
terms of a shifted Zbw frequency. Furthermore, not only each of
the Fourier wave contributes a circular motion in the plane
perpendicular to the direction of electron spin with radius
$\lambdabar_c/2$, rather the shifted Zbw frequency depends equally
on the orientation of the direction as well, i.e. the spin--up or
down. Our results are exactly the same as those derived in
Ref.~\cite{Villavicencio-etal2000} via the Heisenberg picture.
However, contrary to Ref.~\cite{Villavicencio-etal2000}, we have
explicitly shown the distinct modifications for the spin--up and
spin--down states, as one may intuitively expect. Besides, in
order to compare the classic and quantum approaches to the Zbw in
an external magnetic field, this characteristic of our results are
more appropriate.
The BZ classical approach has been
generalized~\cite{SalesiRecami2000} to include the case of an
electron in the presence of an external uniform static magnetic
field. Hence, it is shown that in the weak field approximation,
the motion of electron is described by the three characteristic
frequencies, where one of them reflects the effect of spin on the
cyclotron frequency and the other two are the modified forms of
$\omega_{_{\rm zbw}}$. We have noticed that the latter two derived
frequencies are very rough approximations. Thus, we have indicated
the more accurate solutions, solutions~(\ref{Shif2}). However,
contrary to the quantum approaches -- the approach employed in
this work via the expectation value and the one used in
Ref.~\cite{Villavicencio-etal2000} via the Heisenberg picture --
the solutions~(\ref{Shif2}) are controversial and break the usual
symmetry of free particles and antiparticles states. We have made
this point more obvious by writing the solutions~(\ref{Shif2}) via
the BZ~\cite{SalesiRecami1997} spinning particle and antiparticle
interpretation. Indeed, these solutions,
relations~(\ref{Shif22unew}) and~(\ref{Shif22dnew}), indicate that
the frequency shifts depend nontrivially on the orientation of
spin and also on whether one is dealing with particles or
antiparticles. Hence, CP violation occurs in the classical
electrodynamics domain. Therefore, regarding the Zbw frequency of
electron, one may conclude that the BZ classical approach in the
presence of an external magnetic field is unlikely to correctly
describe the spin of electron, while the quantum approach does.
That is, the quantum approach respects the CP symmetry, as
expected.
\setcounter{equation}{0}
\renewcommand{\theequation}{A.\arabic{equation}}
\section*{Appendix}
\indent
Using the initial state~(\ref{Int}) and
approximations~(\ref{Norm4}),~(\ref{commut}) and $(\pi_x
h^{\uparrow\downarrow}_{\pm}\simeq
h^{\uparrow\downarrow}_{\pm}\pi_x)$, the exact coefficients in
relation~(\ref{Spi}) are found roughly to be
\begin{eqnarray}\label{Coff}
{\cal
A}\!\!&=&\!\!\frac{h^\uparrow_-h^\downarrow_+h^\downarrow_-+h^\uparrow_-
h^\downarrow_+h^\downarrow_-K^2_1\pi^2_z+h^\downarrow_-K_1K_2\pi_-h^\downarrow_+\pi_+h^\uparrow_-}
{\Gamma}f(\pi/\pi_{_{o}})\simeq\frac{f(\pi/\pi_{_{o}})}{h^\uparrow_+}\ ,\cr
{\cal
B}\!\!&=&\!\!\frac{-h^\downarrow_-K^2_2\pi_+h^\uparrow_-\pi_zh^\uparrow_+
+h^\uparrow_-h^\downarrow_-K_1K_2\pi_z\pi_+h^\uparrow_+}{\Gamma}f(\pi/\pi_{_{o}})\simeq 0
\ ,\cr
{\cal
C}\!\!&=&\!\!-\frac{K^2_1K_2\pi_-h^\downarrow_+\pi_zh^\downarrow_-\pi_+h^\uparrow_+
+h^\uparrow_+h^\downarrow_+h^\downarrow_-K^2_1K_2\pi^3_z
+h^\uparrow_+h^\downarrow_+h^\downarrow_-K_2\pi_z}{\Gamma}
f(\pi/\pi_{_{o}})\cr
\!\!&\simeq &\!\!\frac{-K\pi_z}{h^\uparrow_-}f(\pi/\pi_{_{o}})\ ,\cr
{\cal
D}\!\!&=&\!\!-\frac{h^\uparrow_+h^\downarrow_+K_1K^2_2\pi^2_z\pi_+h^\uparrow_-
+K_1K^2_2\pi_+h^\uparrow_+\pi_-h^\downarrow_+\pi_+h^\uparrow_-
+h^\uparrow_-h^\downarrow_+K_2\pi_+h^\uparrow_+}{\Gamma}f(\pi/\pi_{_{o}})\cr
\!\!&\simeq &\!\!\frac{-K\pi_+}{h^\downarrow_-}f(\pi/\pi_{_{o}})\ ,
\end{eqnarray}
where
\begin{eqnarray}\label{gamma}
\Gamma\!\!&\equiv&\!\!h^\uparrow_+h^\uparrow_-h^\downarrow_+h^\downarrow_-
+h^\uparrow_+h^\downarrow_+K_1^2K_2^2\pi_z^2\pi_-h^\downarrow_-\pi_+h^\uparrow_-
+h^\uparrow_-h^\downarrow_-K_1^2K_2^2\pi_z^2\pi_-h^\downarrow_+\pi_+h^\uparrow_+\cr
&&\!\!\!\!+h^\uparrow_+h^\uparrow_-h^\downarrow_+h^\downarrow_-K_1^2\pi_z^2
+h^\uparrow_+h^\uparrow_-h^\downarrow_+h^\downarrow_-K_1^2K_2^2\pi_z^4
+h^\uparrow_+h^\uparrow_-h^\downarrow_+h^\downarrow_-K_2^2\pi_z^2\cr
&&\!\!\!\!+h^\uparrow_+h^\downarrow_-K_1K_2\pi_-h^\downarrow_+\pi_+h^\uparrow_-
+K_1^2K_2^2\pi_-h^\downarrow_+\pi_-h^\downarrow_-\pi_+h^\uparrow_-\pi_+h^\uparrow_+
+h^\uparrow_-h^\downarrow_+K_1K_2\pi_+h^\uparrow_+\pi_-h^\downarrow_-\, .
\end{eqnarray}
\section*{Acknowledgement}
\indent
We thank the Research Office of the Shahid Beheshti University for
financial support. M.Z.--A. would like to appreciate his
supervisor M.F. for very useful comments which have made the
completion of this work possible.
|
\section{Introduction}
The gravitational lens is an important tool in astrophysics for probing
mass distributions and determining the cosmological
parameters.\cite{esf}
Although previous studies of gravitational lensing have been mostly based on
simple, spherically symmetric lens models,
some generalizations which include rotation\cite{rf:2}\tocite{ak}
and higher-order general
relativistic effects\cite{eoak,aky} have also been developed.
In this paper, we investigate another generalization, i.e., the effect
of non-spherical distortion of the gravitational potential in a
compact lens object on the gravitational
lensing.
A pioneering work on the non-spherically deformed compact lens was
done by Asada.\cite{asada}
He studied analytically the gravitational lensing caused by a
non-spherically deformed star, where the non-spherical property of the
gravitational potential was modeled by a quadrupole moment.
He employed a rigorous analytic approach and presented solutions of
the image positions for a source on the principal axis, an expression of the caustics and the critical curves.
However, because of the highly non-linear nature of the lens
equation, general solutions of the image positions for a off-axis
source and their amplification factors have not been presented.
In this paper, we take a more practical approach, namely the
perturbative approximation, and solve analytically the lens equation
and obtain the approximate solutions up to the lowest order of the
quadrupole moment.
We also calculate the Jacobian of the lens mapping and indicate how
the quadrupole moment changes the amplification factors.
\section{Lens equation}
In this section,
we summarize the lens equation for a compact object with a quadrupole
moment. See also Asada\cite{asada} for detail.
The gravitational potential of a non-spherically deformed compact
object is modeled by use of a monopole and a quadrupole moment:
\begin{equation}
\phi = \phi_0 + \phi_2 = -\frac{GM}{r} -
\frac{G}{2}\left(3\frac{x^i x^j}{r^5} - \frac{\delta^{ij}}{r^3}
\right) I_{ij},
\end{equation}
where
\begin{equation}
M = \int \rho\, d^3x,\quad
I_{ij} = \int \rho\, x_i x_j d^3 x,
\end{equation}
and $\rho$ is the mass density of the lens object.
Using the above form of the potential, the
deflection angle 2-vector $\boldsymbol\alpha$ is calculated as follows:
\begin{equation}
\alpha^i = \frac{4GM}{c^2}\frac{\xi^i}{|{\boldsymbol\xi}|^2}
+ \frac{8G}{c^2} \left(
2Q_{jk}\frac{\xi^j\xi^k \xi^i}{|{\boldsymbol\xi}|^6} -
Q_{ij}\frac{\xi^j}{|{\boldsymbol\xi}|^4}\right),
\end{equation}
where the 2-dimensional vector
$\boldsymbol \xi$ denotes the image position, and
\begin{equation}
Q_{ij} = \int\rho\,\left(X_iX_j -
\frac{1}{2}\delta_{ij}|{\boldsymbol X}|^2
\right) d^3 X
\end{equation}
denotes the trace-free quadrupole moment.
Without loss of generality, we can assume the normalized quadrupole moment is
diagonalized as follows,
\begin{equation}
\tilde{Q}_{ij} \equiv \frac{c^2 D_S}{2GM^2 D_L D_{LS}} Q_{ij}
= \left(
\begin{array}{cc}
e & 0\\
0 & -e\\
\end{array}
\right),
\end{equation}
where $D_S, D_L, D_{LS}$ are the angular diameter distances from the
observer to the source, from the observer to the lens, and from the
lens to the source, respectively.
Hereafter, we assume $e>0$.
Finally, the lens equation is
\begin{equation}\label{eq:leq1}
\beta_x = x - \frac{x}{x^2 + y^2} - e
\frac{(x^2-3y^2) x}{(x^2+y^2)^3},
\end{equation}
\begin{equation}\label{eq:leq2}
\beta_y = y - \frac{y}{x^2 + y^2} - e
\frac{(3x^2-y^2) y}{(x^2+y^2)^3},
\end{equation}
where ${\boldsymbol\beta}=(\beta_x,\beta_y)$ and
${\boldsymbol\theta}=(x, y)$ are the source and the image positions,
respectively, normalized by the Einstein radius $\theta_E =
\sqrt{4GMD_{LS}/c^2 D_L D_S}$.
\section{Perturbative approach}
A comprehensive, algebraic study of the above lens equation
Eqs.~(\ref{eq:leq1}) and (\ref{eq:leq2}) was done by Asada\cite{asada} using
polar coordinates. He showed the lens equation for this type as a
single real 10th-order algebraic equation and gave analytic solutions
for a source located exactly on the principal axes.
He also estimated the order of magnitude of the normalized quadrupole
moment as
\begin{equation}
e \sim 10^{-5} \left(\frac{M_{\odot}}{M}\right)
\left(\frac{R}{10^6 \,\mbox{km}}\right)^3
\left(\frac{10^7 \,\mbox{km}}{R_{\rm{E}}}\right)^2
\left(\frac{v}{10\,{\mbox{km s$^{-1}$}}}\right)^2,
\end{equation}
where $R$ denotes the typical size of a lens object, $R_{\rm{E}}$
is the Einstein ring radius, and $v$ is a rational surface velocity
on the equatorial plane of the lens star.
Then, for a nearby solar-type star at 10 pc, i.e., $R\sim 10^6$ km,
$R_{\rm{E}} \sim 10^7$ km, with a rotational
velocity of $v\sim 10$ km s$^{-1}$ which is much faster than that of
the Sun, the estimated value is $e \sim 10^{-5}$ , which is
sufficiently smaller
than unity.
Because
of the higher polynomial nature of the lens equations, however, general
analytic solutions for a source not on the axes have not been given
yet (or cannot be given for more than 4 images). Also, the physical
properties of the lensed images, such as the image amplification
factors, are not yet investigated.
Instead of such an exact algebraic treatment, in this paper, we take another, a
more tractable approach to the quadrupole lens equations.
Using the fact $0<e\ll 1$,
we employ a
perturbative approach to the lens equations and obtain approximate
solutions to the image positions $(x,y)$, up to the lowest order of
the eigenvalue $e$ of the quadrupole moment.
Concerning the perturbative approach to the gravitational lensing,
for example,
Alard\cite{a1}\tocite{a3} wrote a series of papers in a slightly different
context from ours.
His main interest is the non-spherical perturbations of extended lens
models and elongated arc images, whereas in this paper we concentrate our
attention on a non-spherically distorted compact lens model and multiple
images.
We start from the following set of the lens equation, which is
obtained by clearing the fraction of Eqs.~(\ref{eq:leq1}) and
(\ref{eq:leq2}):
\begin{equation}\label{lenseq1}
(x^2+y^2)^2 \left\{
(x^2+y^2)(\beta_x - x) + x\right\} = -e (x^2-3 y^2) x,
\end{equation}
\begin{equation}
\label{lenseq2}
(x^2+y^2)^2 \left\{
(x^2+y^2)(\beta_y - y) + y\right\} = -e (3 x^2-y^2) y.
\end{equation}
Setting $e=0$, we obtain the zeroth-order solutions,
\begin{equation}
\label{0th-sols}
x=x_0^{\pm} \equiv f^{\pm} \beta_x, \quad y=y_0^{\pm} \equiv f^{\pm} \beta_y,
\end{equation}
where
\begin{equation}
f^{\pm} \equiv \frac{1\pm \sqrt{1+4\beta^{-2}}}{2},
\end{equation}
and $\beta=\sqrt{\beta_x^2+\beta_y^2}$.
The above solutions are just the solutions for a point mass lens.
We may also find a trivial solution $(x, y)=(0,0)$, but it is inadequate
because the denominators in the original lens equation
Eqs.~(\ref{eq:leq1}) and (\ref{eq:leq2}) vanish.
Next, we put the following form into Eqs.~(\ref{lenseq1}) and
(\ref{lenseq2}),
\begin{equation}
x=x_0 + x_1, \quad y=y_0 + y_1,
\end{equation}
then, up to the linear order of $x_1$ and $y_1$, we obtain
\begin{equation}
(2x_0x _1 + 2 y_0 y_1)(\beta_x-x_0)- (x_0^2+y_0^2) x_1 + x_1 =
-e\frac{(x_0^2-3y_0^2)x_0}{(x_0^2+y_0^2)^2},
\end{equation}
\begin{equation}
(2x_0x _1 + 2 y_0 y_1)(\beta_y-y_0)- (x_0^2+y_0^2) y_1 + y_1 =
-e\frac{(3x_0^2-y_0^2)y_0}{(x_0^2+y_0^2)^2}.
\end{equation}
The solutions are expressed in terms of $\beta_x$ and $\beta_y$ as
follows:
\begin{equation}\label{majorx}
x^{\pm}=x_0^{\pm}+x_1^{\pm} = f^{\pm}\beta_x +
e\frac{(4\beta_x^2-3\beta^2)(f^{\pm})^2-1}
{\beta^2(f^{\pm}\beta^2+1)(f^{\pm}\beta^2+2)}\beta_x,
\end{equation}
\begin{equation}\label{majory}
y^{\pm}=y_0^{\pm}+y_1^{\pm} = f^{\pm}\beta_y +
e\frac{(3\beta^2-4\beta_y^2)(f^{\pm})^2+1}
{\beta^2(f^{\pm}\beta^2+1)(f^{\pm}\beta^2+2)}\beta_y.
\end{equation}
In the case of a point mass lens model ($e=0$), the number of the images is
two. As shown by Asada,\cite{asada} the number of the images by a
quadrupole lens ($e\neq 0$), which
depends on the value of $e$ and the position of the source, is
generally more than two, in most cases four.
We can obtain another, new solutions for $e>0$ case, which are not
existent for a point mass lens case, as a perturbation
around the trivial solution $(x,y)=(0,0)$, namely,
\begin{equation}
x=0 + x_1, \quad y=0 + y_1.
\end{equation}
Up to the lowest non-trivial order,
the lens equation is
\begin{equation}\label{eq311}
(x_1^2+y_1^2)^2 x_1=-e(x_1^2-3y_1^2) x_1,
\end{equation}
\begin{equation}\label{eq312}
(x_1^2+y_1^2)^2 y_1=-e(3x_1^2-y_1^2) y_1.
\end{equation}
In the case of $x_1\neq 0, y_1\neq 0$, Eqs.~(\ref{eq311}) and
(\ref{eq312}) imply
$x_1^2 = -y_1^2$, which does not have a real root.
In the case $x_1\neq 0, y_1=0$, Eq.~(\ref{eq311}) implies
$x_1^2 = -e$, which again does not have a real root since we have
assumed $e>0$. Finally in the case $x_1=0, y_1\neq 0$,
Eq.~(\ref{eq312})
reduces to $y_1^2=e$, from which we obtain
\begin{equation}
x = x_1 = 0, \quad y=y_1 = \pm \sqrt{e}.
\end{equation}
The above solutions are always on the $y$-axis, independent of the
source position $(\beta_x,\beta_y)$, and order of $O(\sqrt{e})$, not
$O(e)$. Therefore, we perform one more iteration and obtain slightly
higher order solutions in the following form,
\begin{equation}
x=0+x_2, \quad y=0+y_1 + y_2 = \pm \sqrt{e} + y_2.
\end{equation}
Up to the lowest order of $x_2$ and $y_2$, the lens equation is now
\begin{equation}
y_1^6\beta_x + y_1^4 x_2 = 3ey_1^2 x_2,
\end{equation}
\begin{equation}
y_1^6 (\beta_y -y_1) + 5y_1^4 y_2 = 3 e y_1^2 y_2,
\end{equation}
and the solution is
\begin{equation}\label{minor}
x = x_2 = \frac{1}{2} e \beta_x, \quad
y = y_1 + y_2 = \pm \sqrt{e}\left(1+\frac{1}{2}e\right)
- \frac{1}{2} e \beta_y.
\end{equation}
These images always appear very close to the $y$-axis, are very dim
because the amplitude of the amplification factor is $O(e^2)$, and
disappear when $e\rightarrow 0$. We call them as the ``minor''
images, and may safely neglect them in the analysis of the image
amplifications.
\begin{table}[htb]
\caption{Comparison of our approximate solutions with numerical ones
for $\beta_x=0$}
\label{table1}
\centering
\begin{tabular}{cccrrrrrr}
\hline\hline
$e$ & $\beta_x$ & $\beta_y$ &
$x_{\rm num}$ & $x_{\rm appr}$ &
(1) $y_{\rm num}$ & (2) $y_{\rm appr}$ &
$|(2)-(1)|$ \\
\hline
0.01 & 0.0 & 0.2 & 0.0 & 0.0 & 1.10087 & 1.10091 & $3.9\times 10^{-5}$ \\
& & & 0.0 & 0.0 & -0.89881 & -0.89891 & $1.0\times 19^{-4}$ \\
& & & 0.0 & 0.0 & 0.09950 & 0.099500 & $3.7\times 10^{-6}$ \\
& & & 0.0 & 0.0 & -0.10157 & -0.10150 & $6.8\times 10^{-5}$ \\
\hline
0.01 & 0.0 & 0.5 & 0.0 & 0.0 & 1.27780 & 1.27782 & $1.8\times 10^{-5}$ \\
& & & 0.0 & 0.0 & -0.77262 & -0.77282 & $2.0\times 10^{-4}$ \\
& & & 0.0 & 0.0 & 0.09809 & 0.09800 & $8.5\times 10^{-5}$ \\
& & & 0.0 & 0.0 & -0.10327 & -0.10300 & $2.7\times 10^{-4}$ \\
\hline
0.02 & 0.0 & 0.2 & 0.0 & 0.0 & 1.09668 & 1.09684 & $1.6\times 10^{-4}$ \\
& & & 0.0 & 0.0 & -0.89241 & -0.89284 & $4.2\times 10^{-4}$ \\
& & & 0.0 & 0.0 & 0.14084 & 0.14084 & $2.5\times 10^{-7}$ \\
& & & 0.0 & 0.0 & -0.14510 & -0.14484 & $2.7\times 10^{-4}$ \\
\hline
0.02 & 0.0 & 0.5 & 0.0 & 0.0 & 1.27479 & 1.27486 & $7.3\times 10^{-5}$ \\
& & & 0.0 & 0.0 & -0.76402 & -0.76486 & $8.4\times 10^{-4}$ \\
& & & 0.0 & 0.0 & 0.13802 & 0.13784 & $1.8\times 10^{-4}$ \\
& & & 0.0 & 0.0 & -0.14878 & -0.14784 & $9.5\times 10^{-4}$ \\
\hline
\end{tabular}
\end{table}
In order to check the accuracy of our approximate solutions
Eqs.~(\ref{majorx}), ({\ref{majory}), and (\ref{minor}),
we compare them with numerical solutions. Table \ref{table1} and
\ref{table2} show the result for some sets of typical values of the
parameters.
Table \ref{table1} shows the calculated values for the source on the
$y$-axis, i.e., $\beta_x=0$, and Table \ref{table2} for the source on
the $x$-axis, $\beta_y=0$. The maximal error is about $10^{-4}$, which is
the same order of magnitude as $O(e^2)$.
\begin{table}[htb]
\centering
\caption{Comparison of our approximate solutions with numerical ones
for $\beta_y=0$}
\label{table2}
\begin{tabular}{cccrrrrrr}
\hline\hline
$e$ & $\beta_x$ & $\beta_y$ &
(1) $x_{\rm num}$ & (2) $x_{\rm appr}$ &
$|(2)-(1)|$ &
(3) $y_{\rm num}$ & (4) $y_{\rm appr}$ &
$|(4)-(3)|$ \\
\hline
0.01 & 0.2 & 0.0 &
1.10902 & 1.10906 & $3.8\times 10^{-5}$ &
0.0 & 0.0 & 0.0 \\
& & &
-0.91097 & -0.91106 & $9.7\times 10^{-5}$ &
0.0 & 0.0 & 0.0 \\
& & &
0.00102 & 0.00100 & $2.0\times 10^{-5}$ &
0.10048 & 0.10050 & $1.7\times 10^{-5}$ \\
& & &
0.00102 & 0.00100 & $2.0\times 10^{-5}$ &
-0.10048 & -0.10050 & $1.7\times 10^{-5}$ \\
\hline
0.01 & 0.5 & 0.0 &
1.28372 & 1.28373 & $1.8\times 10^{-5}$ &
0.0 & 0.0 & 0.0 \\
& & &
-0.78855 & -0.78873 & $1.9\times 10^{-4}$&
0.0 & 0.0 & 0.0 \\
& & &
0.00254 & 0.00250 & $3.8\times 10^{-5}$ &
0.10035 & 0.10050 & $1.5\times 10^{-4}$ \\
& & &
0.00254 & 0.00250 & $3.8\times 10^{-5}$ &
-0.10035 & -0.10050 & $1.5\times 10^{-4}$ \\
\hline
0.02 & 0.2 & 0.0 &
1.11299 & 1.11314 & $1.5\times 10^{-4}$ &
0.0 & 0.0 & 0.0 \\
& & &
-0.91676 & -0.91714 & $3.8\times 10^{-4}$ &
0.0 & 0.0 & 0.0 \\
& & &
0.00208 & 0.00200 & $8.1\times 10^{-5}$&
0.14281 & 0.14284 & $2.5\times 10^{-5}$ \\
& & &
0.00208 & 0.00200 & $8.1\times 10^{-5}$&
-0.14281 & -0.14284 & $2.5\times 10^{-5}$ \\
\hline
0.02 & 0.5 & 0.0 &
1.28662 & 1.28669 & $7.0\times 10^{-5}$&
0.0 & 0.0 & 0.0 \\
& & &
-0.79598 & -0.79669 & $7.1\times 10^{-4}$ &
0.0 & 0.0 & 0.0 \\
& & &
0.00516 & 0.00500 & $1.6\times 10^{-4}$ &
0.14241 & 0.14284 & $4.2\times 10^{-4}$ \\
& & &
0.00516 & 0.00500 & $1.6\times 10^{-4}$ &
-0.14241 & -0.14284 & $4.2\times 10^{-4}$ \\
\hline
\end{tabular}
\end{table}
\section{Amplification factor}
Once we express the solutions for the image positions $(x,y)$ in terms
of the source position $(\beta_x,\beta_y)$, we can directly calculate
the amplification factor from the Jacobian
\begin{equation}
A^{\pm} = \left| \det\frac{\partial(x^{\pm},y^{\pm})}
{\partial(\beta_x,\beta_y)}\right|,
\end{equation}
where $A^{+}$ and $A^{-}$ represent the amplification factors for the
images with positive and negative parity, respectively.
The calculation is straightforward, and from Eqs.~(\ref{majorx}) and
(\ref{majory}) we obtain the following results up to the linear order
of $e$:
\begin{equation}
A^{+}=
A_0^{+}
\left( 1 - e \frac{4(\beta_x^2-\beta_y^2)
\left(\beta(\beta^2+6) -(\beta^2+4)\sqrt{\beta^2+4}\right)}
{\beta^3(\beta^2+4)\left(\sqrt{\beta^2+4}+\beta\right)^2}
\right),
\end{equation}
\begin{equation}
A^{-}=
A_0^{-}
\left( 1 - e \frac{4(\beta_x^2-\beta_y^2)
\left(\beta(\beta^2+6)+(\beta^2+4)\sqrt{\beta^2+4}\right)}
{\beta^3(\beta^2+4)\left(\sqrt{\beta^2+4}-\beta\right)^2}
\right),
\end{equation}
where
\begin{equation}
A_0^{\pm} \equiv
\frac{\left(\sqrt{\beta^2+4}\pm \beta\right)^2}{4\beta\sqrt{\beta^2+4}}
\end{equation}
represent the amplification factors for a point mass lens, namely
$e=0$ case.
We can also define the Pad\'e approximants for $A^{\pm}$ of order
$[0/1]$\footnote{
Generally speaking, the Pad\'e approximant of order $[m/n]$ is an
approximation of a function by a ratio of two polynomials of order $m$
and $n$. Then, for a function $A(x)$ whose Taylor expansion is given
by $A(x) \simeq A_0(1+ a_1 x+\cdots)$, the Pad\'e approximant of order
$[0/1]$ is simply $A_0/(1-a_1 x)$. At the same order, Pad\'e
approximations are usually
superior to Taylor expansions especially when functions contain
poles. }
with respect to $e$ as follows:
\begin{equation}\label{eq:ppade}
A_{\rm P}^{+}\equiv
A_0^{+}
\left( 1 + e \frac{4(\beta_x^2-\beta_y^2)
\left(\beta(\beta^2+6)-(\beta^2+4)\sqrt{\beta^2+4}\right)}
{\beta^3(\beta^2+4)\left(\sqrt{\beta^2+4}+\beta\right)^2}
\right)^{-1},
\end{equation}
\begin{equation}\label{eq:mpade}
A_{\rm P}^{-}\equiv
A_0^{-}
\left( 1 + e \frac{4(\beta_x^2-\beta_y^2)
\left(\beta(\beta^2+6)+(\beta^2+4)\sqrt{\beta^2+4}\right)}
{\beta^3(\beta^2+4)\left(\sqrt{\beta^2+4}-\beta\right)^2}
\right)^{-1} .
\end{equation}
\begin{table}[b]
\centering
\caption{Comparison of our approximate amplification factors with
numerical ones}
\label{table3}
\begin{tabular}{ccccrrcrc}
\hline\hline
$e$ & $\beta_x$ & $\beta_y$ & parity & (1) $A_{\rm num}$ &
(2) $A$ &$|(2)-(1)|/(1)$ & (3) $A_{\rm P}$ & $|(3)-(1)|/(1)$\\
\hline
0.01 & 0.0 & 0.2 & $+$ & 2.83741 &
2.82454 & 0.45\% & 2.83847 & 0.04\% \\
& & & $-$ & 2.37779 &
2.32454 & 2.24\% & 2.37166 & 0.26\% \\
\hline
0.01 & 0.0 & 0.5 & $+$ & 1.56603 &
1.56568 & 0.02\% & 1.56609 & 0.00\% \\
& & & $-$ & 0.65303 &
0.64568 & 1.13\% & 0.65116 & 0.29\% \\
\hline
0.02 & 0.0 & 0.2 & $+$ & 2.66004 &
2.61173 & 1.82\% & 2.66404 & 0.15\% \\
& & & $-$ & 2.87313 &
2.61173 & 9.10\% & 2.83724 & 1.25\% \\
\hline
0.02 & 0.0 & 0.5 & $+$ & 1.54134 &
1.53994 & 0.09\% & 1.54156 & 0.01\% \\
& & & $-$ & 0.73395 &
0.69994 & 4.63\% & 0.72434 & 1.31\% \\
\hline
0.01 & 0.2 & 0.0 & $+$ & 3.26482 &
3.25015 & 0.45\% & 3.26619 & 0.04\% \\
& & & $-$ & 1.78910 &
1.75015 & 2.18\% & 1.78564 & 0.19\% \\
\hline
0.01 & 0.5 & 0.0 & $+$ & 1.61749 &
1.61714 & 0.02\% & 1.61758 & 0.00\% \\
& & & $-$ & 0.54292 &
0.53714 & 1.06\% & 0.54170 & 0.22\% \\
\hline
0.02 & 0.2 & 0.0 & $+$ & 3.52534 &
3.46296 & 1.77\% & 3.53232 & 0.20\% \\
& & & $-$ & 1.59981 &
1.46296 & 8.55\% & 1.58928 & 0.66\% \\
\hline
0.02 & 0.5 & 0.0 & $+$ & 1.64434 &
1.64288 & 0.09\% & 1.64460 & 0.02\% \\
& & & $-$ & 0.50371 &
0.48288 & 4.14\% & 0.49971 & 0.80\% \\
\hline
\end{tabular}
\end{table}
As shown in Table \ref{table3}, the Pad\'e approximants
Eqs.~(\ref{eq:ppade}) and (\ref{eq:mpade}) generally give the
approximate amplifications with better accuracy of about $1\%$ or
less.
Hereafter, we use Eqs.~(\ref{eq:ppade}) and (\ref{eq:mpade}) as our
approximate formulae for the amplification factors.
\section{Changes in
the image properties}
For a point mass lens model, we have the following ``universal''
relations for the image properties. First, the sum of the two
image positions is
\begin{equation}\label{univ1}
x^{+}_0 + x^{-}_0 = \beta_x, \quad y^{+}_0 + y^{-}_0 = \beta_y.
\end{equation}
Namely, the sum of the image positions is always equal to the original
source position.
Second, the difference of the amplification factors is always unity:
\begin{equation}\label{univ2}
A_0^{\rm diff} \equiv A^{+}_0 - A^{-}_0 = 1.
\end{equation}
This also means, for a point mass lens, the image with positive parity
is always brighter than that with negative parity.
Just for reference, we also show that the image separation and the total
amplification are expressed in the following way:
\begin{equation}
\Delta x_0\equiv x^{+}_0 - x^{-}_0 = \frac{\sqrt{\beta^2+4}}{\beta} \beta_x, \quad
\Delta y_0\equiv y^{+}_0 - y^{-}_0 = \frac{\sqrt{\beta^2+4}}{\beta} \beta_y,
\end{equation}
\begin{equation}
A_0^{\rm tot}\equiv A^{+}_0 + A^{-}_0
= \frac{\beta^2+2}{\beta\sqrt{\beta^2+4}}.
\end{equation}
In this section, we investigate how the quadrupole moment changes the
above image properties.
From Eqs.~(\ref{majorx}) and (\ref{majory}), we obtain the sum as
\begin{equation}
x^{+} + x^{-} = \beta_x - e \left(1+\frac{4\beta_y^2}{\beta^4}\right)
\beta_x,\quad
y^{+} + y^{-} = \beta_y + e \left(1+\frac{4\beta_x^2}{\beta^4}\right) \beta_y,
\end{equation}
The change due to $e$ is getting larger as $\propto
4e\beta^{-2}$ for $|\beta| \ll 1$.
Therefore, even if $|e|\ll 1$, the ``universal'' relation
Eq.~(\ref{univ1}) may be significantly broken if $|\beta| \ll 1$.
On the contrary, the image separation is
\begin{equation}
\Delta x \equiv x^{+} - x^{-} =
\Delta x_0 \left\{
1 - \left(
\frac{2(\beta_x^2-\beta_y^2)}{\beta^2(\beta^2+4)}-1
\right) e \right\},
\end{equation}
\begin{equation}
\Delta y \equiv y^{+} - y^{-} =
\Delta y_0 \left\{
1 - \left(
\frac{2(\beta_x^2-\beta_y^2)}{\beta^2(\beta^2+4)}+1
\right) e \right\}.
\end{equation}
It is evident that, even if $|\beta| \ll 1$, the change due to $e$ in
the image separation is $\sim e\times O(1)$. Therefore, we may
conclude that the change due to $e$ in the image separation is small for $|e|\ll
1$.
Similar situation occurs in the analysis of the image amplification.
From Eqs.~(\ref{eq:ppade}) and (\ref{eq:mpade}), the amplification
difference for $|\beta| \ll 1$ is
\begin{equation}\label{eq:diff}
A^{\rm diff} \equiv A^{+}_P - A^{-}_P \simeq \frac{1+\frac{2}{\beta^2}e'}
{1+ \frac{7}{2}e' - \left(\frac{2}{\beta}e'\right)^2},
\end{equation}
where
\begin{equation}
e' = \left( \left(\frac{\beta_x}{\beta}\right)^2
-\left(\frac{\beta_y}{\beta}\right)^2
\right) e.
\end{equation}
The change due to $e$ is getting larger as $\propto
2e\beta^{-2}$ for $|\beta| \ll 1$.
Therefore, even if $|e|\ll 1$, the ``universal'' relation
Eq.~(\ref{univ2}) may be significantly broken if $|\beta| \ll 1$.
As is already shown numerically in Table \ref{table3}, there is even a case
of $A^{\rm diff} < 0$ for $e=0.02, \beta_x=0.0, \beta_y=0.2$.
The total amplification is
\begin{equation}\label{eq:5-10}
A^{\rm tot}\equiv A^{+}_P + A^{-}_P \simeq \frac{A^{\rm tot}_0}
{1+\frac{3}{2}e' - \left(\frac{2}{\beta}e'\right)^2}.
\end{equation}
Compared to $A^{\rm diff}$, the change due to $e$ is relatively mild
for $|\beta| \ll 1$.
We give some comments on the range of the parameters $e$ and
$\beta$. Since we have started from the linear perturbation with
respect to $e$, the range
of validity is $0\leq e \ll 1$. If we assume additionally that the
maximal error in the amplification factors should be less than, say, $1\%$,
the range of validity may be $0 \leq e \lesssim 0.01$, which we can guess
from Table \ref{table3}.
The ranges of $\beta_x$ and $\beta_y$ are more important.
From Eqs.~(\ref{eq:diff}) and (\ref{eq:5-10}), we can obtain the constraint
$|e'| \lesssim \beta^2$ and $|e'| \ll \beta$.
If not, then the term of $O(e^2)$ in the denominators of
Eqs.~(\ref{eq:diff}) and (\ref{eq:5-10}) can make a dominant
contribution over the linear term, which is outside of the validity of
linear perturbation theory.
\section{Summary}
In this paper, we have investigated the gravitational lens effect
caused by a non-spherical, compact object.
The non-spherical property of the gravitational potential is expressed
by the quadrupole moment.
Eqs.~(\ref{eq:leq1}) and (\ref{eq:leq2}) are the lens equations for a
compact lens object with a quadrupole moment. In these lens equations,
the effect of the non-spherical contribution is expressed by $e$, the
eigenvalue of the normalized quadrupole moment tensor $\tilde Q_{ij}$.
Under the assumption that the quadrupole moment $e$ is small, we have
perturbatively solved the lens equation and obtained the solutions for
the image positions. Eqs.~(\ref{majorx}) and (\ref{majory}) represent
the perturbative solutions for the positions of the ``major'' images,
which are reduced to the solutions for a point mass lens model in the
limit $e\rightarrow 0$. We have also found two new perturbative
solutions for the ``minor'' images, which are given in
Eq.~(\ref{minor}). The minor images are very dim, always appear very
close to the $y$-axis, near $(0,\pm \sqrt{e})$, and vanish in the
limit $e\rightarrow 0$. The accuracy of our approximate solutions is
numerically checked and summarized in Tables \ref{table1} and
\ref{table2} for some sets of parameters. It is shown that the
maximal error in the image position is about $10^{-4}$, which is the
same order of magnitude as $O(e^2)$.
Then, we have calculated the amplification factors for the major
images from the Jacobian and obtained the Pad\'e approximants of them
in Eqs.~(\ref{eq:ppade}) and (\ref{eq:mpade}). The accuracy of our
approximate formula for the amplification factor is numerically
estimated and summarized in Table \ref{table3} for some sets of
parameters. It is shown that the typical relative error is $1\%$ or
less.
For a point mass lens model, we know the simple ``universal''
relations for the image properties, Eqs.~(\ref{univ1}) and
(\ref{univ2}). We have investigated how the quadrupole contribution
$e$
changes such ``universal'' relations.
We have found that the change due to a non-zero $e$ in the image
separation is $\sim e\times O(1)$.
However, the difference of the amplifications $A^{\rm diff} =
A^{+}-A^{-}$, which is always unity in the case of a point mass lens,
may be significantly changed due to the quadrupole contribution, which
is shown in Eq.~(\ref{eq:diff}).
Mao and Schneider\cite{ms} discussed the ``anomalous'' flux ratio between the
images caused by a lens galaxy.
The lens system they discussed was QSO B 1422 + 231. Whereas the
well-known
``simple'' lens models, such as a singular isothermal ellipsoid, could
fit the observed image positions very accurately, they all failed to
obtain the observed flux ratios. The possibility they argued is that the
discrepancy between the observed and model-predicted flux ratios is due
to substructure or perturbation in the simple lens models, which only
slightly
changes the image positions but significantly affects the flux
ratios.
Since the lens model in this paper is different from
that of Mao and Schneider,\cite{ms}
we have no intention of directly applying our results to their lens
systems. Still, it should be noted that
the results in our paper show some formal similarity.
It indicates that even a tiny non-spherical distortion of the lens
potential may also cause significant amount of flux
anomalies
in the
lensed images, whereas it only slightly changes the image positions.
|
\section{Introduction}
\label{S:intro}
This paper concerns the Monge transportation problem in geodesic spaces, i.e. metric spaces with a geodesic structure.
Given two Borel probability measure $\mu,\nu \in \mathcal{P}(X)$, where $(X,d)$ is a Polish space, we study the minimization of the functional
\[
\mathcal{I}(T) = \int d_{L}(x,T(x)) \mu(dy)
\]
where $T$ varies over all Borel maps $T:X \to X$ such that $T_{\sharp}\mu = \nu$ and $d_{L}$ is a Borel distance that makes $(X,d_{L})$ a non branching geodesic space.
Before giving an overview of the paper and of the existence result, we recall which are the main results concerning the Monge problem.
In the original formulation given by Monge in 1781 the problem was settled in $\mathbb{R}^{n}$,
with the cost given by the Euclidean norm and the measures $\mu, \nu$ were 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}
solved 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 have been 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}.
Afterwards M. Feldman and R.J. McCann \cite{feldcann:mani} extended 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,
has been 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}.
\subsection{Overview of the paper}
\label{Ss:over}
The presence of $1$-dimensional sets (the geodesics) along which the cost is linear is a strong degeneracy for transport problems. This degeneracy is equivalent to the following problem in $\mathbb{R}$: if $\mu$ is concentrated on $(-\infty,0]$, and $\nu$ is concentrated on $[0,+\infty)$, then every transference plan is optimal for the $1$-dimensional distance cost $|\cdot|$. In fact, every $\pi \in \Pi(\mu,\nu)$ is supported on the set $(-\infty,0] \times [0,+\infty)$, on which $|x-y| = y-x$ and thus
\[
\int |x-y| \pi(dxdy) = - \int x \mu(dx) + \int y \nu(dy).
\]
Nevertheless, for this easy case an explicit map $T : \mathbb{R} \to \mathbb{R}$ can be constructed if $\mu$ is continuous (i.e. without atoms): the easiest choice is the monotone map, a minimizer of the quadratic cost $|\cdot|^2$.
The approach suggested by the above simple case is the following:
\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}
While the last point can be seen as an application of selection principles in Polish spaces, the first two points are more subtle.
The geodesics used by a given transference plan $\pi$ to transport mass can be obtained from a set $\Gamma$ on which $\pi$ is concentrated. If $\pi$ wants to be a minimizer, then it certainly chooses the shortest paths: however the metric space can be branching, i.e. geodesics can bifurcate.
In this paper we assume that the space is non branching.
Under this assumption, a cyclically monotone plan $\pi$
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 non branching assumption and the cyclical monotonicity of $\Gamma$ imply that the geodesics used by $\pi$ are a partition on $\mathcal T$.
In general in $a$
there are points from which more than geodesic starts and in $b$ there are points in which more than one geodesic ends, hence
being on a geodesic can't be an equivalence relation on the set $a\cup b$.
For example one can think to the unit circle with $\mu = \delta_0$ and $\nu = \delta_\pi$.
We note here that $\pi$ gives also a direction along each component of $R$, as the one dimensional example above shows.
Even if we have a natural partition $R$ in $\mathcal T$ and $\mu(a \cup b) = 0$, 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 $R$. In this case, one can write
\[
m := f_\sharp \mu, \quad \mu = \int \mu_y m(dy), \quad \mu_y(f^{-1}(y)) = 1,
\]
i.e. the conditional probabilities $\mu_y$ are concentrated on the counterimages $f^{-1}(y)$ (which are single geodesics). We can 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 $|\cdot|$, the length on the geodesic. At this point we can study the problem of the regularity of the conditional probabilities $\mu_y$.
The fact that there exists a strongly consistent disintegration is a property of the geodesics of the metric space. In the setting considered in this paper, $(X,d_L)$ is a non branching geodesic space, not necessarily Polish. To assure that standard measure theory can be used, there exists a second distance $d$ on $X$ which makes $(X,d)$ Polish, and $d_L$ is a Borel function on $X \times X$ with the metric $d \times d$.
Note that we do not require $d_L$ to be l.s.c., so the existence of an optimal plan $\pi$ is not assured, but we consider a $d_L$-cyclically monotone transference plan $\pi$.
It is worth notice that we do not use the existence of optimal potentials $(\phi,\psi)$, as well as the optimality of $\pi$.
Thus, let $\pi$ be a $d_L$-cyclically monotone transference plan. The strong consistency of the disintegration of $\mu$ along the geodesic used by $\pi$
is a consequence of the topological properties of the geodesics of $d_L$ considered as curves in $(X,d)$: in fact we require that they are $d$-continuous and locally compact. Under this assumption, on $\mathcal T$ (the transport set without end points) it is possible to disintegrate $\mu$. Moreover, a natural operation on sets can be considered: the translation along geodesics. If $A$ is a subset of $\mathcal T$, we denote by $A_t$ the set translated by $t$ in the direction determined by $\pi$.
It turns out that the fact that $\mu(a \cup b) = 0$ and the measures $\mu_y$ are continuous depends on how the function $t \mapsto \mu(A_t)$ behaves. We can now state the main result.
\begin{theorem}[Lemma \ref{L:puntini} and Proposition \ref{P:nonatoms}]
\label{T:-1}
If $\sharp \{t > 0: \mu(A_t) > 0\}$ is uncountable for all $A$ Borel such that $\mu(A) > 0$, then $\mu(a \cup b) = 0$ and the conditional probabilities $\mu_y$ are continuous.
\end{theorem}
This is sufficient to solve the Monge problem, i.e. to find a transport map which has the same cost as $\pi$. A second result concerns a stronger regularity assumption.
\begin{theorem}[Theorem \ref{teo:a.c.}]
\label{T:1}
Assume that $\mathcal L^1(\{t > 0: \mu(A_t) > 0\}) > 0$ for all $A$ Borel such that $\mu(A) > 0$. Then $\mu(a \cup b) = 0$ and $\mu_y$ is a.c. w.r.t. the $1$-dimensional Hausdorff measure $\mathcal H^1_{d_L}$ induced by $d_L$.
\end{theorem}
The assumption of the above theorem and the assumption $d_L \geq d$ allows to define a current in $(X,d)$ which represents the vector field corresponding to the translation $A \mapsto A_t$, and moreover to solve the equation
\[
\partial U = \mu - \nu
\]
is the sense of current in metric space.
The final results of the paper are the stability of these conditions under Measure-Gromov-Hausdorff like convergence of structures $(X_n,d_n,d_{L,n},\pi_{n})$. The conclusion is that a sort of uniform integrability condition on the conditional probability w.r.t. $\mathcal H^1_{d_{L,n}}$ passes to the limit, so that one can verify by approximation if Theorem \ref{T:1} holds.
In the case $d = d_L$, considering a reference measure $\eta \in \mathcal{P}(X)$ such that
$(X,d,\eta)$ is a geodesic measure space satisfying the $MCP(K,N)$ for $K\in \mathbb{R}$ and $N\geq1$,
the application of the above results together with $MCP$-condition implies that Assumption \ref{A:NDE} holds for
$\eta$ w.r.t. the optimal flow induced by any $d$-monotone plan $\pi\in \Pi(\mu,\nu)$.
Hence if $\mu \ll \eta$,
the existence of a minimizer for the Monge minimization problem with marginal $\mu$ and $\nu$ follows.
To conclude this introduction, we observe that it is probably possible to extend these results to the case where $-d_L$ is a Souslin function on $(X \times X,d \times d)$:
this means that $d_{L}^{-1}(-\infty,t)$ is an analytic set in the sense of Souslin.
The interested reader can refer for example to the analysis of \cite{biacar:cmono}.
\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 in Section \ref{ss:Metric} we define the geodesic structure $(X,d,d_L)$ which is studied in this paper. Finally, Section \ref{Ss:General Facts} recalls some fundamental results in optimal transportation theory.
The next three sections are the key ones.
Section \ref{S:Optimal} shows how 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 (Lemma \ref{L:analGR} and Proposition \ref{P:equiv}): $xGy$ 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 (Proposition \ref{P:equiv}), we can assume that the set of final points $b$ can be taken $\mu$-negligible (Lemma \ref{L:finini0}), and in two final remarks we study what happens in the case more regularity on the cost $d_L$ is assumed, Remark \ref{R:lsccase1} and Remark \ref{R:TR}.\\
Notice that in the case $d=d_{L}$ the existence of a Lipschitz potential $\varphi$, one can take
\[
\Gamma = G = \Big\{ (x,y) : \varphi(x)-\varphi(y)=d(x,y) \Big\}.
\]
Thus the main result of this section is that these sets can be defined even if the potential does not exist.
Section \ref{S:partition} proves that the continuity and local compactness of geodesics imply that the disintegration induced by $R$ on $\mathcal T$ is strongly consistent (Proposition \ref{P:sicogrF}): as Example \ref{Ex:nonsection}
shows, the strong consistency of the disintegration is a non trivial property of the metric spaces we are considering.\\
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:regurlr} we prove Theorem \ref{T:-1} and Theorem \ref{T:1}. We first introduce the operation $A \mapsto A_t$, the translation along geodesics \eqref{E:At}, and show that $t \mapsto \mu(A_t)$ is a Souslin function if $A$ is analytic (Lemma \ref{L:measumuAt}). \\
Next we show that under the assumption
\[
\mu(A) > 0 \quad \Longrightarrow \quad \sharp \big\{ t > 0: \mu(A_t) > 0 \big\} > \aleph_0
\]
the set of initial points $a$ is $\mu$-negligible (Lemma \ref{L:puntini}) and the conditional probabilities $\mu_y$ are continuous. \\
Finally, we show that under the stronger assumption
\begin{equation}
\label{E:intro0}
\mu(A) > 0 \quad \Longrightarrow \quad \int_{\mathbb{R}^+} \mu(A_t) dt > 0,
\end{equation}
the conditional probabilities $\mu_y$ are a.c. w.r.t. $\mathcal H^1_{d_L}$ (Theorem \ref{teo:a.c.}). A final result shows that actually Condition \eqref{E:intro0} yields that $t \mapsto \mu(A_t)$ has more regularity than just integrability (Proposition \ref{P:peloso}) it is in fact continuous
After the above results, the solution of the Monge problem is routine, and it is done in Theorem \ref{T:mongeff} of Section \ref{S:Solution}.
Under Condition \ref{E:intro0} and $d \leq d_L$, in Section \ref{S:div} we give a dynamic interpretation to the transport along geodesics. In Definition \ref{D:dotgamma} we define the current $\dot g$ in $(X,d)$, which represents the flow induced by the transference plan $\pi$. Not much can be said of this flow, unless some regularity assumptions are considered. These assumptions are the natural extensions of properties of transportation problems in finite dimensional spaces. \\
If there exists a background measure $\eta$ whose disintegration along geodesics satisfies
\[
\eta = \int q_y \mathcal{H}^1_{d_L} m(dy), \quad q_y \in \text{\rm BV}, \ \int \text{\rm Tot.Var.}(q_y) m(dy) < +\infty,
\]
then $\dot g$ is a normal current, i.e. its boundary is a bounded measure on $X$ (Lemma \ref{L:normalcurr}). \\
We can also consider the problem $\partial U = \mu - \nu$ in the sense of currents: Proposition \ref{P:bidual} gives a solution, and in the case $q_y(t) > 0$ for $\mathcal H^1_{d_L}$-a.e. $t$ we can write represent $U = \rho \dot g$, i.e. the flow $\dot g$ multiplied by a scalar density $\rho$ (Corollary \ref{C:regudual}).
In Section \ref{S:limite} we address the stability of the assumptions under Measure-Gromov-Hausdorff-like convergence of
structures $(X_n,d_n,d_{L,n},\pi_{n})$.
Under a uniform integrability condition of $\mu_{y,n}$ w.r.t. $\mathcal H^1_{d_{L,n}}$ and
a uniform bound on the $\pi_{n}$ transportation cost
(Assumption \ref{A:equi} of Section \ref{Ss:GHC}), we show that the marginal $\mu$ can be represented
as the image of a measure $r m \otimes \mathcal L^1$ by a Borel function $h : \mathcal T \times \mathbb{R} \to \mathcal T_e$,
with $r \in L^1(m \otimes \mathcal L^1)$ (Proposition \ref{P:graphnn}).
The key feature of $h$ is that $t \mapsto h(y,t)$ is a geodesic of $\mathcal T$ for $m$-a.e. $y \in \mathcal T$. \\
Thus while $h(0,\mathcal T)$ is not a cross section for $R$ (in that case we would have finished the proof), in Proposition \ref{P:finalapr} we show which conditions on $h$ imply that $\mu$ can be disintegrated with a.c. conditional probabilities, and we verify that this is our case in Theorem \ref{T:aproxxgg}. \\
In two remarks we suggest how to pass also uniform estimates on the disintegration on $(X_n,d_n,d_{L,n})$ to the transference problem in $(X,d,d_L)$ (Remark \ref{R:morereg} and Remark \ref{R:reguadd}).
In Section \ref{S:mcp} we consider an application of the results obtained in the previous sections.
We assume $d=d_{L}$ and the existence of background probability measure $\eta$ such that $(X,d,\eta)$
satisfies $MCP(K,N)$ (Definition \ref{D:mcp}).
In this framework we prove that for any $d$-cyclically monotone transference plan $\pi$, $\eta$ admits a disintegration along the geodesics
used by $\pi$ with marginal probabilities absolutely continuous w.r.t. $\mathcal{H}^{1}$ (Theorem \ref{T:regularity}).
This implies directly (Corollary \ref{C:conclu}) that if $\mu\ll \eta$ the Monge minimization problem with marginals $\mu$ and $\nu$
admits a solution.
The final result of the section (Lemma \ref{L:mhdregul}) shows that we can solve the dynamical problem $\partial U = \mu - \nu$
with $U=\rho \dot g$, and if the support of $\mu$ and $\nu$ are disjoint $U$ is a normal current.
The last section contains two important examples. In Example \ref{Ex:nonsection} we show that if the geodesics are not locally compact, then in general the disintegration along transport rays is not strongly supported. In Example \ref{Ex:nooptir} we show that under our assumptions the $c$-monotonicity is not sufficient for optimality.
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 $f^{-1}(t,+\infty] \in \Sigma^{1}_{1}$.
\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}
\begin{proof}
If $\mu \in \mathcal{M}(X)$, then $f_\sharp \mu \in \mathcal{M}(Y)$, so for $U \subset Y$ universally measurable there exist Borel sets $B$, $B'$ such that $B \subset U \subset B'$ and
\[
0 = (f_\sharp \mu)(B' \setminus B) = \mu \big( f^{-1}(B') \setminus f^{-1}(B) \big).
\]
Since $f^{-1}(B), f^{-1}(B') \subset X$ are universally measurable, there exists Borel sets $C$, $C'$ such that
\[
C \subset f^{-1}(B) \subset f^{-1}(U) \subset f^{-1}(B') \subset C'
\]
and $\mu(C' \setminus C) = 0$. The conclusion follows.
\end{proof}
\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{Metric setting}
\label{ss:Metric}
In this section we refer to \cite{burago}.
\begin{definition}
\label{D:lengthstr}
A \emph{length structure} on a topological space $X$ is a class $\mathtt{A}$ of admissible paths, which is a subset of all continuous paths in X, together with a map $L: \mathtt{A} \to [0,+\infty]$: the map $L$ is called \emph{length of path}. The class $\mathtt{A}$ satisfies the following assumptions:
\begin{description}
\item[closure under restrictions] if $\gamma : [a,b] \to X$ is admissible and $a \leq c\leq d \leq b$, then $\gamma \llcorner_{[c,d]}$ is also admissible.
\item[closure under concatenations of paths] if $\gamma : [a,b] \to X$ is such that its restrictions $\gamma_1, \gamma_2$ to $[a,c]$ and $[c,b]$ are both admissible, then so is $\gamma$.
\item[closure under admissible reparametrizations] for an admissible path $\gamma : [a,b] \to X$ and a for $\varphi: [c,d]\to[a,b]$, $\varphi \in B$, with $B$ class of admissible homeomorphisms that includes the linear one, the composition $\gamma(\varphi(t))$ is also admissible.
\end{description}
The map $L$ satisfies the following properties:
\begin{description}
\item[additivity] $L(\gamma \llcorner_{[a,b]}) = L(\gamma \llcorner_{[a,c]}) + L(\gamma \llcorner_{[c,b]})$ for any $c\in [a,b]$.
\item[continuity] $L(\gamma \llcorner_{[a,t]})$ is a continuous function of $t$.
\item[invariance] The length is invariant under admissible reparametrizations.
\item[topology] Length structure agrees with the topology of $X$ in the following sense: for a neighborhood $U_x$ of a point $x \in X$, the length of paths connecting $x$ with points of the complement of $U_x$ is separated from zero:
\[
\inf \big\{ L(\gamma) : \gamma(a)=x, \gamma(b) \in X\setminus U_{x} \big\} >0.
\]
\end{description}
\end{definition}
Given a length structure, we can define a distance
\[
d_L(x,y) = \inf \Big\{ L(\gamma): \gamma:[a,b]\to X, \gamma \in \mathtt{A}, \gamma(a)=x, \gamma(b)=y \Big\},
\]
that makes $(X,d_{L})$ a metric space (allowing $d_{L}$ to be $+\infty$). The metric $d_{L}$ is called \emph{intrinsic}.
It follows from Proposition 2.5.9 of \cite{burago} that every admissible curve of finite length admits a constant speed parametrization, i.e.
$\gamma$ defined on $[0,1]$ and $L(\gamma\llcorner[t,t'])= v (t'-t)$, with $v$ velocity.
\begin{definition}
A length structure is said to be \emph{complete} if for every two points $x,y$ there exists an admissible path joining them whose length $L(\gamma)$ is equal to $d_{L}(x,y)$.
\end{definition}
In other words, a length structure is complete if there exists a shortest path between two points.
Intrinsic metrics associated with complete length structure are said to be \emph{strictly intrinsic}. The metric space $(X,d_L)$ with $d_L$ strictly intrinsic is called a \emph{geodesic space}. A curve whose length equals the distance between its end points is called \emph{geodesic}.
\begin{definition}\label{D:strettaconv}
Let $(X,d_L)$ be a metric space. The distance $d_L$ is said to be \emph{strictly convex} if, for all $r \geq 0$, $d_L(x,y) = r/2$ implies that
\[
\{ z : d_L(x,z) = r \} \cap \{ z : d_L(y,z) = r/2 \}
\]
is a singleton.
\end{definition}
The definition can be restated in geodesics spaces as: geodesics cannot bifurcate in the interior, i.e. \emph{the geodesic space $(X,d_L)$ is not branching}.
An equivalent requirement is that if $\gamma_1 \not=\gamma_2$ and $\gamma_1(0) = \gamma_2(0)$, $\gamma_1(1) = \gamma_2(1)$, then $\gamma_1((0,1)) \cap \gamma_2((0,1)) = \emptyset$ and such geodesics do not admit a geodesic extension i.e. they are not a part of a longer geodesic.
From now on we assume the following: \label{P:assumpDL}
\medskip
\begin{enumerate}
\item $(X,d)$ Polish space;
\item $d_L : X \times X \to [0,+\infty]$ Borel distance;
\item $(X,d_L)$ is a non-branching geodesic space;
\item geodesics are continuous w.r.t. $d$;
\item \label{Cond:XdL5}geodesics are locally compact in $(X,d)$: if $\gamma$ is a geodesic for $(X,d_L)$, then for each $x \in \gamma$ there exists $r$ such that $\gamma^{-1}(\bar B_r(x))$ is compact in $\mathbb{R}$.
\end{enumerate}
\medskip
Since we have two metric structures on $X$, we denote the quantities relating to $d_L$ with the subscript $L$: for example
\[
B_r(x) = \big\{ y : d(x,y) < r \big\}, \quad B_{r,L}(x) = \big\{ y : d_L(x,y) < r \big\}.
\]
In particular we will use the notation
\[
D_L(x) = \big\{ y : d_L(x,y) < + \infty \big\},
\]
$(\mathcal{K},d_H)$ for the compact sets of $(X,d)$ with the Hausdorff distance $d_H$ and $(\mathcal{K}_L,d_{H,L})$ for the compact sets of $(X,d_L)$ with the Hausdorff distance $d_{H,L}$. We recall that $(\mathcal{K},d_H)$ is Polish.
We write
\begin{equation}
\label{E:gammarre}
\gamma_{[x,y]} := \Big\{ \gamma \in \textrm{Lip}_{d_L}([0,1];X): \gamma(0) = x, \gamma(1)=y, L(\gamma) = d_L(x,y) \Big\}.
\end{equation}
With a slight abuse of notation, we will write
\begin{equation}
\label{E:opeclgeo}
\gamma_{(x,y)} = \bigcup_{\gamma \in \gamma_{[x,y]}} \gamma((0,1)), \quad \gamma_{[x,y]} = \bigcup_{\gamma \in \gamma_{[x,y]}} \gamma([0,1]).
\end{equation}
We will also use the following definition.
\begin{definition}
\label{D:geoconvx}
We say that $A \subset X$ is \emph{geodesically convex} if for all $x,y \in A$ the minimizing geodesic $\gamma_{[x,y]}$ between $x$ and $y$ is contained in $A$:
\[
\Big\{ \gamma((0,1)) : \gamma(0) = x, \gamma(1) = y, L(\gamma) = d(x,y), x,y \in A \Big\} \subset A.
\]
\end{definition}
\begin{lemma}
\label{L:measclos}
If $A$ is analytic in $(X,d)$, then $\{x : d_L(A,x) < \epsilon \}$ is analytic for all $\epsilon > 0$.
\end{lemma}
\begin{proof}
Observe that
\[
\big\{ x : d_L(A,x) < \epsilon \big\} = P_1 \Big( X \times A \cap \big\{ (x,y) :d_L(x,y) < \epsilon \big\} \Big),
\]
so that the conclusion follows from the invariance of the class $\Sigma^1_1$ w.r.t. projections.
\end{proof}
In particular, $\overline{A}^{_{d_L}}$, the closure of $A$ w.r.t. $d_L$, is analytic if $A$ is analytic.
\begin{remark}
\label{R:compact}
During the paper, whenever more regularity is required, we will assume also the following hypothesis:
\begin{itemize}
\item[(2')] $d_L : X \times X \to [0,+\infty]$ l.s.c. distance,
\item[(4')] $d_L(x,y) \geq d(x,y)$,
\item[(5')] \label{Point:globalcomp} $\cup_{x \in K_1, y \in K_2} \gamma_{[x,y]}$ is $d$-compact if $K_1$, $K_2$ are $d$-compact, $d_L \llcorner_{K_1 \times K_2}$ uniformly bounded.
\end{itemize}
A simple computation shows that $d_L(x,y) \geq d(x,y)$ implies the following
\begin{enumerate}
\item $d_L$-compact sets are $d$-compact;
\item $d$-Lipschitz functions are $d_{L}$-Lipschitz with the same constant.
\end{enumerate}
\end{remark}
An application of Theorem \ref{T:vanneuma}, in the setting of Remark \ref{R:compact}, gives a Borel function which selects a single geodesic $\gamma\in \gamma_{[x,y]}$
for any couple $(x,y)$.
\begin{lemma}\label{L:geodetiche}
Assume that $d_{L}$ is l.s.c.. Then
there exists a Borel function $\varUpsilon: X \times X \to \textrm{\rm Lip}_{d}([0,1],X)$ such that
up to reparametrization $\varUpsilon(x,y)\in \gamma_{[x,y]}$.
\end{lemma}
\begin{proof}
Let
\begin{equation*}
\begin{array}{ccccc}
F& : & X\times X& \to & \textrm{Lip}_{d}([0,1],X) \\
&& (x,y)& \mapsto & \gamma_{[x,y]}
\end{array}
\end{equation*}
with $\textrm{Lip}_{d}([0,1],X)$ endowed with the uniform topology and $\gamma_{[x,y]}$ defined in \eqref{E:gammarre}.
The result follows by Theorem \ref{T:vanneuma} observing that $\textrm{graph}(F)$ is the set
\[
\Big\{ (x,y,\gamma)\in X\times X \times \textrm{Lip}_{d}([0,1],X), L(\gamma)= d_{L}(x,y) \Big\}.
\]
which is Borel by the l.s.c. of the map $\gamma \mapsto L(\gamma)$, and this is implied by the l.s.c. of $d_{L}$.
\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}.
\section{Optimal transportation in geodesic spaces}
\label{S:Optimal}
Let $\mu, \nu \in \mathcal{P}(X)$ and consider the transportation problem with cost $c(x,y)= d_L(x,y)$, 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}
In this section, 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.
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 $0$.
\begin{lemma}
\label{L:gGamma}
The following holds:
\begin{enumerate}
\item $\Gamma \subset \Gamma' \subset \{d_L(x,y) < +\infty\}$;
\item if $\Gamma$ is analytic, so is $\Gamma'$;
\item if $\Gamma$ is $d_L$-cyclically monotone, so is $\Gamma'$.
\end{enumerate}
\end{lemma}
\begin{proof}
For the first point, set $I=0$ and $(w_{n,0},z_{n,0}) = (x,y)$ for the first inclusion. If $d_L(x,y) = +\infty$, then $(x,y) \notin \Gamma$ and all finite set of points in $\Gamma$ are bounded.
For the second point, observe that
\begin{align*}
\Gamma' =&~ \bigcup_{I \in \mathbb{N}_0} P_{12} (A_I) \crcr
=&~ \bigcup_{I \in \mathbb{N}_0} P_{12} \bigg( \prod_{i=0}^I \Gamma \cap \bigg\{ \prod_{i=1}^I (w_i,z_i) : \sum_{i=0}^I d_L(w_{i+1},z_i) - d_L(w_i,z_i) = 0, w_{I+1} = w_0 \bigg\} \bigg).
\end{align*}
For each $I \in \mathbb{N}_0$, since $d_L$ is Borel, it follows that
\begin{align*}
\bigg\{ \prod_{i=1}^I (w_i,z_i) : \sum_{i=0}^I d_L(w_{i+1},z_i) - d_L(w_i,z_i) = 0, w_{I+1} = w_0 \bigg\}
\end{align*}
is Borel in $\prod_{i=0}^I (X \times X)$, so that for $\Gamma$ analytic each set $A_{n,I}$ is analytic. Hence $P_{12}(A_I)$ is analytic, and since the class $\Sigma^1_1$ is closed under countable unions and intersections it follows that $\Gamma'$ is analytic.
For the third point, observe that for all $(x_j,y_j) \in \Gamma'$, $j=0,\dots,J$, there are $(w_{j,i},z_{j,i}) \in \Gamma$, $i = 0,\dots,I_j$, such that
\begin{align*}
d_L(x_j,y_j) + \sum_{i=0}^{I_j-1} d_L(w_{j,i+1},z_{j,i}) - \sum_{i=0}^{I_j} d_L(w_{j,i},z_{j,i}) = 0.
\end{align*}
Hence we can write for $x_{J+1} = x_0$, $w_{j,I_j+1} = w_{j+1,0}$, $w_{J+1,0} = w_{0,0}$
\begin{align*}
\sum_{j=0}^J d_L(x_{j+1},y_j) - d_L(x_j,y_j) =&~ \sum_{j=0}^J \sum_{i=0}^{I_j} d_L(w_{j,i+1},z_{j,i}) - d_L(w_{j,i},z_{j,i}) \geq 0,
\end{align*}
using the $d_L$-cyclical monotonicity of $\Gamma$.
\end{proof}
\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}
\begin{lemma}
\label{L:analGR}
The following holds:
\begin{enumerate}
\item $G$ is $d_L$-cyclically monotone;
\item $\Gamma' \subset G \subset \{d_L(x,y) < +\infty\}$;
\item the sets $G$, $R := G \cup G^{-1}$ are analytic.
\end{enumerate}
\end{lemma}
\begin{proof}
The second point follows by the definition: if $(x,y) \in \Gamma'$, just take $(w,z) = (x,y)$ in the r.h.s. of \eqref{E:trG}.
The third point is consequence of the fact that
\[
G = P_{34} \Big( \big( \Gamma' \times X \times X \big) \cap \Big\{ (w,z,x,y) : d_L(w,x) + d_L(x,y) + d_L(y,z) = d_L(w,z) \Big\} \Big),
\]
and the result follows from the properties of analytic sets.
The first point follows from the following observation: if $(x_i,y_i) \in \gamma_{[w_i,z_i]}$, then from triangle inequality
\begin{align*}
d_L(x_{i+1},y_i) - d_L(x_i,y_i) + d_L(x_i,y_{i-1}) \geq&~ d_L(x_{i+1},z_i) - d_L(z_i,y_i) - d_L(x_i,y_i) + d_L(x_i,y_{i-1}) \crcr
=&~ d_L(x_{i+1},z_i) - d_L(x_i,z_i) + d_L(x_i,y_{i-1}) \crcr
\geq&~ d_L(x_{i+1},z_i) - d_L(x_i,z_i) + d_L(w_i,y_{i-1}) - d_L(w_i,x_i) \crcr
=&~ d_L(x_{i+1},z_i) - d_L(w_i,z_i) + d_L(w_i,y_{i-1}).
\end{align*}
Repeating the above inequality finitely many times one obtain
\[
\sum_i d_L(x_{i+1},y_i) - d_L(x_i,y_i) \geq \sum_i d_L(w_{i+1},z_i) - d_L(w_i,z_i) \geq 0.
\]
Hence the set $G$ is $d_L$-cyclically monotone.
\end{proof}
\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 $G$ it is fairly easy to prove that $\mathcal{T}$, $\mathcal{T}_e$ are analytic sets. The subscript $e$ refers to the endpoints of the geodesics: clearly we have
\begin{equation}
\label{E:RTedef}
\mathcal{T}_e = P_1(R \setminus \{x = y\}).
\end{equation}
The following lemma shows that we have only to study the Monge problem in $\mathcal{T}_e$.
\begin{lemma}
\label{L:mapoutside}
It holds $\pi(\mathcal{T}_e \times \mathcal{T}_e \cup \{x = y\}) = 1$.
\end{lemma}
\begin{proof}
If $x \in P_1(\Gamma \setminus \{x=y\})$, then $x \in G^{-1}(y) \setminus \{y\}$ for some $y\in X$. Similarly, $y \in P_2(\Gamma \setminus \{x=y\})$ implies that $y \in G(x) \setminus \{x\}$
for some $x\in X$. Hence $\Gamma \setminus \mathcal{T}_e \times \mathcal{T}_e \subset \{x = y\}$.
\end{proof}
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}
We now use the non branching assumption.
\begin{lemma}
\label{L:uniqr}
If $x \in \mathcal{T}$, then $R(x)$ is a single geodesic.
\end{lemma}
\begin{proof}
Since $x \in \mathcal{T}$, there exists $(w,x), (x,z) \in G \setminus \{x=y\}$: from the $d_L$-cyclical monotonicity and triangular inequality, it follows that
\[
d_L(w,z) = d_L(w,x) + d_L(x,z),
\]
so that $(w,z) \in G$ and $x \in \gamma_{(w,z)}$. Hence from the non branching assumption the set
\[
R(x) = \bigcup_{y \in G(x)} \gamma_{[x,y]} \cup \bigcup_{z \in G^{-1}(x)} \gamma_{[z,x]}
\]
is a single geodesic.
\end{proof}
\begin{proposition}
\label{P:equiv}
The set $R \cap \mathcal{T} \times \mathcal{T}$ is an equivalence relation on $\mathcal{T}$. The set $G$ is a partial order relation on $\mathcal{T}_e$.
\end{proposition}
\begin{proof}
Using the definition of $R$, one has in $\mathcal{T}$:
\begin{enumerate}
\item $x \in \mathcal{T}$ implies that
\[
\exists y \in G(x) \setminus \{x = y\},
\]
so that from the definition of $G$ it follows $(x,x) \in G$;
\item if $y \in R(x)$, $x,y \in \mathcal{T}$, then from Lemma \ref{L:uniqr} there exists $(w,z) \in G$ such that $x , y \in \gamma_{(w,z)}$. Hence $x \in R(y)$;
\item if $y \in R(x)$, $z \in R(y)$, $x,y,z \in \mathcal{T}$, then from Lemma \ref{L:uniqr} it follows again there exists $(w,z) \in G$ such that $x , y , z \in \gamma_{(w,z)}$. Hence $z \in R(x)$.
\end{enumerate}
The second part follows similarly:
\begin{enumerate}
\item $x \in \mathcal{T}_e$ implies that
\[
\exists (x,y) \in \big( G \setminus \{x=y\} \big) \cup \big( G^{-1} \setminus \{x=y\} \big),
\]
so that in both cases $(x,x) \in G$;
\item as in Lemma \ref{L:uniqr}, $(x,y), (y,z) \in G \setminus \{x=y\}$ implies by $d_L$-cyclical monotonicity that $(x,z) \in G$.
\end{enumerate}
\end{proof}
\begin{remark}
\label{R:orderX}
Note that $G \cup \{x=y\}$ is a partial order relation on $X$.
\end{remark}
\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{figure}
\label{Fi:mongemetri3}
\psfrag{Gamma}{\color{MidnightBlue}$\Gamma$}
\psfrag{Gamma'}{\color{Brown}$\Gamma'$}
\psfrag{b}{\color{Purple}$b$}
\psfrag{a}{\color{Goldenrod}$a$}
\psfrag{G}{\color{Cyan}$G$}
\psfrag{G-1}{\color{CarnationPink}$G^{-1}$}
\psfrag{mu}{\color{Red}$\mu$}
\psfrag{nu}{\color{Green}$\nu$}
\psfrag{T}{\color{Cyan}$\mathcal{T}$}
\centerline{\resizebox{12cm}{12cm}{\includegraphics{mongemetric3.eps}}}
\caption{Construction of the sets $\Gamma$, $\Gamma'$, $G$, $G^{-1}$, $a$, $b$ in a 1-dimensional example with $d_{L}=|\cdot|$.}
\end{figure}
\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}
\begin{proof}
Define
\[
C := \Big\{ (x,y,z) \in \mathcal{T}_e \times \mathcal{T}_e \times \mathcal{T}_e: y \in G(x), z \in G(y) \Big\} = (G \times X) \cap (X \times G) \cap \mathcal{T}_e \times \mathcal{T}_e \times \mathcal{T}_e,
\]
that is clearly analytic. Then
\[
b = \Big\{ (x,y) \in G : y \in G(x), G(y) \setminus \{y\} = \emptyset \Big\} = G \setminus P_{1,2}(C \setminus X \times \{y=z\}),
\]
\[
b(A) = \Big\{ y: y \in G(x), G(y) \setminus \{y\} = \emptyset, x \in A \Big\} = P_2(G \cap A \times X) \setminus P_2(C \setminus X \times \{y=z\}).
\]
A similar computation holds for $a$:
\[
a = G^{-1} \setminus P_{23}(C \setminus \{x=y\} \times X), \quad a(A) = P_1(G \cap X \times A) \setminus P_1(C \setminus \{x=y\} \times X)
\]
Hence $a,b \in \mathcal{A}(X \times X)$, $a(A), b(A) \in \mathcal{A}(X)$, being the intersection of an analytic set with a coanalytic one.
If $x \in \mathcal{T}$, then from Lemma \ref{L:uniqr} it follows that $a(x)$, $b(x)$ are empty or singletons and $a(x) \not= b(x)$. If $x \in \mathcal{T}_e \setminus \mathcal{T}$, then it follows that the geodesic $\gamma_{[w,z]}$, $(w,z) \in G$, to which $x$ belongs cannot be prolonged in at least one direction: hence $x \in a(x) \cup b(x)$.
The other point follows easily.
\end{proof}
We finally show that we can assume that the $\mu$-measure of final points and the $\nu$-measure of the initial points are $0$.
\begin{lemma}
\label{L:finini0}
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.
\end{lemma}
\begin{proof}
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 $a$.
\end{proof}
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
\[
\mu(b(\mathcal{T})) = \nu(a(\mathcal{T})) = 0.
\]
\begin{remark}
\label{R:lsccase1}
In the case considered in Remark \ref{R:compact}, it is possible to obtain more regularity for the sets introduced so far.
Recall that we are now assuming
\begin{itemize}
\item[(2')] $d_L : X \times X \to [0,+\infty]$ l.s.c. distance,
\item[(4')] $d_L(x,y) \geq d(x,y)$,
\item[(5')] $\cup_{x \in K_1, y \in K_2} \gamma_{[x,y]}$ 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 (5') 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 if we are in the case of Remark \ref{R:compact}, $\Gamma$, $\Gamma'$, $G$, $G^{-1}$, $a$ and $b$ are $\sigma$-compact sets.
\end{remark}
\begin{remark}
\label{R:TR}
Many simplifications occur in the case the disintegration w.r.t. the partition $\{D_L(x)\}_{x \in X}$ is strongly consistent. Recall that
$D_L(x) = \big\{ y : d_L(x,y) < + \infty \big\}$. Let
\[
\pi = \int_0^1 \pi_\alpha m(d\alpha), \quad \mu = \int_0^1 \mu_\alpha m(d\alpha), \ \nu = \int_0^1 \nu_\alpha m(d\alpha)
\]
be strongly consistent disintegrations such that
\[
\mu_\alpha(D_L(x_\alpha)) = \nu_\alpha(D_L(x_\alpha)) = 1, \quad \pi_\alpha \in \Pi(\mu_\alpha,\nu_\alpha).
\]
We have used the fact that the partition $\{D_L(x) \times D_L(x)\}_{x \in X}$ has the crosswise structure, and then we can apply the results of \cite{biacar:cmono}.
\medskip
\noindent {\it 1) Optimality of $\pi_\alpha$.} Since $\pi$ is $d_L$-cyclically monotone, also the $\pi_\alpha$ are $d_L$-cyclically monotone: precisely they are concentrated on the sets
\[
\Gamma_\alpha = \Gamma \cap D_L(x_\alpha) \times D_L(x_\alpha),
\]
if $\Gamma$ is $d_L$-cyclically monotone and $\pi(\Gamma) = 1$.
Using the fact that $(D_L(x_\alpha),d_L)$ is a metric space, then we can construct a potential $\varphi(x,x_\alpha)$ using the formula
\[
\varphi(x,x_\alpha) = \inf \bigg\{ \sum_{i=0}^I d_L(x_{i+1},y_i) - d_L(x_i,y_i), (x_i,y_i) \in \Gamma_\alpha, x_{I+1} = x, (x_0,y_0) = (x_\alpha,x_\alpha) \bigg\}.
\]
and since this is bounded on $(D_L(x_\alpha),d_L)$, we see that $\pi_\alpha$ and hence $\pi$ are optimal.
\medskip
\noindent {\it 2. Potential for $\pi$.} Extend $\varphi(x,x_\alpha)$ to $X$ by setting $\varphi(x,x_\alpha) = +\infty$ if $x \notin D_L(x_\alpha)$. If $\{(x_\alpha,x_\alpha)\}_{\alpha \in [0,1]}$ is a Borel section, then the function
\[
\varphi(x) = \inf_\alpha\{\varphi(x,\alpha)\}
\]
is easily seen to be analytic. This function is clearly a potential for $\pi$. In particular, it follows again from \cite{biacar:cmono} that $\pi$ is optimal if it is $d_L$-cyclically monotone.
\medskip
\noindent {\it 3.Transport set.} We can then define the set of oriented transport rays as the set
\[
G = \Big\{ (x,y) \in X\times X: \varphi (x) - \varphi(y) = d_L(x,y) \Big\}.
\]
In general, this sets is larger than the one of definition \ref{D:Gray}.
\end{remark}
\section{Partition of the transport set $\mathcal{T}$}
\label{S:partition}
In this section we use the continuity and local compactness of geodesics to show 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 $R$.
Let $\{x_i\}_{i \in \mathbb{N}}$ be a dense sequence in $(X,d)$.
\begin{lemma}
\label{L:reguloclco}
The sets
\[
W_{ijk} := \Big\{ x \in \mathcal{T} \cap \bar B_{2^{-j}}(x_i): L(G(x)),L(G^{-1}(x)) \geq 2^{2-k}, L \big( R(x) \cap \bar B_{2^{1-j}}(x_i) \big) \leq 2^{-k} \Big\}
\]
form a countable covering of $\mathcal{T}$ of class $\mathcal{A}$.
\end{lemma}
\begin{proof}
We first prove the measurability. We consider separately the conditions defining $W_{ijk}$.
{\it Point 1.} The set
\[
A_{ij} := \mathcal{T} \cap \bar B_{2^{-j}}(x_i)
\]
is clearly analytic.
{\it Point 2.} The set
\[
B_k := \big\{ x \in \mathcal{T}: L(G(x)) \geq 2^{2-k} \big\} = P_1 \Big( G \cap \big\{ d_L(x,y) \geq 2^{2-k} \big\} \Big)
\]
is again analytic, being the projection of an analytic set. Similarly, the set
\[
C_k := \big\{ x \in \mathcal{T}: L(G^{-1}(x)) \geq 2^{2-k} \big\} = P_1 \Big( G^{-1} \cap \big\{ d_L(x,y) \geq 2^{2-k} \big\} \Big)
\]
is again analytic.
{\it Point 3.} The set
\begin{align*}
D_{jk} :=&~ \big\{ x \in \mathcal{T}: L \big( R(x) \cap \bar B_{2^{-j}}(x_i) \big) \leq 2^{-k} \big\} \crcr
=&~ \mathcal{T} \setminus P_1 \Big( R \cap \big( \{(x,y): d(x_i,y) \leq 2^{1-j}\} \cap \{d_L(x,y) > 2^{-k}\} \big) \Big)
\end{align*}
is in the $\mathcal{A}$-class, being the difference of two analytic sets.
We finally can write
\begin{align*}
W_{ijk} = A_{ij} \cap B_k \cap C_k \cap D_{jk},
\end{align*}
and the fact that $\mathcal{A}$ is a $\sigma$-algebra proves that $W_{ijk} \in \mathcal{A}$.
To show that it is a covering, notice that for all $x \in \mathcal{T}$ it holds
\[
\min \big\{ L(G(x)), L(G^{-1}(x)) \big\} \geq 2^{2-\bar k}
\]
for some $\bar k \in \mathbb{N}$.
From the local compactness of geodesics, Assumption \eqref{Cond:XdL5} of page \pageref{P:assumpDL}, it follows that if $\gamma^{-1}(\bar B_r(x))$ is compact, then the continuity of $\gamma$ implies that $\gamma^{-1}(\bar B_{r'}(x))$ is also compact for all $r' \leq r$, and $\textrm{diam\,}_{d_L}(\gamma \cap \bar B_{r'}(x)) \to 0$ and $r' \to 0$. In particular there exists $\bar j \in \mathbb{N}$ such that
\[
L \big( R(x) \cap \bar B_{2^{1-\bar j}}(x) \big) \leq 2^{-\bar k},
\]
with $\bar k$ the one chosen above.
Finally, one choose $x_{\bar i}$ such that $d(x,x_{\bar i}) < 2^{-1-\bar j}$, so that $x \in \bar B_{2^{-\bar j}}(x_{\bar i}) \subset \bar B_{2^{1-\bar j}}(x)$ and thus
\[
L \big( R(x) \cap \bar B_{2^{-\bar j}}(x_{\bar i}) \big) \leq 2^{-\bar k}.
\]
\end{proof}
\begin{lemma}
\label{L:partitilW}
There exist $\mu$-negligible sets $N_{ijk} \subset W_{ijk}$ such that the family of sets
\begin{align*}
\mathcal{T}_{ijk} = R^{-1}(W_{ijk} \setminus N_{ijk})
\end{align*}
is a countable covering of $\mathcal{T} \setminus \cup_{ijk} N_{ijk}$ into saturated analytic sets.
\end{lemma}
\begin{proof}
First of all, since $W_{ijk} \in \mathcal{A}$, then there exists $\mu$-negligible set $N_{ijk} \subset W_{ijk}$ such that $W_{ijk} \setminus N_{ijk} \in \mathcal{B}(X)$. Hence $\{W_{ijk} \setminus N_{ijk}\}_{i,j,k \in \mathbb{N}}$ is a countable covering of $\mathcal{T} \setminus \cup_{ijk} N_{ijk}$. It follows immediately that $\{\mathcal{T}_{ijk}\}_{i,j,k \in \mathbb{N}}$ satisfies the lemma.
\end{proof}
\begin{remark}
\label{R:compFF}
Observe that $\bar B_{2^{-j}}(x_i) \cap R(x)$ is compact for all $x \in \mathcal{T}_{ijk}$: in fact, during the proof of Lemma \ref{L:reguloclco} we have already shown that $\gamma^{-1}(\bar B_{2^{-j}} (x_i))$ is compact.
\end{remark}
From any analytic countable covering, we can find a countable partition into $\mathcal{A}$-class saturated sets by defining
\begin{equation}
\label{E:Zkije}
\mathcal{Z}_{m,e} := \mathcal{T}_{i_mj_mk_m} \setminus \bigcup_{m' = 1}^{m-1} \mathcal{T}_{i_{m'}j_{m'}k_{m'}}, \quad \mathcal{Z}_{0,e} := \mathcal{T}_e \setminus \bigcup_{m \in \mathbb{N}} \mathcal{Z}_{m,e},
\end{equation}
where
\[
\mathbb{N} \ni m \mapsto (i_m,j_m,k_m) \in \mathbb{N}^3
\]
is a bijective map. Intersecting the above sets with $\mathcal{T}$, we obtain the countable partition of $\mathcal{T}$ in $\mathcal{A}$-sets
\begin{equation}
\label{E:Zkij}
\mathcal{Z}_m := \mathcal{Z}_{m,e} \cap \mathcal{T}, \quad m \in \mathbb{N}_0.
\end{equation}
Now we use this partition to prove the strong consistency of the disintegration.
On $\mathcal{Z}_m$, $m > 0$, we define the closed values map
\begin{equation}
\label{E:mapTijkF}
\mathcal{Z}_m \ni x \mapsto F(x) := R(x) \cap \bar B_{2^{-j_m}}(x_{i_m}) \in \mathcal{K} \big( \bar B_{2^{-{j_m}}}(x_{i_m}) \big),
\end{equation}
where $\mathcal{K}(\bar B_{2^{-{j_m}}}(x_{i_m}))$ is the space of compact subsets of $\bar B_{2^{-{j_m}}}(x_{i_m})$.
\begin{proposition}
\label{P:sicogrF}
There exists a $\mu$-measurable cross section $f : \mathcal{T} \to \mathcal{T}$ for the equivalence relation $R$.
\end{proposition}
\begin{proof}
First we show that $F$ is $\mathcal{A}$-measurable: for $\delta > 0$,
\begin{align*}
F^{-1}(B_\delta(y)) =&~ \Big\{ x \in \mathcal{Z}_m: R(x) \cap B_{\delta}(y) \cap \bar B_{2^{-{j_m}}}(x_{i_m}) \not= \emptyset \Big\} \crcr
=&~ \mathcal{Z}_m \cap P_1 \Big( R \cap \big( X \times B_{\delta}(y) \cap \bar B_{2^{-{j_m}}}(x_{i_m}) \big) \Big).
\end{align*}
Being the intersection of two $\mathcal{A}$-class sets, $F^{-1}(B_\delta(y))$ is in $\mathcal{A}$.
By Corollary \ref{C:weelsupprr} there exists a $\mathcal{A}$-class section $f_m : \mathcal Z_m \to \bar B_{2^{-{j_m}}}(x_{i_m})$. The proposition follows by setting $f \llcorner_{\mathcal Z_m} = f_m$ on $\cup_m \mathcal Z_m$, and defining it arbitrarily on $\mathcal T \setminus \cup_m \mathcal Z_m$: the latter being negligible, $f$ is $\mu$-measurable.
\end{proof}
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.
\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}
\begin{proof}
For the first point just observe that
\begin{align*}
g^+ =&~ \Big\{ (y,t,x): y \in S, t \in \mathbb{R}^+, x \in G(y) \cap \{d_L(x,y) = t\} \Big\} \crcr
=&~ S \times R^+ \times X \cap \{(y,t,x):(y,x) \in G\} \cap \{(y,t,x):d_L(x,y) = t\} \in \Sigma^1_1.
\end{align*}
Similarly
\[
g^- = \Big\{ (y,t,x): y \in S, t \in \mathbb{R}^-, x \in G^{-1}(y) \cap \{d_L(x,y) = -t\} \Big\} \in \Sigma^1_1.
\]
Since $\mathcal{S} \subset \mathcal{T}$ and $R(y)$ is a subset of a single geodesic for $y \in \mathcal{S} \subset \mathcal{T}$,
$g$ is the graph of a map.
Note that for any $x \in\mathcal{T}_{e}$ there exists $z \in \mathcal{T}$ such that $x\in R(z)$: hence $x\in R(f(z))$, and therefore the range of the map is
the whole $\mathcal{T}_{e}$.
The third point is a direct consequence of the definition.The fourth point follows by substitution.
\end{proof}
\begin{figure}
\label{Fi:mongemetri4}
\psfrag{W}{$W_{ijk}$}
\psfrag{S}{\color{Red}$\mathcal{S}$}
\psfrag{g}{$g$}
\psfrag{R}{$\mathbb{R}$}
\centerline{resizebox{15cm}{5cm}{\includegraphics{mongemetric.4.eps}}}
\caption{The ray map $g$.}
\end{figure}
We finally prove the following 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}
We will use the disintegration technique exploited also in the next section. We observe that another proof can be the direct composition of the transference plan with itself, using the fact that the mass moves along geodesics and the disintegration makes the problem one dimensional.
\begin{proof}
We have already shown that we can take
\[
\mu(P_2(b)) = \nu(P_2(a)) = 0,
\]
so that $\mu \wedge \nu$ is concentrated on $\mathcal{T}_S$.
{\it Step 1.} On $\mathcal{T}$ we can use the Disintegration Theorem to write
\begin{equation}
\label{E:disintT1}
\mu \llcorner_{\mathcal T} = \int_S \mu_y m(dy), \quad m = f_\sharp (\mu \llcorner_{\mathcal T}), \ \mu_y \in \mathcal{P}(R(y) \cap \mathcal T).
\end{equation}
In fact, the existence of a Borel section is equivalent to the strong consistency of the disintegration. Since $\{R(y) \times X\}_{y \in \mathcal T}$ is also a partition on $\mathcal T \times X$, we can similarly write
\[
\pi \llcorner_{\mathcal T \times X} = \int_S \pi_y m(dy), \quad \pi_y(R(y)\times R(y)) = 1.
\]
We write moreover
\begin{equation}
\label{E:nuy1}
\nu_y := (P_2)_\sharp (\pi \llcorner_{\mathcal T \times X}), \quad \tilde \nu := \int_S \nu_y m(dy) = \int_S (P_2)_\sharp \pi_y m(dy).
\end{equation}
Clearly the rest of the mass starts from $a(\mathcal T)$, so we have just to show how to rearrange the transference plan in $\mathcal T$ in order to obtain $\mu \perp \nu$. Using $g$, we can reduce the problem to a transport problem on $S \times \mathbb{R}$ with cost
\[
c((y,t),(y',t')) =
\begin{cases}
|t - t'| & y = y' \crcr
+ \infty & y \not= y'
\end{cases}
\]
By standard regularity argument, we can assume that $S \ni y \mapsto \pi_y \in \mathcal{P}(R(y) \times R(y))$ is $\sigma$-continuous, i.e. its graph is $\sigma$-compact.
{\it Step 2.} Using the fact that $(\mu,\nu) \mapsto \mu \wedge \nu$ is Borel w.r.t. the weak topology \cite{biacar:cmono}, we can assume that $S \ni y \mapsto \mu_y \wedge \nu_y \in \mathcal{P}(R(y))$ is $\sigma$-continuous, so that also the map
\[
S \ni y \mapsto (\mu_y - \mu_y \wedge \nu_y, \nu_y - \mu_y \wedge \nu_y) \in \mathcal{P}(R(y)) \times \mathcal{P}(R(y))
\]
is $\sigma$-continuous.
{\it Step 3.} Since in each $R(y)$ the problem is one dimensional, one can take the unique transference plan
\[
\tilde \pi_y \in \Pi \big( \mu_y - \mu_y \wedge \nu_y, \nu_y - \mu_y \wedge \nu_y \big)
\]
concentrated on a monotone set: clearly
\[
\int d_L \tilde \pi_y = \int d_L \pi_y.
\]
{\it Step 4.} If we define the left-continuous distribution functions
\[
H(y,s) := \big( \mu_y - \mu_y \wedge\nu_y \big)(-\infty,s), \quad F(y,t) := \big( \nu_y - \mu_y \wedge\nu_y \big)(-\infty,t),
\]
and
\[
G(y,s,t) := \tilde \pi_y \big( (-\infty,s) \times (-\infty,t) \big),
\]
then the measure $\tilde \pi_y$ is uniquely determined by $G(y,s,t) = \min \{ H(y,s), F(y,t) \}$.
The $\sigma$-continuity of $y \mapsto (\mu_y - \mu_y \wedge \nu_y, \nu_y - \mu_y \wedge \nu_y)$ yields that $H$, $F$ are again $\sigma$-l.s.c., so that $G$ is Borel, and finally $y \mapsto \tilde \pi_y$ is $\sigma$-continuous up to a $f_\sharp \mu$-negligible set.
{\it Step 5.} Define
\[
\hat \pi_y := \tilde \pi_y + (\mathbb{I},\mathbb{I})_\sharp (\mu_y \wedge \nu_y) \in \Pi(\mu_y,\nu_y).
\]
The above steps show that $\hat \pi$ is $m$-measurable, and thus we can define the measure
\[
\hat \pi := \pi \llcorner_{(\mathcal T_e \setminus \mathcal T) \times X} + \int \hat \pi_y m(dy).
\]
It is routine to check that $\hat \pi$ has the required properties.
\end{proof}
\section{Regularity of the disintegration}
\label{S:regurlr}
This section is divided in two parts.
In the first one we consider the translation of Borel sets by the optimal geodesic flow, we introduce a first regularity assumption (Assumption \ref{A:NDEatom}) on the measure $\mu$ and we show that an immediate consequence is that the set of initial points is negligible. A second consequence is that the disintegration of $\mu$ w.r.t. $R$ has continuous conditional probabilities.
In the second part we consider a stronger regularity assumption (Assumption \ref{A:NDE}) which gives that the conditional probabilities are absolutely continuous with respect to $\mathcal{H}^{1}$ along geodesics.
\subsection{Evolution of Borel sets}
\label{Ss:evolution}
Let $A \subset \mathcal{T}_e$ be an analytic set and define for $t \in \mathbb{R}$ the \emph{$t$-evolution $A_t$ of $A$} by
\begin{equation}
\label{E:At}
A_t := g \big( g^{-1}(A) + (0,t) \big).
\end{equation}
\begin{lemma}
\label{L:evolution}
The set $A_t \cap g(S \times \mathbb{R})$ is analytic, and $A_t$ is $\mu$-measurable for $t \geq 0$.
\end{lemma}
\begin{proof}
Divide $A$ into two parts:
\[
A_S := A \cap g(S \times \mathbb{R}) \quad \text{and} \quad A_N : = A \setminus A_S.
\]
From Point (1) of Proposition \ref{P:gammaclass} it follows that $A_S$ is analytic. We consider the evolution of the two sets separately.
Again by Point (1) of Proposition \ref{P:gammaclass}, the set $(A_S)_t$ is analytic, hence universally measurable for all $t \in \mathbb{R}$.
Since $\mathcal{T}_N$ is $\mu$-negligible (see \eqref{E:TNngel}), it follows that $(A_N)_t$ is $\mu$-negligible for all $t>0$, and by the assumptions it is clearly measurable for $t=0$.
\end{proof}
We can show that $t \mapsto \mu(A_t)$ is measurable.
\begin{lemma}
\label{L:measumuAt}
Let $A$ be analytic. The function $t \mapsto \mu(A_t)$ is Souslin for $t \geq 0$.
If $A \subset g(S \times \mathbb{R})$, then $t \mapsto \mu(A_t)$ is Souslin for $t \in \mathbb{R}$.
\end{lemma}
\begin{proof}
As before, we split the $A$ into the sets
\[
A_S := A \cap g(S \times \mathbb{R}) \quad \text{and} \quad A_N : = A \setminus A_S.
\]
The function
\[
t \mapsto \mu(A_{N,t}) =
\begin{cases}
\mu(A_N) & t = 0 \crcr
0 & t > 0
\end{cases}
\]
is clearly Borel. Observe that since $\mathcal{T}_{N}\subset \mathcal{T}$ and the $\mu$-measure of final points is 0,
the value of $\mu(A_{N,t})$ is known only for $t>0$.
Since $A_S$ is analytic, then $g^{-1}(A_S)$ is analytic, and the set
\[
\tilde A_S := \big\{ (y,\tau,t) : (y,\tau-t) \in g^{-1}(A_S) \big\}
\]
is easily seen to be again analytic.
Define the analytic set $\hat A_{S} \subset X \times \mathbb{R}$ by
$$
\hat{A}_{S} : = (g,\mathbb{I}) (\tilde A_{S}).
$$
Clearly $(A_{S})_{t} = \hat A_{S}(t)$. We now show in two steps that the function $t \mapsto \mu((A_{S})_{t})$ is analytic.
{\it Step 1.} Define the closed set in $\mathcal{P}(X \times [0,1])$
$$
\Pi(\mu) : = \big\{ \pi \in \mathcal{P}(X\times [0,1]) : (P_{1})_{\sharp}(\pi)= \mu \big\}
$$
and let $B \subset X\times \mathbb{R} \times [0,1]$ a Borel set such that $P_{12}(B)= \hat A_{S}$.
Consider the function
\[
\mathbb{R} \times \Pi(\mu) \ni (t, \pi) \mapsto \pi(B(t)).
\]
A slight modification of Lemma 4.12 in \cite{biacar:cmono} shows that this function is Borel.
{\it Step 2.}
Since supremum of Borel function are Souslin, pag. 134 of \cite{Sri:courseborel},
the proof is concluded once we show that
\[
\mu((A_{S})_{t}) = \mu (\hat A_{S}(t)) = \sup_{ \pi \in \Pi(\mu) } \pi(B(t)).
\]
From the Disintegration Theorem, for all $\pi \in \Pi(\mu)$ we have
\[
\pi(B(t)) = \int \pi_{x}(B(t)) \mu(dx) \leq \int_{P_{1}(B(t))} \mu(dx) = \mu (\hat A_{S}(t)).
\]
On the other hand from Theorem \ref{T:vanneuma}, there exists an $\mathcal{A}$-measurable section $u: \hat A_{S}(t) \to B(t)$.
Clearly for $\pi_{u} = (\mathbb{I},u)_{\sharp}(\mu)$ it holds $\pi_{u}(B(t))= \mu (\hat A_{S}(t))$.
\end{proof}
The next assumption is the first fundamental assumption of the paper.
\begin{assumption}[Non-degeneracy assumption]
\label{A:NDEatom}
For all Borel sets $A$ such that $\mu(A) > 0$ the set $\{t \in \mathbb{R}^+: \mu(A_t) > 0\}$ has cardinality $> \aleph_0$.
\end{assumption}
By inner regularity, it is clearly enough to verifies Assumption \ref{A:NDEatom} only for compact sets.
Note that since for analytic set Cantor Hypothesis holds true, Theorem 4.3.5, pag. 142 of \cite{Sri:courseborel} ,
Assumption \ref{A:NDEatom} implies that the cardinality of $\{t \in \mathbb{R}^+: \mu(A_t) > 0\}$ is $\mathfrak{c}$.
An immediate consequence of the Assumption \ref{A:NDEatom} is that the measure $\mu$ is concentrated on $\mathcal{T}$.
\begin{lemma}
\label{L:puntini}
If $\mu$ satisfies Assumption \ref{A:NDEatom} then
\[
\mu(\mathcal{T}_e\setminus \mathcal{T}) = 0.
\]
\end{lemma}
\begin{proof}
If $A \subset a(X)$, then $A_t \cap A_s = \emptyset$ for $0 \leq s < t$. Hence
\[
\sharp \big\{ t \in \mathbb{R}^+: \mu(A_t) > 0 \big\} \leq \aleph_0,
\]
because of the boundedness of $\mu$. This contradicts the assumptions.
\end{proof}
Once we know that $\mu(\mathcal{T}) = 1$, 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.
\begin{figure}
\label{Fi:mongem5}
\psfrag{mu}{$\mu$}
\psfrag{nu}{$\nu$}
\psfrag{B}{$B$}
\psfrag{Bt}{$B_t$}
\psfrag{gamma}{$\gamma$}
\resizebox{11cm}{6cm}{\includegraphics{mongemetric.6.eps}}
\caption{The evolution of a set $B$ through the optimal flow.}
\end{figure}
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}
The conditional probabilities $\mu_y$ are continuous for $m$-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 $<\epsilon$ such that $L(R(y)) > \epsilon$ for all $y \in K$. 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 it is Borel, and by Lusin Theorem (Theorem 5.8.11 of \cite{Sri:courseborel}) it is the countable union of Borel graphs: setting in case $c_i(y) = 0$, we can consider them as Borel functions on $S$ and order them w.r.t. $G$,
\[
\mu_{y,\textrm{atomic}} = \sum_{i \in \mathbb{Z}} c_i(y) \delta_{x_i(y)}, \quad x_{i+1}(y) \in G(x_i(y)), \ i \in \mathbb{Z}.
\]
{\it Step 2.} Define the sets
\[
S_{ij}(t) := \Big\{ y \in K: x_i(y) = g \big( g^{-1}(x_j(y)) + t \big) \Big\} \cap \mathcal T.
\]
Since $K \subset S$, to define $S_{ij}$ we are using the graph $g \cap S \times \mathbb{R} \times \mathcal{T}$, which is analytic: hence $S_{ij} \in \Sigma^1_1$.
For $A_j := \{x_j(y), y \in K\}$ and $t \in \mathbb{R}^+$ we have that
\begin{align*}
\mu((A_j)_t) =&~ \int_K \mu_y((A_j)_t) m(dy) = \int_K \mu_{y,\textrm{atomic}}((A_j)_t) m(dy) \crcr
=&~ \sum_{i \in \mathbb{Z}} \int_K c_i(y) \delta_{x_i(y)} \big( g(g^{-1}(x_j(y)) + t) \big) m(dy) = \sum_{i \in \mathbb{Z}} \int_{S_{ij}(t)} c_i(y) m(dy).
\end{align*}
We have used the fact that $A_j \cap R(y)$ is a singleton.
{\it Step 3.} For fixed $i,j \in \mathbb{N}$, again from the fact that $A_j \cap R(y)$ is a singleton
\[
S_{ij}(t) \cap S_{ij}(t') =
\begin{cases}
S_{ij}(t) & t = t' \crcr
\emptyset & t \not= t'
\end{cases}
\]
so that
\[
\sharp \big\{ t : m(S_{ij}(t)) > 0 \big\} \leq \aleph_0.
\]
Finally
\[
\mu((A_j)_t) > 0 \quad \Longrightarrow \quad t \in \bigcup_i \big\{ t : m(S_{ij}(t)) > 0 \big\},
\]
whose cardinality is $\leq \aleph_0$, contradicting Assumption \ref{A:NDEatom}.
\end{proof}
\subsection{Absolute continuity}
\label{Ss:regolarita'}
We next assume a stronger regularity assumption.
\begin{assumption}[Absolute continuity assumption]
\label{A:NDE}
For every Borel set $A \subset \mathcal{T}_{e}$
\[
\mu(A) > 0 \quad \Longrightarrow \quad \int_0^{+\infty} \mu(A_t) dt > 0.
\]
\end{assumption}
Again by inner regularity, Assumption \ref{A:NDE} can be verified only for compact sets.
Note that the condition is meaningful by Lemma \ref{L:measumuAt}. Observe moreover that Assumption \ref{A:NDE} implies Assumption \ref{A:NDEatom}, so that in the following we will restrict the map $g$ to the set $g^{-1}(\mathcal{T})$, where it is analytic. Moreover, we can consider shift $t \mapsto A_t$ for $t \in \mathbb{R}$, because of Lemma \ref{L:measumuAt}.
\begin{remark}\label{R:NDE}
An equivalent form of the Assumption \ref{A:NDE} is the following:
\[
\mu(A)>0 \quad \Longrightarrow \quad \int_{t,s \geq 0} \mu(A_t \cap A_s) dtds > 0.
\]
In fact, due to $\mu(X) = 1$, in the set $I_n := \{t:\mu(A_t) > 2^{-n}\}$ the set $\{s \in I_n : \mu(A_s \cap A_t) = 0, t \in I_n\}$ has cardinality at most $2^{-n}$. Hence, since for some $n$ $\mathcal{L}^1(I_n) > 0$ by Assumption \ref{A:NDE}, it follows that
\[
\mathcal{L}^2(I_n \times I_n) = \big( \mathcal{L}^1(I_n) \big)^2 > 0.
\]
The opposite implication is a consequence of Fubini theorem.
\end{remark}
The next results show regularity of the Radon-Nikodym derivative of $\mu_y$ w.r.t. $(\mathcal{H}^1_L)\llcorner_{ f^{-1}(y)}$, where $\mathcal{H}^1_L$ is the $1$-dimensional Hausdorff measure w.r.t. the $d_L$-distance. Note that along $d_L$ $1$-Lipschitz geodesics, $\mathcal{H}^1_L$ is equivalence to $g(y,\cdot)_\sharp \mathcal{L}^1$: in the following we will use both notations.
\begin{lemma}
\label{Lem:dec}
Let $\mu$ be a Radon measure and
\[
\mu_y = r(y,\cdot) g(y,\cdot)_\sharp \mathcal{L}^1 + \omega_y, \quad \omega_y \perp g(y,\cdot)_\sharp \mathcal{L}^1
\]
be the Radon-Nikodym decomposition of $\mu_y$ w.r.t. $g(y,\cdot)_\sharp \mathcal{L}^1$. Then there exists a Borel set $C\subset X$ such that
\[
\mathcal{L}^{1} \big( g^{-1} (C) \cap (\{y\} \times \mathbb{R}) ) \big) = 0
\]
and $\omega_y = \mu_y \llcorner_C$ for $m$-a.e. $y \in [0,1]$.
\end{lemma}
\begin{proof}
Consider the measure
\[
\lambda = g_\sharp (m \otimes \mathcal L^1),
\]
and compute the Radon-Nikodym decomposition
\[
\mu = \frac{D \mu}{D \lambda} \lambda + \omega.
\]
Then there exists a Borel set $C$ such that $\omega = \mu \llcorner_C$ and $\lambda(C)=0$. The set $C$ proves the Lemma. Indeed $C = \cup_{y \in [0,1]} C_{y}$ where $C_{y} = C \cap f^{-1}(y)$ is such that $\mu_y \llcorner_{C_{y}} = \omega_{y} $ and $g(y,\cdot)_{\sharp}\mathcal{L}^{1}(C_{y})=0$ for $m$-a.e. $y \in [0,1]$.
\end{proof}
\begin{theorem}
\label{teo:a.c.}
If $\mu$ satisfies Assumption \ref{A:NDE}, then for $m$-a.e. $y \in [0,1]$ the conditional probabilities $\mu_y$ are absolutely continuous w.r.t. $g(y,\cdot)_\sharp \mathcal{L}^1$.
\end{theorem}
The proof is based on the following simple observation.
\medskip
\noindent Let $\eta$ be a Radon measure on $\mathbb{R}$. Suppose that for all $A \subset \mathbb{R}$ Borel with $\eta(A)>0$ it holds
\[
\int_{\mathbb{R}^+} \eta(A+t) dt = \eta \otimes\mathcal{L}^1 \big( \{ (x,t): t \geq 0, x - t \in A \} \big) > 0.
\]
Then $\eta \ll \mathcal{L}^1$.
\begin{proof}
The proof will use Lemma \ref{Lem:dec}: take $C$ the set constructed in Lemma \ref{Lem:dec} and suppose by contradiction that
\[
\mu(C) > 0 \quad \text{and} \quad m \otimes \mathcal{L}^1 (g^{-1}(C)) = 0.
\]
In particular, for all $t \in \mathbb{R}$ it follows that
\[
m \otimes \mathcal{L}^1 (g^{-1}(C_t)) = m \otimes \mathcal{L}^1 (g^{-1}(C) + (0,t)) = 0.
\]
By Fubini-Tonelli Theorem
\begin{align*}
0< &~ \int_{\mathbb{R}^+} \mu(C_t) dt = \int_{\mathbb{R}^+} \bigg( \int_{g^{-1}(C_t)} (g^{-1})_\sharp \mu(dyd\tau) \bigg) dt \crcr
=&~ \big( (g^{-1})_\sharp \mu \otimes \mathcal{L}^1 \big) \Big( \Big\{ (y,\tau,t): (y,\tau) \in g^{-1}(\mathcal{T}), (y,\tau-t) \in g^{-1}(C) \Big\} \Big) \crcr
\leq&~ \int_{S \times \mathbb{R}} \mathcal{L}^1 \big( \big\{\tau - g^{-1}(C \cap f^{-1}(y)) \big\} \big) (g^{-1})_{\sharp} \mu (dyd\tau) \crcr
=&~ \int_{S \times \mathbb{R}} \mathcal{L}^1 \big( g^{-1}(C \cap f^{-1}(y)) \big) (g^{-1})_{\sharp} \mu (dyd\tau) \crcr
=&~ \int_{S} \mathcal{L}^1 \big( g^{-1}(C \cap f^{-1}(y)) \big) m(dy) = 0.
\end{align*}
That gives a contradiction.
\end{proof}
Now we will study the regularity of the map $t \mapsto \mu(A_t)$ under Assumption \ref{A:NDE}. We will use the following notation:
\[
\mu(A) = \int_S \mu_y(A) m(dy) = \int_S \bigg( \int_{g(y,\cdot)^{-1}(A)} r(y,\tau) d\tau \bigg) m(dy) = g_\sharp (r m \otimes \mathcal L^1).
\]
\begin{proposition}
\label{P:peloso}
$\mu$ satisfies Assumption \ref{A:NDE} if and only if for all $A$ Borel $t \mapsto \mu(A_t)$ is continuous. Moreover if $A$ is geodesically convex then $\mu(A_t)$ is absolutely continuous.
\end{proposition}
\begin{proof}
It is enough to prove the continuity for $t=0$. Since
\[
\mu(A_t) = \int_S \bigg( \int_{g(y,\cdot)^{-1}(A_t)} r(y,\tau) d\tau \bigg) m(dy),
\]
its continuity is a direct consequence of Lebesgue dominated convergence theorem applied to the function:
\[
t \mapsto \mu_y(A_t) = \int_{g(y,\cdot)^{-1}(A_t)} r(y,\tau) d\tau.
\]
Suppose now $A$ geodesically convex. Each $g(y,\cdot)^{-1}(A)$ is an interval $(\alpha(y),\omega(y))$, so that the map
\[
t \mapsto \int_{g(y,\cdot)^{-1}(A_t)} r(y,\tau) d\tau
\]
is absolutely continuous with derivative
\[
h(y,t) = r(y,\omega(y)+t) - r(y,\alpha(y)+t).
\]
Since $h(y,t) \in L^1(m \otimes \mathcal{L}^1)$ the result follows by a standard computation.
\end{proof}
\section{Solution to the Monge problem}
\label{S:Solution}
In this section we show that Theorem \ref{teo:a.c.} allows to construct an optimal map $T$. We recall the one dimensional result for the Monge problem \cite{villa:Oldnew}.
\begin{theorem}
\label{T:oneDmonge}
Let $\mu$, $\nu$ be probability measures on $\mathbb{R}$, $\mu$ continuous, and let
\[
H(s) := \mu((-\infty,s)), \quad F(t) := \nu((-\infty,t)),
\]
be the left-continuous distribution functions of $\mu$ and $\nu$ respectively. Then the following holds.
\begin{enumerate}
\item The non decreasing function $T : \mathbb{R} \to \mathbb{R} \cup [-\infty,+\infty)$ defined by
\[
T(s) := \sup \big\{ t \in \mathbb{R} : F(t) \leq H(s) \big\}
\]
maps $\mu$ to $\nu$. Moreover any other non decreasing map $T'$ such that $T'_\sharp \mu = \nu$ coincides with $T$ on the support of $\mu$ up to a countable set.
\item If $\phi : [0,+\infty] \to \mathbb{R}$ is non decreasing and convex, then $T$ is an optimal transport relative to the cost $c(s,t) = \phi(|s-t|)$. Moreover $T$ is the unique optimal transference map if $\phi$ is strictly convex.
\end{enumerate}
\end{theorem}
Assume that $\mu$ satisfies Assumption \ref{A:NDEatom}. Then we can disintegrate $\mu$ and $\pi$ respect to the ray equivalence relation $R$ and $R \times X$ as in \eqref{E:disintT},
\begin{equation}
\label{E:muy}
\mu = \int \mu_y m(dy), \ \pi = \int \pi_y m(dy), \quad \mu_y \ \text{continuous}, \ (P_1)_\sharp \pi_y = \mu_y.
\end{equation}
We write moreover
\begin{equation}
\label{E:nuy}
\nu = \int \nu_y m(dy) = \int (P_2)_\sharp \pi_y m(dy).
\end{equation}
Note that $\pi_y \in \Pi(\mu_y,\nu_y)$ is $d_L$-cyclically monotone (and hence optimal, because $R(y)$ is one dimensional) for $m$-a.e. $y$. If $\nu(\mathcal{T}) = 1$, then \eqref{E:nuy} is the disintegration of $\nu$ w.r.t. $R$.
\begin{theorem}
\label{T:mongeff}
Let $\pi \in \Pi(\mu,\nu)$ be a $d_L$-cyclically monotone transference plan, and assume that Assumption \ref{A:NDEatom} holds. Then there exists a Borel map $T: X \to X$ with the same transport cost as $\pi$.
\end{theorem}
\begin{proof}
By means of the map $g^{-1}$, we reduce to a transport problem on $S \times \mathbb{R}$, with cost
\[
c((y,s),(y',t)) =
\begin{cases}
|t - s| & y = y' \crcr
+ \infty & y \not= y'
\end{cases}
\]
It is enough to prove the theorem in this setting under the following assumptions: $S$ compact and $S \ni y \mapsto (\mu_y,\nu_y)$ weakly continuous. We consider here the probabilities $\mu_y$, $\nu_y$ on $\mathbb{R}$.
{\it Step 1.} From the weak continuity of the map $y \mapsto (\mu_y,\nu_y)$, it follows that the maps
\[
(y,t) \mapsto H(y,t) := \mu_y((-\infty,t)), \ \ (y,t) \mapsto F(y,t) := \nu_y((-\infty,t))
\]
are easily seen to be l.s.c.. Both are clearly increasing in $t$. Note also that $H$ is continuous in $t$.
{\it Step 2.} The map $T$ defined as Theorem \ref{T:oneDmonge} by
\[
T(y,s) := \Big( y, \sup \big\{ t : F(y,t) \leq H(y,s) \big\} \Big)
\]
is Borel. In fact, for $A$ Borel,
\[
T^{-1}(A \times [t,+\infty)) = \big\{ (y,s) : y \in A, H(y,s) \geq F(y,t) \big\} \in \mathcal{B}(S \times \mathbb{R}).
\]
{\it Step 3.} Note that $\pi_{y}$ and $T(y,\cdot)$ are both optimal for the transport problem between $\mu_{y}$ and $\nu_{y}$ with cost $d_{L}$ restricted to $R(y)$.
Indeed $d_{L}$ restricted to $R(y)\times R(y)$ is finite. Therefore $\pi_{y}$ and $T(y,\cdot)$ have the same cost.
\end{proof}
\begin{remark}\label{R:monot}
By the definition of the set $G$, it follows that along each geodesic $\mu_y(g(y,(-\infty,t))) \geq \nu_y(g(y,(-\infty,t)))$, because in the opposite case $G$ is not $d_L$-cyclically monotone. Hence $T(s) \geq s$, and $c((y,s),T(y,s)) =P_{2}( T(y,s)) - s$. Hence
\begin{equation}\label{E:costo}
\int d_{L} \pi =\int d_{L}(x,T(x))\mu(dx) = \int_{S\times \mathbb{R}} s \big(g(y,\cdot)^{-1}_{\sharp}(\nu_{y} - \mu_{y})\big)(ds)m(dy)= \int P_{2}(g^{-1}(x)) (\nu - \mu)(dx).
\end{equation}
\end{remark}
\section{Dynamic interpretation}
\label{S:div}
In this section we show how the regularity of the disintegration yields a correct definition of the current $\dot g$ representing the flow along the geodesics of an optimal transference plan. This allows to solve the PDE
\[
\partial U = \mu - \nu
\]
in the sense of currents in metric spaces. In particular, under additional regularity assumptions, one can prove that the boundary $\partial \dot g$ is well defined and satisfies an ODE along geodesics. This gives a dynamic interpretation to the transport problem.
The setting here is slightly different from the previous sections:
\begin{enumerate}
\item $d(x,y) \leq d_L(x,y)$;
\item there exists a probability measure $\eta$, such that it (or more precisely $\eta \llcorner_{\mathcal{T}_e}$) satisfies Assumption \ref{A:NDE} along the transport rays of the transportation problem with marginals $\mu$, $\nu$;
\item $\mu \ll \eta$, so that also $\mu$ satisfies Assumption \ref{A:NDE}.
\end{enumerate}
In particular, $\textrm{Lip}(X) \subset \textrm{Lip}_{d_L}(X)$.
The main reference for this chapter is \cite{ambkir:currentmetric}.
\subsection{Definition of $\dot g$}
\label{Ss:dotgam}
For any Lipschitz function $\omega : X \to \mathbb{R}$ we can define the derivative $\partial_t \omega$ along the geodesic $g(t,y)$ for a.e. $t \in \mathbb{R}$,
\[
\partial_t \omega(g(y,t)) := \frac{d}{dt} \omega(g(t,y)).
\]
Using the disintegration formula
\[
\eta \llcorner_{\mathcal T} = \int (g(y,\cdot))_\sharp (q(y,\cdot) \mathcal L^1) m(dy) = g_\sharp (q m \otimes \mathcal L^1)
\]
for some $q \in L^{1}(m \otimes \mathcal{L}^{1})$ (Theorem \ref{teo:a.c.}), we can define the measure $\partial_t \omega \eta$ as
\[
\int \phi(x) (\partial_{t} \omega \eta)(dx) := \int_S \int_\mathbb{R} \phi(g(y,t)) \partial_t \omega(g(y,t)) q(y,t) dt m(dy).
\]
where $\phi \in C_b(X,\mathbb{R})$.
\begin{definition}
\label{D:dotgamma}
We define the \emph{flow $\dot g$} as the current
\[
\langle \dot{g}, (h, \omega) \rangle = \int_{S \times \mathbb{R}} h(g(y,t)) \partial_t \omega(g(y,t)) q(y,t) dt m(dy)
\]
where $h$, $\omega$ are Lipschitz functions of $(X,d)$ with $h$ bounded.
\end{definition}
It is fairly easy to see that $\dot g$ is a current: in fact,
\begin{enumerate}
\item $\dot g$ has finite mass, namely
\[
\big| \langle \dot g, (h,\omega) \rangle \big| \leq \textrm{Lip}(\omega) \int h \eta;
\]
\item $\dot g$ is linear in $h$, $\omega$;
\item if $\omega_n \to \omega$ pointwise in $X$ with uniformly bounded Lipschitz constant, then by Lebesgue Dominated Convergence Theorem if follows that
\[
\lim_{n \to +\infty} \langle \dot g, (h_n,\omega_n) \rangle = \langle \dot g, (h,\omega) \rangle;
\]
\item $\langle \dot g, (h,\omega) \rangle = 0$ if $\omega$ is constant in $\{h \not= 0\}$.
\end{enumerate}
In general, $\dot g$ is only a current, with boundary $\partial \dot g$ defined by the duality formula
\begin{equation}
\label{E:boundgamma}
\langle \partial \dot g, \omega \rangle = \langle \dot g, (1,\omega) \rangle.
\end{equation}
Under additional assumptions, the current $\dot g$ is a normal current, i.e. $\partial \dot g$ is also a scalar current, in particular it is a bounded measure on $(X,d)$.
\begin{lemma}
\label{L:normalcurr}
Assume that $q(y,\cdot) : \mathbb{R} \to \mathbb{R}$ belongs to $\text{\rm BV}(\mathbb{R})$ for $m$-a.e. $y$ and
\[
\sigma_y := - \frac{d}{dt} q(y,t), \quad \int_S |\sigma_y(\mathbb{R})| m(dy) = \int_S \text{\rm Tot.Var.}(q(y,\cdot) )m(dy) < + \infty.
\]
Then $\dot g$ is a normal current and its boundary is given by
\[
\langle \partial \dot g, \omega \rangle = \int_S \int_\mathbb{R} \omega(g(y,t)) \sigma_y(dt) m(dy).
\]
\end{lemma}
Note that in the above formula we cannot restrict $\sigma_y$ to $g^{-1}(\mathcal T)$: in fact, in general
\[
\int_S (g(y,\cdot)_\sharp \sigma_y)(\mathcal T_e \setminus \mathcal T) m(dy) > 0.
\]
\begin{proof}
First of all, by using the formula $q(y,t) = \sigma_y((t,+\infty))$, it follows that $\sigma_y$ is $m$-measurable, i.e. for all $\phi \in C_b(X,\mathbb{R})$ the integral
\[
\int \bigg( \int \phi(g(y,t)) \sigma_y(dt) \bigg) m(dy)
\]
is meaningful and then
\[
\int \big( g(y,\cdot)_\sharp \sigma_y \big) m(dy)
\]
is a finite measure on $(X,d)$.
A direct computation yields
\begin{align*}
\langle \partial \dot g, \omega \rangle = &~ \langle \dot g, (1,\omega) \rangle = \int_S \int_\mathbb{R} \partial_t \omega(g(t,y)) \sigma_y((t,+\infty)) dt m(dy) = \int_S \int_\mathbb{R} \omega(g(t,y)) \sigma_y(dt) m(dy).
\end{align*}
\end{proof}
\begin{remark}
\label{R:abscurr}
In many cases the measure $\int (g(y,\cdot)_\sharp \sigma_y) \llcorner_{\mathcal{T}} m(dy)$ is absolutely continuous w.r.t. $\eta$, i.e. for $m$-a.e. $y$
\[
\sigma_{y}\llcorner_{\mathcal{T}} = h(g(t,y)) q(y,t) \mathcal{L}^1.
\]
for some $h \in L^1(\eta)$. In that case we obtain that
\begin{align*}
\langle \partial \dot g, \omega \rangle =&~ \int \omega(b(y)) \sigma_y \big( P_2(\{g^{-1}(b(y))\}) \big) m(dy) \crcr
&~ - \int \omega(a(y)) \sigma_y \big( P_2(\{g^{-1}(a(y))\}) \big) m(dy) + \int \omega(x) h(x) \eta(dx).
\end{align*}
\end{remark}
\subsection{Transport equation}
\label{Ss:transpT}
We now consider the problem $\partial U = \mu - \nu$ in the sense of currents:
\[
\langle U, (1,\omega) \rangle = \langle \mu - \nu, \omega \rangle = \int \omega(x) (\mu - \nu)(dx).
\]
Using the disintegration formula and \eqref{E:muy}, \eqref{E:nuy} we can write
\[
\langle U, (1,\omega) \rangle = \int_S \bigg\{ \int_\mathbb{R} \omega(g(y,t)) (g^{-1}(y,\cdot)_\sharp \mu_y)(dt) - \int_\mathbb{R} \omega(g(y,t)) (g^{-1}(y,\cdot)_\sharp \nu_{y})(dt) \bigg\} m(dy).
\]
By integrating by parts we obtain
\begin{align*}
\int_\mathbb{R} \omega(g(y,t)) (g^{-1}(y,\cdot)_\sharp \mu_y)(dt) =&~ - \int_\mathbb{R} \mu_y(g(y,(-\infty,t))) \partial_t \omega(g(y,t)) dt = - \int_\mathbb{R} H(y,t) \partial_t \omega(g(y,t)) dt,
\end{align*}
\begin{align*}
\int_\mathbb{R} \omega(g(y,t)) (g^{-1}(y,\cdot)_\sharp \nu_y)(dt) = - \int_\mathbb{R} \nu_y(g(y,(-\infty,t))) \partial_t \omega(g(y,t)) dt = - \int_\mathbb{R} F(y,t) \partial_t \omega(g(y,t)) dt.
\end{align*}
Observe that the map
\[
S \times \mathbb{R} \ni (y,t) \mapsto F(y,t) - H(y,t) \in \mathbb{R}
\]
is in $L^1(m \otimes \mathcal L^1)$ if the transport cost $\mathcal{I}(\pi)$ is finite: in fact, using the fact that $F(y,t) \leq H(y,t)$ and integrating by parts,
\begin{equation}
\label{E:elle1}
\int_\mathbb{R} H(y,t) - F(y,t) dt = \int_\mathbb{R} (g^{-1}(y,\cdot)_\sharp \mu_y - g^{-1}(y,\cdot)_\sharp \nu_y)(-\infty,t) (dt) = \int_{\mathbb{R}^2} (t - s) \tilde \pi_{y}(ds,dt),
\end{equation}
where $\tilde \pi_{y}$ is the monotone rearrangement.
We deduce the following proposition.
\begin{proposition}
\label{P:bidual}
Under Assumption \ref{A:NDEatom}, a solution to $\partial U = \mu - \nu$ is given by the current $U$ defined as
\[
\langle U, (h,\omega) \rangle = \int_S \bigg( \int_\mathbb{R} (F(y,t) - H(y,t)) h(g(y,t)) \partial_t \omega(g(y,t)) dt \bigg) m(dy).
\]
\end{proposition}
In general, the solution is not unique: just add a boundary free current to our solution.
Some further assumptions allow to represent our solution $U$ as the product of a scalar $\rho$ with the current $\dot g$.
\begin{proposition}
\label{P:dualregu}
Assume that $q(y,t) > 0$ whenever $H(y,t) - F(y,t) > 0$. Then $R = \rho \dot g$, where
\[
\rho(g(y,t)) = \frac{F(y,t) - H(y,t)}{q(y,t)}.
\]
\end{proposition}
\begin{proof}
It is enough to observe that
\begin{align*}
\int_{S \times \mathbb{R}} F(y,t) - H(y,t) dt m(dy) =&~ \int_{S \times \mathbb{R}} \frac{F(y,t) - H(y,t)}{q(y,t)} q(y,t) dt m(dy) \crcr
=&~ \int_{S \times \mathbb{R}} \rho(g(y,t)) q(y,t) dt m(dy) = \int_X \rho(x) \eta(dx),
\end{align*}
and from \eqref{E:elle1} we conclude that $\rho \in L^1(\eta)$.
\end{proof}
\begin{corollary}
\label{C:regudual}
If $q(y,t) \not= 0$ for $m \otimes \mathcal{L}^1$-a.e. $(y,t) \in g^{-1}(\mathcal{T})$, then there exists a scalar function $\rho$ such that $\partial (\rho \dot g) = \mu - \nu$.
\end{corollary}
\section{Stability of the non degeneracy condition}
\label{S:limite}
In this section we prove a general approximation theorem, which will be then applied to the Measure-Gromov-Hausdorff (MGH) convergence: if a uniform estimate holds for the disintegration in the approximating spaces, we deduce the regularity of the disintegration also in the limit.
\subsection{A general stability result}
\label{Ss:genstb}
We consider the following setting:
\begin{enumerate}
\item $\mu_n$ is a sequence of measure converging to $\mu$ weakly;
\item there exists functions $g_n : S_n \times \mathbb{R} \to X$, $S_n \subset X$ Borel, and measures $r_n m_n \otimes \mathcal{L}^1 \in \mathcal P(S_n \times \mathbb{R})$ such that
\begin{equation}\label{enum:erren}
\mu_n = (g_n)_\sharp \big( r_n m_n \otimes \mathcal{L}^1 \big).
\end{equation}
\end{enumerate}
The following is the basic tool for our stability result.
\begin{proposition}
\label{P:limite}
Let $Y$ be a Polish space, $\{\xi_n\}_{n \in \mathbb{N}} \subset \mathcal{P}(Y)$ such that $\xi_n \rightharpoonup \xi$.
Consider $\{r_n\}_{n \in \mathbb{N}}$, $r_n \geq 0$, such that $r_n \in L^{1}( \xi_n )$, $r_n \xi_n \rightharpoonup \zeta$ and the following equintegrability condition holds:
\[
\forall \varepsilon > 0 \ \exists \delta >0 \ \bigg( \forall A \in \mathcal B, \xi_n(A) < \delta \quad \Longrightarrow \quad \int_A r_n \xi_n < \varepsilon \bigg).
\]
Then there exists $r \in L^1(\xi)$ such that $\zeta= r \xi$.
\end{proposition}
\begin{proof}
We will show that $\zeta(B) = 0$ for all $B$ such that $\xi(B) = 0$. Clearly by inner and outer regularity, it is enough to prove the following statement:
\[
\forall \varepsilon > 0 \ \exists \delta >0 \ \bigg( \phi \in C_{b}(Y), \phi\geq 0, \ \int \phi \xi < \delta \quad \Longrightarrow \int \phi \zeta < \varepsilon \bigg).
\]
Fix $\varepsilon > 0$ and take the corresponding $\delta$ given by the equintegrability condition on $r_n$. Clearly w.l.o.g. $\delta\leq\varepsilon$.
Consider $\phi \in C_b(Y)$ positive such that
\[
\int \phi \xi \leq \delta^2/2.
\]
From the weak convergence for $n$ great enough
\[
\int \phi \xi_n \leq \delta^2,
\]
so that we can estimate
\[
\int \phi r_n \xi_n \leq \int_{\phi > \delta} r_n \xi_n + \delta < \varepsilon+\delta .
\]
Hence $\int \phi \zeta < 2 \varepsilon$.
\end{proof}
\begin{theorem}
\label{T:apprograph}
Assume that the family of functions $\{r_n\} \subset L^1(m_n \otimes \mathcal{L}^1)$ given by \eqref{enum:erren} is such that
\[
(\mathbb{I},\mathbb{I},g_n)_\sharp \big( r_n m_n \otimes \mathcal{L}^1 \big) \rightharpoonup (\mathbb{I},\mathbb{I},g)_\sharp \zeta
\]
with $\zeta \in \mathcal{P}(S \times \mathbb{R})$ and $g$ being the ray map (Definition \ref{D:mongemap}). Assume moreover
\[
\forall T\geq 0 \ \forall \varepsilon > 0 \ \exists \delta >0 \ \bigg( A \in \mathcal B(S \times[-T,T]), m_{n}\otimes\mathcal{L}^{1}(A) < \delta \quad \Longrightarrow \quad
\int_A r_n m_n\otimes\mathcal{L}^{1} < \varepsilon \bigg).
\]
Then $\zeta = r m \otimes \mathcal{L}^1$ for some function $r \in L^1(m \otimes \mathcal{L}^1)$, measure $m \in \mathcal{P}(S)$ and the disintegration of $\mu$ is a.c. w.r.t. $\mathcal H^1$ on each geodesic.
\end{theorem}
\begin{proof}
Define for $k \in \mathbb{N}$
\[
\phi_{k} \in C_{c}(\mathbb{R}),\ \phi_{k}\geq 0, \quad
\phi_{k}(t):=
\begin{cases}
1 & |t| \leq k, \crcr
0 & |t| \geq k+1.
\end{cases}
\]
Let $\xi_{n,k}=m_{n}\otimes\mathcal{L}^{1}\llcorner_{[-k-1,k+1]}$ and consider the functions $\tilde r_{n,k}:= r_{n}(y,t)\phi_{k}(t)$.
Since $m_{n}=(P_{1})_{\sharp}(r_{n}m_{n}\otimes\mathcal{L}^{1})$ and hence $m_{n}\rightharpoonup m = (P_{1})_{\sharp}\zeta$, then
\[
\xi_{n,k} \rightharpoonup m\otimes \mathcal{L}^{1}\llcorner_{[-k-1,k+1]}
\]
and the hypothesis of Proposition \ref{P:limite} are verified up to rescaling. So $\zeta = r m \otimes \mathcal{L}^1$.
The fact that $g_\sharp \zeta$ is a disintegration is a consequence of the a.c. of $\zeta$ along each geodesic: in this case the initial points have $\zeta$-measure $0$
and therefore $g$ is invertible on a set of full $\mu$-measure.
\end{proof}
In general the convergence of the graph of $g_n$ is too strong: the next result considers a more general case.
\begin{proposition}
\label{P:finalapr}
Assume that $\tilde \zeta \in \Pi(r m \otimes \mathcal{L}^1, \mu)$ is concentrated on the graph of a Borel function $h : \mathcal{T} \times \mathbb{R} \to \mathcal{T}_e$ such that
\begin{enumerate}
\item \label{Cond:1final} $(y,t) \mapsto e(y) := f(h(y,t)) \in S$ is constant w.r.t. $t$,
\item it holds
\[
h(y,\cdot)_\sharp \big( r(y,\cdot) \mathcal{L}^1 \big) \ll \mathcal{H}^1 \llcorner_{g(e(y),\mathbb{R})}.
\]
\end{enumerate}
Then the disintegration w.r.t. $g$ has absolutely continuous conditional probability.
\end{proposition}
\begin{proof}
We can disintegrate the measure $m$ as follows:
\[
m = \int_S m_z (e_\sharp m)(dz),
\]
and by the second assumption
\[
h(y,\cdot)_\sharp (r(y,\cdot) \mathcal{L}^1) = g(e(y),\cdot)_\sharp (\tilde r(y,\cdot) \mathcal{L}^1),
\]
for $m$-a.e. $y \in \mathcal{T}$. Hence by explicit computation,
\begin{align*}
\mu =&~ \int_{S} h(y,\cdot)_{\sharp}( r(y,\cdot) \mathcal{L}^{1} ) m(dy) = \int_{S} g(e(y),\cdot)_{\sharp}( \tilde r(y,\cdot) \mathcal{L}^{1} ) m(dy) \crcr
=&~ \int_{S} \bigg( \int_{e^{-1}(z)} g(z,\cdot)_{\sharp} (\tilde r(y,\cdot) \mathcal{L}^{1}) m_{z}(dy)\bigg) e_{\sharp}m(dz).
\end{align*}
To conclude the proof observe that
\begin{align*}
\int_{e^{-1}(z)} g(z,\cdot)_{\sharp} (\tilde r(y,\cdot) \mathcal{L}^{1}) m_{z}(dy) = &~
g(z,\cdot)_{\sharp}\bigg( \int_{e^{-1}(z)} \tilde r(y,\cdot) \mathcal{L}^{1} m_{z}(dy) \bigg) \crcr
=&~ g(z,\cdot)_{\sharp}\bigg( \int_{e^{-1}(z)} \tilde r(y,\cdot) m_{z}(dy) \bigg) \mathcal{L}^{1}.
\end{align*}
\end{proof}
\begin{remark}
\label{R:morereg}
Observe that some properties of $r_n$ are preserved passing to the limit $r$.
In relation with the previous section, we consider the following cases: for $A \subset X \times \mathbb{R}$ open
\begin{enumerate}
\item for some $\varepsilon>0$
\[
\big((r_n -\varepsilon)m_{n}\otimes \mathcal{L}^{1}\big)\llcorner_{A} \geq0;
\]
\item there exists $L> 0$ such that
\[
r_n(y,\cdot) \in \textrm{Lip}_{L}(A_{y});
\]
\item there exists $M>0$ such that
\[
TV( r_n(y,\cdot)\llcorner_{A}) \leq M.
\]
\end{enumerate}
The first condition yields that the assumptions of Corollary \ref{C:regudual} holds in $A$.
The second and third conditions imply that we are under the conditions for Remark \ref{R:abscurr} in $A$.
\end{remark}
\subsection{Approximations by metric spaces}
\label{Ss:GHC}
In this section we explain a procedure to verify if the transport problem under consideration satisfies Assumption \ref{A:NDE}. The basic references for this sections are \cite{villott:curv} and \cite{sturm:MGH1,sturm:MGH2}.
We consider the following setting:
\begin{enumerate}
\item $(X,d,d_{L})$, $(X_n,d_n,d_{L,n})$, $n \in \mathbb{N}$, are metric structures satisfying the assumptions of page \pageref{P:assumpDL} and Remark \ref{R:compact}: more precisely,
$d_{L},d_{L,n}$ l.s.c., $d_{L}\geq d, d_{L,n}\geq d_{n}$ and
\[
\bigcup_{x \in K_1, y \in K_2} \gamma_{[x,y]}\ \ \textrm{is $d_{n}$($d$)-compact if $K_1$, $K_2$ are $d_{n}$($d$)-compact, $d_{L,n}(d_{L}) \llcorner_{K_1 \times K_2}$ uniformly bounded}.
\]
\item $\mu_n, \nu_n \in \mathcal{P}(X_n)$, $\mu_n \perp \nu_n$;
\item $\pi_n \in \Pi(\mu_n,\nu_n)$ is a $d_{L,n}$-cyclically monotone transference plan with finite cost.
\end{enumerate}
For $\mu,\nu \in \mathcal{P}(X)$ let $\pi \in \Pi(\mu,\nu)$ be a generic transference plan.
\begin{definition}\label{D:mgh}
We say that the structures $(X_n,d_n,d_{L,n}, \pi_n)$ \emph{converge} to $(X,d,d_L,\pi)$ if the following holds:
there exists $C>0$ such that for all $n \in \mathbb{N}$
\begin{equation*}
\int d_{L,n} \pi_{n}\leq C
\end{equation*}
and there exist Borel sets $A_n \subset X_n$ and Borel maps $\ell_n : A_n \to X$ such that
\begin{equation}
\label{E:conMHD0}
(\ell_n\otimes \ell_n)_\sharp \pi_n \llcorner_{A_n \times A_n} \rightharpoonup \pi,
\end{equation}
\begin{equation}
\label{E:conMHD}
\big| d_L(\ell_n(x),\ell_n(y)) - d_{L,n}(x,y) \big| \leq 2^{-n},
\end{equation}
and if $(\ell_n(x_n),\ell_n(y_n)) \to (x,y)$, then
\begin{equation}
\label{E:conMHD1}
d_L(x,y) = \lim_n d_{L,n}(x_n,y_n).
\end{equation}
\end{definition}
As a first result, we show that also $\pi$ is $d_{L}$-cyclically monotone with finite cost.
\begin{proposition}\label{P:bddmono}
If $(X_n,d_n,d_{L,n},\pi_n)$ converges to $(X,d,d_L,\pi)$ and the plans $\pi_{n}$ have uniformly bounded cost then
also $\pi$ has finite cost and is $d_{L}$-cyclically monotone.
\end{proposition}
\begin{proof}
Since $d_{L}$ is l.s.c.
\begin{align*}
\int d_{L} \pi \leq~& \liminf_{n\to +\infty} \int d_{L} (\ell_{n}\otimes \ell_{n})_{\sharp}\pi_{n} =
\liminf_{n\to +\infty} \int d_{L}(\ell_{n}(x), \ell_{n}(y)) \pi_{n}(dxdy) \crcr
\stackrel{\textrm{\eqref{E:conMHD}}}{\leq}& \liminf_{n\to +\infty} \bigg\{ \int d_{L,n}(x,y) \pi_{n}(dxdy) + 2^{-n} \bigg\} \leq C,
\end{align*}
for some $C<+\infty$.
Now let $\Gamma_{n}$ be a $d_{L,n}$-cyclically monotone set with $\pi_{n}(\Gamma_{n})=1$: by standard regularity of Borel function and by Prokhorov Theorem
we can assume that
\begin{enumerate}
\item $\Gamma_{n}$ is $\sigma$-compact, $\Gamma_{n}=\cup_{m\in \mathbb{N}} \Gamma_{n,m}$ with $\Gamma_{n,m}\subset \Gamma_{n,m+1}$;
\item $(\ell_{n}\otimes \ell_{n})(\Gamma_{n,m})$ is compact and $(\ell_{n}\otimes \ell_{n})(\Gamma_{n,m}) \to \Gamma_{m}$
in the Hausdorff distance $d_{H}$;
\item $\pi_{n}(\Gamma_{n,m})\geq 1- 2^{-m}$.
\end{enumerate}
It follows that: $\pi(\Gamma_{m})\geq 1 - 2^{-m}$, hence
\[
\pi \bigg(\bigcup_{m \in \mathbb{N}}\Gamma_{m} \bigg)=1.
\]
Since each $\Gamma_{m}$ is the limit in Hausdorff distance of $(\ell_{n}\otimes \ell_{n})(\Gamma_{n,m})$,
\eqref{E:conMHD1} implies that $\Gamma_{m}$ (and thus $\cup_{m}\Gamma_{m}$, because $\Gamma_{m}\subset \Gamma_{m+1}$)
is $d_{L}$-cyclically monotone.
\end{proof}
Note that since $\pi$ is $d_{L}$-cyclically monotone, we can define
the sets $\Gamma, \Gamma', G, G^{-1}, R, a, b$ of Section \ref{S:Optimal} as well as
the quotient map $f$ and the ray map $g$ constructed in Section \ref{S:partition}.
The same sets and maps can be given for the structures $(X_{n},d_{n},d_{L,n})$: we
will denote them with the subscript $n$.
For the transport problems in $(X_n,d_n)$ with measures $\mu_n$, $\nu_n$, we assume the following.
\begin{assumption}[Non degeneracy]
\label{A:punti}
The $d_{L,n}$-cyclically monotone plan $\pi_n$ satisfies Assumption \ref{A:NDE} for all $n \in \mathbb{N}$.
\end{assumption}
This allows to write the disintegration of $\mu_n$ w.r.t.
the ray equivalence relation $R_{n}$:
\[
\mu_n = (g_n)_\sharp (r_n m_{n} \otimes \mathcal L^1) = \int g_n(y,\cdot)_\sharp (r_{n}(y,\cdot) \mathcal L^1) m_{n}(dy),
\]
with $f_{n\,\sharp}\mu_{n}=m_{n}$ and $r_n \in L^1(m_{n} \otimes \mathcal L^1)$.
\begin{lemma}\label{L:gnid}
If $(X_n,d_n,d_{L,n},\pi_n)$ converges to $(X,d,d_L,\pi)$ then the structures $(S_{n}\times \mathbb{R},\tilde d_{n}, \tilde d_{L,n}, \tilde \pi_{n} )$, where
\[
\tilde d_{n} = d_{n} \circ (g_{n}\otimes g_{n}), \quad \tilde \pi_{n}=(g_{n}^{-1}\otimes g_{n}^{-1})_{\sharp} (\pi_{n}),
\quad \tilde d_{L,n}((y,t),(y',t')) =
\begin{cases}
|t-t'| & y=y' \crcr
+\infty & y\neq y',
\end{cases}
\]
converges to $(X,d,d_{L},\pi)$.
\end{lemma}
\begin{proof}
It is enough to observe that $\pi_{n}(G_{n})=1$, $\tilde d_{L,n} = d_{L,n} \circ (g_{n}\otimes g_{n})$ on $G_{n}$ and to replace the map $\ell_{n}$
with the map $\ell_{n}\circ g_{n}$.
\end{proof}
By Lemma \ref{L:gnid}, in the following we assume that the ray map $g_{n}$ is the identity map.
The next assumption is the fundamental one.
\begin{assumption}[Equintegrability]
\label{A:equi}
The $L^{1}$-functions $r_n$ are equintegrable w.r.t. the measure $m_n \otimes \mathcal L^1$:
\[
\forall \varepsilon>0 \ \exists \delta>0 \bigg( (m_n \otimes \mathcal L^1)(A) < \delta \ \Rightarrow \ \int_A r_n m_n \otimes \mathcal L^1 < \varepsilon \bigg).
\]
\end{assumption}
From now on we will assume that $(X_n,d_n,d_{L,n}, \pi_n) \to (X,d,d_L,\pi)$ in the sense of Definition \ref{D:mgh},
$(X_n,d_n,d_{L,n}, \pi_n)$ verifies Assumption \ref{A:punti} and Assumption \ref{A:equi}.
Our aim is to prove that the structure $(X,d,d_L,\pi)$ satisfies Assumption \ref{A:NDE}, which is equivalent to the fact that
the marginal probabilities of the disintegration of $\mu$ w.r.t the ray equivalence relation $R$ are a.c. w.r.t. $\mathcal{H}^{1}$.
The next lemma shows that in order to obtain our purpose we can perform some reductions without losing generality.
We will write $\mu_{k}\nearrow \mu$ for $\mu_{k}\leq \mu_{k+1}$ and $\mu = \sup_{k} \mu_{k}$.
\begin{lemma}\label{L:approx}
Let $\{\mu_{k}\}_{k\in\mathbb{N}} \subset \mathcal{M}(X)$, $\mu_k \geq 0$, be such that $\mu_{k} \nearrow \mu$ and assume that
\[
\mu_{k}=g_{\sharp}( r_{k}m_{k}\otimes \mathcal{L}^{1}), \quad r_{k}\geq0,
\]
where $g$ is the ray map on $\mathcal{T}$. Then there exist $m \in \mathcal{P}(X)$, $r \in L^{1}(m\otimes \mathcal{L}^{1})$, $r\geq0$ such that the same formula holds for $\mu$:
\[
\mu=g_{\sharp}( r m \otimes \mathcal{L}^{1}).
\]
\end{lemma}
\begin{proof}
Since $\int r_{k}(y,t)dt=1$ it follows that $P_{1\,\sharp}(r_{k}m_{k}\otimes \mathcal{L}^{1})= m_{k}$ and therefore $m_{k} \nearrow m$ with
$m=f_{\sharp}\mu$ (recall that $f$ is a section for the ray equivalence relation $R$).
The convergence $\mu_{k}\nearrow \mu$ yields
\[
\bigg( r_{k}\frac{dm_{k}}{d m} \bigg) m\otimes \mathcal{L}^{1} \nearrow \zeta,
\]
where $\mu=g_{\sharp}\zeta$. We conclude $\zeta =r m\otimes \mathcal{L}^{1}$ with $r:= \sup_{k} r_{k}\frac{dm_{k}}{d m}$.
\end{proof}
A first reduction is given by the following lemma.
\begin{lemma} \label{L:ptdist}
To prove that there exist $m \in \mathcal{P}(X)$, $r \in L^{1}(m\otimes \mathcal{L}^{1})$, $r\geq0$ such that
\[
\mu=g_{\sharp}( r m \otimes \mathcal{L}^{1}),
\]
we can assume w.l.o.g. that there exist $\bar x, \bar y \in X$ and $q \geq0$ such that
\[
\pi \Big( \Big\{ (x,y): d(\bar x,\bar y) > 8q, d(x,\bar x), d(y,\bar y) \leq q \Big\} \Big) =1.
\]
Moreover the $d_{L}$-cyclically monotone set $\Gamma$ and the set of oriented transport rays $G$ can be assumed to be compact subsets of $X\times X$.
\end{lemma}
\begin{proof}
{\it Step 1.}
Since $\pi(\{x=y \})=0$ we can assume that $\Gamma \cap \{x=y \}= \emptyset$.
Take two dense sequences $\{x_{i}\}_{i\in \mathbb{N}} \subset X$, $\{q_{i} \}_{i\in \mathbb{N}}\subset \mathbb{R}^+$ and consider the family of closed sets
\[
\Gamma_{ijk}:=\Big\{ (x,y): d( x_{i},x_{j}) \geq 8q_{k}, d(x,x_{i}), d(y,x_{j}) \leq q_{k} \Big\}.
\]
Then $\Gamma_{ijk}$ is a countable covering of $X\times X \setminus \{x=y\}$.
Suppose now to have proven that for all $\mu_{ijk} = P_{1\,\sharp}( \pi\llcorner_{\Gamma_{ijk}})$ the disintegration formula holds with
$\mathcal H^{1}$-a.c. marginal probabilities, then the same $\mathcal H^1$-a.c. property is true if we replace $\Gamma_{ijk}$ with the finite union of
sets $\Gamma_{i'j'k'}$.
Define
\[
\tilde \Gamma_{m} := \bigcup_{n < m} \Gamma_{i_{n}j_{m}k_{m}},
\]
where
\[
\mathbb{N} \ni m \mapsto (i_m,j_m,k_m) \in \mathbb{N}^3
\]
is a bijective map, and consider $\mu_{m} = P_{1\,\sharp}(\pi\llcorner_{\tilde \Gamma_{m}})$, then $\{\mu_{m}\}_{m\in\mathbb{N}}$
verifies the hypothesis of Lemma \ref{L:approx}.
{\it Step 2.} It remains to show how to construct the approximating structure $\tilde \pi_{n} \in \mathcal{P}(X_{n})$
converging in the sense of Definition \ref{D:mgh} to $\pi \llcorner_{\Gamma_{ijk}}$.
Since $\Gamma_{ijk}$ is closed, there exists a sequence
$\phi_{l} \in C_{c}(X\times X,[0,1])$ such that $\phi_{l}\searrow \chi_{\Gamma_{ijk}}$.
Now $\phi_{l} \pi \searrow \pi\llcorner_{\Gamma_{ijk}}$ as $l\to +\infty$ and
\[
\phi_{l} (\ell_{n}\otimes \ell_{n} )_{\sharp}\pi_{n} \rightharpoonup \phi_{l} \pi.
\]
Hence there exists a subsequence $\{\phi_{l_i} (\ell_{n_i}\otimes \ell_{n_i} )_{\sharp}\pi_{n_i} \}_{i \in \mathbb{N}}$ satisfying \eqref{E:conMHD0} with weak limit $\pi\llcorner_{\Gamma_{ijk}}$. If one defines
\[
\tilde \pi_{i} = \big( \phi_{l_i} \circ \ell_{n_i} \big) \pi_{n_i},
\]
then it is straightforward to show that $(X_i,d_i,d_{L,i},\tilde \pi_i)$ converges to $(X,d,d_L,\pi\llcorner_{\Gamma_{ijk}})$ in the sense of Definition \ref{D:mgh}.
{\it Step 3.}
Since Remark \ref{R:lsccase1} yields that $\Gamma$, $G$ are $\sigma$-compact, let $\Gamma= \cup_{k}\Gamma_{k}$, $G=\cup_{k}G_{k}$
with $\Gamma_{k}$, $G_{k}$ compact and consider $\pi\llcorner_{\Gamma_{k}}$.
The same reasoning done in Step 1 and Step 2. yields that it is enough to prove the a.c. of disintegration for $\pi\llcorner_{\Gamma_{k}}$.
\end{proof}
Therefore from now on we will assume that $\pi$ is concentrated on the set
\[
\Big\{ (x,y): d(\bar x,\bar y) > 8q, d(x,\bar x), d(y,\bar y) \leq q \Big\}.
\]
Using the same reasoning of Lemma \ref{L:ptdist} one can also prove the following.
\begin{lemma} \label{L:cpt}
We can assume w.l.o.g. that the sets $A_{n}\subset X_{n}$ are compact and the maps $\ell_{n}:A_{n} \to X$ are continuous.
Moreover $\ell_{n}(A_{n})$ converges in Hausdorff distance to a compact set $K$ on which $\mu$ and $\nu$ are concentrated.
\end{lemma}
\begin{proof}
By Lusin Theorem and inner regularity of measures it follows that there exist $B_{n}\subset A_{n}$ such that
\begin{itemize}
\item $A_{n}\setminus B_{n}$ is compact;
\item $\mu_{n}(B_{n})\leq 1/n$;
\item the map $\ell_{n}: A_{n}\setminus B_{n} \to X$ is continuous.
\end{itemize}
To prove the first part of the claim just observe that $(\ell_{n} \otimes \ell_{n} )_{\sharp} \pi_{n}\llcorner_{A_{n}\setminus B_{n} \times A_{n}\setminus B_{n}} \rightharpoonup \pi$.
The second part of the statement can be proven following the line of the second part of the proof of the Proposition \ref{P:bddmono}.
\end{proof}
By Lemma \ref{L:cpt} it is straightforward that for all $n$ great enough we have
\[
\big(\ell_{n\,\sharp}\mu_{n}\big)(B_{2q}(\bar x) )= \big(\ell_{n\,\sharp}\nu_{n}\big)(B_{2q}(\bar y) ) =1.
\]
\begin{lemma}\label{L:rette}
We can assume that the measure $m_{n}$ is concentrated on a compact subset of
\[
\Big\{ y \in S_{n} : \exists t,s>\delta : (y,-s),(y,0),(y,t) \in P_{12}( \textrm{\rm graph} (\ell_{n}) ) \Big\}
\]
for some fixed $\delta>0$ and
\begin{equation}\label{E:rette}
\mu_{n}( S_{n}\times (-\infty,0]) =1, \quad \nu_{n}( S_{n}\times (4q,+\infty)) =1.
\end{equation}
\end{lemma}
\begin{proof}
{\it Step 1.} Defining
\[
A_\delta := \Big\{ (y,t)\in S_{n} \times \mathbb{R} : |a(y)-t|<\delta \Big\},
\]
by Fubini Theorem
\[
m_{n}\otimes\mathcal{L}^{1}(A_{\delta})
\leq \delta,
\]
hence by Assumption \ref{A:equi}, for any $\varepsilon>0$ we can choose $\delta>0$ such that
$r_{n} m_{n} \otimes \mathcal{L}^{1}(A_{\delta})<\varepsilon$ for al $n \in \mathbb{N}$.
Therefore we can assume that $r_{n} m_{n} \otimes \mathcal{L}^{1}$ is concentrated on a compact subset $B_{n}$
of $\ell_{n}^{-1}(\bar B (\bar x, \frac{3}{2}q) ) \setminus A_{\delta}$.
{\it Step 2.}
Define the u.s.c. selection of $B_{n}$ $t_{n}: S_{n} \to \mathbb{R} $ in the following way:
\[
y \mapsto t_{n}(y):= \max \big\{ t \in \mathbb{R}, (y,t) \in B_{n} \big\}.
\]
By removing a set of arbitrarily small measure we can assume that for all $y \in P_{1}(\textrm{graph} (t_{n}))$ there exists $t>4q$ such that
\[
(y,t_{n}(y)+t) \in \ell_{n}^{-1}\bigg( \bar B\bigg(\bar y, \frac{3}{2}q \bigg) \bigg).
\]
{\it Step 3.}
The Borel transformation
\begin{align*}
B_{n} \ni (y,t) \mapsto (y, t-t_{n}(y))
\end{align*}
maps $m_{n}\otimes \mathcal{L}^1$ into itself and in the new coordinates the section $S_{n}$ satisfies the first part of the claim.
By the definition of $G_{n}$ and $\mu_{n}\perp\nu_{n}$ it follows that $\mu_{n}$ and $\nu_{n}$ satisfy \eqref{E:rette}, see Remark \ref{R:monot}.
\end{proof}
Define the map
\[
\begin{array}{ccccc}
h_n &:& S_{n} \times \mathbb{R} &\to& X \times \mathbb{R} \crcr
&& (y,t) &\mapsto& (\ell_{n} (y,0),t) .
\end{array}
\]
and the measure $h_{n\,\sharp} ( r_{n}m_{n}\otimes \mathcal{L}^{1} ) = \tilde r_{n} \tilde m_{n} \otimes \mathcal{L}^{1}$,
with $\tilde m_{n} = \ell_{n}(\cdot,0)_{\sharp} m_{n}$.
\begin{lemma} \label{L:tight}
The family of measures $\{\tilde r_{n} \tilde m_{n} \otimes \mathcal{L}^{1}\}_{n\in \mathbb{N}} \subset \mathcal{P}( \ell_{n}(S_{n}\times \{0\}) \times \mathbb{R})$ is tight and $\tilde r_{n}$
is equintegrabile w.r.t. $\tilde m_{n} \otimes \mathcal{L}^{1}$.
\end{lemma}
\begin{proof}
Performing the same calculation of \eqref{E:costo}
\[
C\geq \int d_{L,n} \pi_{n} = \int s \nu_{n} - \int s \tilde r_{n} \tilde m_{n}\otimes \mathcal{L}^{1}.
\]
From \eqref{E:rette}, Lemma \ref{L:rette}, it follows that $s \leq 0$, $\tilde r_{n} \tilde m_{n}\otimes \mathcal{L}^{1}$-a.e..
Hence $\tilde r_{n} \tilde m_{n}\otimes \mathcal{L}^{1} \in \mathcal{P}( \ell_{n}(S_{n}\times \{0\}) \times (-\infty,0]))$ and
\[
0 \leq - \int s \tilde r_{n} \tilde m_{n}\otimes \mathcal{L}^{1} \leq C,
\]
therefore $\tilde r_{n} \tilde m_{n}\otimes \mathcal{L}^{1}$ is tight. Recall in fact that $\{S_n\}_{n \in \mathbb{N}}$ is a precompact sequence w.r.t. the Hausdorff distance by Lemma \ref{L:cpt}.
The equintegrability is straightforward:
\[
\int_{A} \tilde r_{n} \tilde m_{n}\otimes \mathcal{L}^{1} = \int_{(h_{n})^{-1}(A)} r_{n} m_{n}\otimes \mathcal{L}^{1}
\]
and $m_{n}\otimes \mathcal{L}^{1}((h_{n})^{-1}(A)) = \tilde m_{n}\otimes \mathcal{L}^{1} (A)$.
\end{proof}
Consider the following measure
\[
\zeta_n := (h_{n}, \ell_{n})_\sharp ( r_n m_n \otimes \mathcal L^1) \in \Pi \Big( \tilde r_n \tilde m_n \otimes \mathcal L^1,
(\ell_n )_{\sharp} (\mu_{n}) \Big) \in \mathcal{P}(X\times \mathbb{R} \times X).
\]
\begin{proposition}
\label{P:graphnn}
Up to subsequences, $\zeta_n \rightharpoonup \zeta$, where $\zeta \in \Pi(r m \otimes \mathcal L^1, \mu)$ is supported on a Borel graph $h : \mathcal T \times \mathbb{R} \to \mathcal T_e$ such that $t \mapsto h(y,t)$ is the $d_L$ $1$-Lipschitz curve $R(y)$ for $m$-a.e. $y \in X$.
\end{proposition}
\begin{proof}
{\it Step 1.}
The convergence to the correct marginals is a consequence of \eqref{E:conMHD0}
\[
(P_2)_\sharp \zeta_n = (\ell_n \circ g_n)_\sharp (r_n m_n \otimes \mathcal L^1) =
(\ell_n)_\sharp \mu_n \rightharpoonup \mu,
\]
and by Lemma \ref{L:tight}
\[
(P_1)_\sharp \zeta_n= \tilde r_n \tilde m_n \otimes \mathcal L^1 \rightharpoonup r m \otimes \mathcal L^{1}.
\]
{\it Step 2.} Since up to subsequence $\zeta_{n} \rightharpoonup \zeta$, using the same technique of Lemma \ref{P:bddmono},
we can assume that $K_{n}:= (h_{n},\ell_{n})(S_{n}\times \mathbb{R})$ is compact and $d_{H}(K_{n}, \textrm{graph}(h)) \to 0$ where $\textrm{graph}(h)$ is a compact set supporting $\zeta$ and $h$ is the associated multivalued function.
{\it Step 3.}
Let $(y,t,x) \in \textrm{graph}(h)$, then by the definition of convergence in the Hausdorff metric,
there exists a sequence $(\ell_{n}(y_{n},0), t_{n},\ell_{n} (y_{n},t_{n})) \to (y,t,x)$.
Hence from
\[
d_{L,n}\big( (y_{n},t_{n}),(y_{n},0) \big)=|t_{n}| \to |t|,
\]
we deduce by \eqref{E:conMHD1} that $d_{L}(x,y)=|t|$. In particular this implies that if $t=0$ then $x=y$.
{\it Step 4.} Let $(y,t,x), (y,t',x') \in \textrm{graph}(h)$ with $t < 0$ and $t' > 0$.
Again by the Hausdorff convergence there exist two sequences satisfying
\[
\big(\ell_{n}(y_{n},0), t_{n},\ell_{n}(y_{n},t_{n})\big) \to(y,t,x), \quad\big(\ell_{n}(y'_{n},0), t'_{n}, \ell_{n}(y'_{n},t'_{n})\big) \to(y,t',x').
\]
Since $d_{L}(y,y)=0$, from \eqref{E:conMHD1} we deduce
\[
d_{L,n}\big((y_{n},0),(y'_{n},0)\big) \to 0
\]
hence by the definition of $d_{L,n}$, for $n$ great enough $y_{n}=y'_{n}$.
Therefore
\[
d_{L}(x,x')= \lim_{n \to + \infty} d_{L,n}((y_{n},t_{n}),(y'_{n},t'_{n})) = |t|+t',
\]
and by Step 3 we conclude that $d_{L}(x,x')=d_{L}(x,y)+d_{L}(y,x')$.
{\it Step 5.} Let $(y,t,x) \in \textrm{graph}(h)$: we now show that
\[
t\geq0\ \Rightarrow\ (y,x)\in G, \quad -t\geq0\ \Rightarrow\ (y,x)\in G^{-1}.
\]
We will prove only the first implication for $t>0$.
Since following Lemma \ref{L:cpt} we can take $G_{n}$ compact such that
\begin{enumerate}
\item $(\ell_{n}\otimes \ell_{n})(G_{n}) \to \hat G$ in the Hausdorff metric;
\item $\hat G \subset G$,
\end{enumerate}
it is enough to show that there exists a sequence $(\ell_{n}(y_{n},0),t_{n},\ell (y_{n},t_{n}))\to (y,t,x)$
so that $(y_{n},x_{n}) \in G_{n}$ for all $n$, but this last implication is straightforward.
{\it Step 6.} We next show that for any $y \in P_{1}(\textrm{graph} (h))$ there exist $t_{-},t_{+} \geq \delta$
and $x_{-},x_{+}$ such that $(y,-t_{-},x_{-}),(y,t_{+},x_{+}) \in \textrm{graph}(h)$.
In fact
we recall that for all $y_{n} \in S_{n}$ there exist $t_{-,n},t_{+,n}\geq \delta$,
for some strictly positive constant $\delta$,
such that
\[
\big( (y_{n}, -t_{-,n}),(y_{n},0) \big), \big((y_{n},0),(y_{n},t_{+,n})\big) \in G_{n}.
\]
Hence chose $y_{n} \in S_{n}$ such that $\ell_{n}(y_{n})\to y$
and pass to converging subsequences to obtain the claim.
{\it Step 7.} Since for $y \in P_{1}(\textrm{graph}(h))$ there exist $x,x'$ such that $(x,y), (y,x')\in G \setminus \{x=y\}$, then $(x,x')\in G$, $y \in \mathcal{T}$
and $h$ is single valued.
The same computation of Point 5 yields that
\[
\big\{ (y,h(t,y)), t\geq 0\big\} \cup \big\{ (h(t,y),y), t \leq 0\big\}\subset G,
\]
and from this it follows that $h(y,\mathbb{R})\subset R(y)$.
Again from Point 5 one obtains that $d_{L}(y,h(t,y))=|t|$ and therefore
$t\mapsto g^{-1}(h(y,t))= g^{-1}(y)+t$.
\end{proof}
\begin{theorem}
\label{T:aproxxgg}
Let $(X_{n},d_{n},d_{L,n},\pi_{n}) \to (X,d,d_{L},\pi)$ and $(X_{n},d_{n},d_{L,n},\pi_{n})$, $n \in \mathbb{N}$,
verifies Assumption \ref{A:punti} and Assumption \ref{A:equi}. Then the marginal measure $\mu = P_{1\,\sharp}(\pi)$ satisfies Assumption \ref{A:NDE}.
\end{theorem}
\begin{proof}
The measure $\zeta$ constructed in the Proposition \ref{P:graphnn} satisfies the hypothesis of Proposition \ref{P:finalapr}.
Therefore the marginal probabilities of the disintegration of $\mu$ are absolutely continuous with respect to $\mathcal{H}^{1}$ and therefore $\mu$
verifies Assumption \ref{A:NDE}.
\end{proof}
\begin{remark}
\label{R:reguadd}
As in Remark \ref{R:morereg}, if we know more regularity of the disintegrations for the approximating problems, we can pass them to the limit. Here the key observation is that geodesics converge to geodesics, so uniform continuous functions on them converge pointwise to continuous functions.
\end{remark}
A special case is when $d_L = d$: a natural approximation is by transport plans where $\nu$ is atomic, with a finite number of atoms.
This case can be studied with more standard techniques, we refer to the analysis contained in \cite{biaglo:HJ1}.
\section{Applications}
\label{S:mcp}
In this section we recall the definition of Measure Contraction Property ($MCP$) and then we
prove that for a metric measure space $(X,d,\eta)$ satisfying $MCP$, the Monge minimization problem with
marginal measures
$\mu$ and $\nu$ with $\mu \ll \eta$ and cost $d$ admits a solution. We show moreover that the hypotheses of Corollary \ref{C:regudual} hold, and if $\text{\rm supp} \mu$ and $\text{\rm supp} \nu$ are at positive distance then the assumptions of Lemma \ref{L:normalcurr} are satisfied, i.e. the current $\dot g$ is normal. The main reference for this section is \cite{ohta:mcp}.
From now on $d=d_{L}$ and $\eta \in \mathcal{M}^{+}(X)$ is a locally finite measure on $X$.
Since $d_{L}=d$ there exists a Lipschitz function $\varphi$ potential for the transport problem:
hence in the following we will set
\[
\Gamma = \Gamma' = G = \Big\{ (x,y) \in X\times X : \varphi(x)-\varphi(y) = d(x,y) \Big\},
\]
where $\phi$ is a potential for the transport problem.
Let $H$ be the set of all geodesics: we regard $H$ as a subset of $\textrm{Lip}_1([0,1],X)$ with the uniform topology.
Define the evaluation map $e_{t}(\gamma)$ by
\begin{equation}\label{E:evalua}
\begin{array}{ccccc}
e &:& [0,1] \times H &\to& X \crcr
&& e_t(\gamma) &\mapsto& \gamma(t)
\end{array}
\end{equation}
It is immediate to see that $e_t(\gamma)$ is continuous.
A \emph{dynamical transference plan} $\Xi$ is a Borel probability measure on $H$, and
the path $\{ \xi_{t}\}_{t \in [0,1]} \subset \mathcal{P}^{2}(X)$ given by $\xi_{t}= (e_{t})_{\sharp}\Xi$
is called \emph{displacement interpolation} associated to $\Xi$.
We recall that $\mathcal{P}^{2}(X)$ is the set of Borel probability measures $\xi$
satisfying $\int_{X}d^{2}(x,y)\xi(dy)< \infty$ for some (and hence all) $x \in X$.
Define for $K\in \mathbb{R}$ the function $s_{K} : [0,+\infty) \to \mathbb{R}$ (on $[0,\pi/\sqrt{K})$ if $K>0$)
\begin{equation}\label{E:sk}
s_{K}(t):=
\begin{cases}
(1/\sqrt{K})\sin(\sqrt{K}t) & \rm{if}\ K>0, \crcr
t & \rm{if}\ K=0, \crcr
(1/\sqrt{-K})\sinh(\sqrt{-K}t) &\rm{if}\ K<0,
\end{cases}
\end{equation}
and let $N \in \mathbb{N}$.
\begin{definition}\label{D:mcp}
A metric measure space $(X,d,\eta)$ is said to satisfies the $(K,N)$-\emph{measure contraction property} ($MCP(K,N)$)
if for every point $x \in X$ and $\eta$-measurable set $A \subset X$ with $\eta(A)>0$
there exists a displacement interpolation $\{\xi_{t}\}_{t\in [0,1]}$ associated to
a dynamical transference plan $\Xi = \Xi_{x,A}$ satisfying the following:
\begin{enumerate}
\item We have $\xi_{0}=\delta_{x}$ and $\xi_{1}=\eta(A)^{-1} \eta_{\llcorner A}$;
\item \label{E:mcp} for $t \in [0,1]$
\[
\eta \geq (e_{t})_{\sharp} \bigg( t \bigg\{ \frac{s_{K}(t d(x,\gamma(1)) )}{s_{K}( d(x,\gamma(1)) )} \bigg\}^{N-1} \eta(A) \Xi \bigg),
\]
where we set $0/0=1$.
\end{enumerate}
\end{definition}
From now on we will assume the metric measure space $(X,d,\eta)$ to satisfies $MCP(K,N)$ for some $K\in \mathbb{R}$ and $N\in \mathbb{N}$.
Recall that $MCP(K,N)$ implies that $(X,d)$ is locally compact, Lemma 2.4 of \cite{ohta:mcp}.
The strategy to prove Assumption \ref{A:NDE} for any $d$-cyclically monotone plan is the following:
first we prove that for any $\pi \in \Pi(\mu,\delta_{x})$ $d$-monotone with $x$ arbitrary,
the marginal probabilities of $\eta$
obtained by the disintegration induced by the ray map $g$ are absolutely continuous w.r.t. $\mathcal{H}^{1}$ and their densities
satisfy some uniform estimates.
Then we observe that these estimates hold true also for any $\pi\in \Pi(\mu,\sum_{i\leq I} c_{i}\delta_{x_{i}})$ $d$-monotone.
Finally we show that the same estimates hold for general transference plans and therefore
we deduce that the densities of the marginals obtained by disintegrating $\eta$ w.r.t. any $d$-monotone plan $\pi$
are absolutely continuous w.r.t. $\mathcal{H}^{1}$.
By Lemma \ref{L:approx}, it is enough to assume that there exists $K_{1}, K_{2} \subset X$ compact set,
such that $\mu(K_{1}) =\nu(K_{2})=1$ and $d_{H}(K_{1},K_{2})< + \infty$.
Hence we can assume that $\textrm{diam\,}(X)< + \infty$ and $\eta(X)=1$.
\begin{lemma}
\label{L:punto}
Consider $\bar x\in X$ and let $\pi \in \Pi(\mu,\delta_{\bar x})$ be the unique $d$-cyclically monotone transference plan.
Then $\eta$ and the optimal flow induced by $\pi$ verify Assumption \ref{A:NDE}:
more precisely, $\eta = g_{\sharp} ( q m\otimes \mathcal{L}^{1} )$
and the density $q$ satisfies the estimate
\begin{equation}\label{E:marg}
q(y,t) \geq \bigg\{ \frac{s_{K}( d(g(y, t) , \bar x))}{s_{K}( d(g(y,s) , \bar x)) } \bigg\}^{N-1} q(y,s)
\end{equation}
for $m$-a.e. $y \in S$, for any $s\leq t$ such that $d(g(y,t),x)>0$.
\end{lemma}
We recall that $S$ is a section for the ray equivalence relation. Since $\mu \ll \eta$, \eqref{E:marg} implies that $\mu = g_{\sharp}( r m \otimes \mathcal{L}^{1})$ with $r \leq q$.
\begin{proof}
First observe that the potential for the transport problem is
\[
\varphi(x) : = \varphi(\bar x) + d(x,\bar x),
\]
so that the geodesics used by $\pi$ are exactly $H_{\bar x}:= H \cap e_{0}^{-1}(\bar x)$,
in the sense that
\[
G = \Big\{ \big(\gamma(1-s),\gamma(1-t)\big), s \leq t, \gamma \in H_{\bar x} \Big\}.
\]
{\it Step 1.}
We first prove that the set of initial points $A=a(X)$ has $\eta$-measure zero.
Suppose by contradiction that $\eta(A)>0$ and let $\Xi_{\bar x,A}$ be the dynamical transference plan associated:
we can assume that $\Xi_{\bar x,A}$ is supported on the set
$H_{\bar x,A}:= H_{\bar x} \cap e_{1}^{-1}(A)$.
Then the evolution of $A$ by the geodesics of $H_{\bar x, A}$ can be defined as
\[
A^{s}: = e_{1-s}(H_{\bar x,A}).
\]
By Condition \ref{E:mcp} of Definition \ref{D:mcp} and the fact that $e_{1-s}^{-1}(A^{s})=H_{\bar x, A}$
\begin{equation}\label{E:acaso}
\eta(A^{s}) \geq \eta(A) \int_{H_{\bar x,A}}
(1- s) \bigg\{ \frac{s_{K}((1- s) d(\bar x,\gamma(1)) ) }{s_{K}( d(\bar x,\gamma(1)) )} \bigg\} ^{N-1} \Xi_{x,A}(d\gamma) > 0,
\end{equation}
for all $s \in [0,1)$.
Since all $A^{s}$ are disjoint being the space non branching, it follows that $\eta(A)=0$.
{\it Step 2.} For $A$ with $\eta(A)>0$ let $\Xi_{\bar x,A}$ be the dynamical transference plan concentrated on a set
$H_{\bar x,A}: =H_{\bar x} \cap e_{1}^{-1}(A)$.
Denote as before $A^{s}: = e_{1-s}( H_{\bar x, A})$.
Observe that since the set initial point has $\eta$-measure zero, we can disintegrate $\eta$ w.r.t. the ray equivalence relation:
using the disintegration formula $\eta = \int \eta_{y} m(dy)$ the same estimate as in \eqref{E:acaso} yields
\[
\int \eta_{y}(A^{s}) m(dy) \geq \int \eta_{y}(A)m(dy) \bigg( \int_{ H_{\bar x,A}} (1- s) \bigg\{ \frac{s_{K}((1- s) d(\bar x,\gamma(1)) ) }{s_{K}( d(\bar x,\gamma(1)) )} \bigg\} ^{N-1} \Xi_{\bar x,A}(d\gamma) \bigg).
\]
By evaluating the above formula on sets of the form $A=g( S \times [t_{1},t_{2}] )$, where $g$ is the ray map
such that $g(y,0)=\bar x$ for all $y$, gives
\begin{align*}
\int_{S} \eta_{y} \big( g(y,[t_{1},t_{2}] (1-s)) \big) m(dy) \geq &~ \int_{S} \eta_{y} \big( g(y,[t_{1},t_{2}]) \big) m(dy) \crcr
&~ \qquad \cdot \bigg( \int_{ H_{\bar x,A}} (1- s) \bigg\{ \frac{s_{K}((1- s) d(\bar x,\gamma(1)) ) }{s_{K}( d(\bar x,\gamma(1)) )} \bigg\} ^{N-1} \Xi_{\bar x,A}(d\gamma) \bigg) \crcr
\geq &~ \int_{S} \eta_{y}\big(g(y,[t_{1},t_{2}])\big) m(dy) \min_{c\in [t_{1},t_{2}]} \bigg\{ (1-s) \frac{s_{K}((1- s) |c| ) }{s_{K}( |c| )} \bigg\} ^{N-1}.
\end{align*}
and therefore for $m$-a.e. $y$ and every $t_{1},t_{2}$
\begin{equation}\label{E:evo}
\eta_{y} \big( g(y,[t_{1},t_{2}] (1-s)) \big) \geq \eta_{y} \big( g(y,[t_{1},t_{2}]) \big) \min_{c\in [t_{1},t_{2}]} \bigg\{ (1-s) \frac{s_{K}((1- s) |c| ) }{s_{K}( |c| )} \bigg\} ^{N-1}.
\end{equation}
{\it Step 3.}
For $t_{1}<0$ consider the family of disjoint open sets
\[
t_{1} \bigg( 1 - \frac{k}{2n}, 1 - \frac{k+1}{2n}\bigg), \quad k = \{0,1,\dots, n-1 \}.
\]
The above estimate and the fact that $\eta_{y}$ is probability yield
\[
\eta_{y} \bigg\{ g\bigg(y, t_{1}\bigg(1,1-\frac{1}{2n}\bigg) \bigg) \bigg\} \leq \frac{1}{n}
\max_{c\in [t_{1},t_{1}/2]} \bigg\{ 2 \frac{s_{K}( |c| ) }{s_{K}( 2|c| )} \bigg\} ^{N-1}.
\]
Hence $\eta_{y} = q \mathcal{H}^{1}\llcorner_{g(y,\mathbb{R})}$ and $q$ satisfies \eqref{E:marg}.
\end{proof}
\begin{lemma}\label{L:piupunti}
Let $\pi\in \Pi(\mu,\sum_{i \leq I} c_{i}\delta_{x_{i}})$ $d$-cyclically monotone.
Then the conditional probabilities of the disintegration of $\eta$ w.r.t. the ray equivalence relation induced by $\pi$
are absolutely continuous w.r.t. $\mathcal{H}^{1}$ and the density $q(y,\cdot)$ satisfies
\[
q(y,t) \geq \bigg\{ \frac{s_{K}( d(g(y, t) , b(y)))}{s_{K}( d(g(y,s) , b(y))) } \bigg\}^{N-1} q(y,s).
\]
\end{lemma}
\begin{proof}
Let $\varphi$ be a potential for the transport problem with marginal $\mu$ and $\nu$.
Define
\[
E_{i}:= \bigg\{z \in \mathcal{T}_{e} : \varphi(z)-\varphi(x_{i})=d(z,x_{i})\bigg\}.
\]
Now each $E_{i}$ is sent by the optimal geodesic flow to $x_{i}$, so we can perform exactly the same calculations done in Lemma \ref{L:punto}. Indeed $E_{i}\cap E_{j} \subset a(X)$ which has $\eta$-measure zero,
$\eta_{\llcorner E_{i}}$ verifies \eqref{E:mcp} of Definition \ref{D:mcp} along the geodesic flow connecting $E_{i}$ to $x_{i}$.
\end{proof}
Given $\tilde H \subset \textrm{Lip}_1([0,1],X)$ a set of geodesics and $A\subset X$, define
\begin{equation}\label{E:evodue}
A^{s,\tilde H}:= e_{1-s}( e^{-1}_{1}(A) \cap \tilde H).
\end{equation}
\begin{lemma}\label{L:fine}
Assume that there exists two compact sets $K_{1}, K_{2}\subset X$ such that
\begin{enumerate}
\item $\mu(K_{1})=\nu(K_{2})=1$;
\item there exist $0<a \leq b < + \infty$ such that
\[
a = \min_{x_{1} \in K_{1}, x_{2}\in K_{2}} d(x_{1},x_{2}) \leq \max_{x_{1} \in K_{1}, x_{2}\in K_{2}} d(x_{1},x_{2}) ;
\]
\item $K_{2}$ is a section of $R$.
\end{enumerate}
Then if
\begin{equation}\label{E:geodue}
H(G) : = \Big\{ \gamma \in H : \exists y \in K_{2} \Big( \varphi(\gamma(0)) - \varphi(\gamma(1))= d(\gamma(0),\gamma(1))\, \wedge \,\gamma(0)=y \Big) \Big\},
\end{equation}
where $\varphi$ is the potential for the transport problem with marginal $\mu$ and $\nu$, then
\[
\eta(K_{1}^{s,H(G)}) \geq \eta(K_{1}) \min_{a \leq c \leq b} \bigg\{ (1- s) \frac{s_{K}((1- s) c ) }{s_{K}( c )} \bigg\} ^{N-1}.
\]
\end{lemma}
\begin{proof}
{\it Step 1.}
It follows directly from Lemma \ref{L:piupunti} that
the statement holds for $\nu = \sum_{i\leq I}c_{i} \delta_{y_{i}}$.
We thus consider the sequence of approximating problem constructed as follows:
let $\{y_{i}\}_{i \in \mathbb{N}}$ be a dense sequence in $K_{2}$ and for $I \in \mathbb{N}$
define
$$
\varphi_{I} (x) : = \min \Big\{ \varphi(y) + d(x,y), y \in \{y_{1},\cdots, y_{I}\} \Big\},
$$
$$
E_{i,I} := \Big\{ x \in X : \varphi_{I}(x) -\varphi_{I}(y_{i}) = d(x,y_{i}), i \leq I \Big\},
$$
$$
\nu_{I}= \sum_{i\leq I} c_{i,I} \delta_{y_{i}}, \quad \textrm{where} \quad c_{i,I}= \mu \bigg(E_{i,I}\setminus \bigcup_{j\neq i} E_{j,I} \bigg).
$$
Clearly $\varphi_{I}$ is a potential for the transport problem with marginal $\mu$ and $\nu_{I}$ and let
$$
H(G_{I}) : =
\Big\{ \gamma \in H : \varphi_{I}(\gamma(0)) - \varphi_{I}(\gamma(1))= d(\gamma(0),\gamma(1))\, \wedge \,\gamma(0) \in
\{y_{1},\cdots, y_{I}\} \Big\}.
$$
{\it Step 2.} Observe that $K_{1}^{s,H(G)}$ is compact.
In fact, since $K_{1}$ and $K_{2}$ are compact,
$H(G)\cap e_{1}^{-1}(K_{1})$ is compact and
since $e_{1-s}$ is continuous $K_{1}^{s,H(G)} = e_{1-s}(H(G))$ is compact.
For the same reasons the sets $K_{1}^{s,H(G_{I})}$ are compact.
{\it Step 3.} $K_{1}^{s,H(G_{I})}$ is contained in a compact set and $\varphi_{I} \to \varphi$ as $I \to + \infty$, so that
up to subsequences $K_{1}^{s,H(G_{I})}$ converges in Hausdorff distance to a compact subset of $K_{1}^{s,H(G)}$.
By the upper semicontinuity of Borel bounded measures with respect to Hausdorff convergence for compact sets
the claim follows.
\end{proof}
\begin{theorem}\label{T:regularity}
If $\pi\in \Pi(\mu,\nu)$ $d$-monotone then $\eta \llcorner_{\mathcal T_e} = g_{\sharp} (q m \otimes \mathcal{L}^{1})$, where $\mathcal T_e$ is the transport set with end points \eqref{E:TRe}, and for $m$-a.e. $y$ and $s\leq t$ it holds
\begin{equation}\label{E:bigreg}
\bigg\{ \frac{s_{K}( d(g(y, t) , b(y)))}{s_{K}( d(g(y,s) , b(y))) } \bigg\}^{N-1} \leq \frac{q(y,t)}{q(y,s)} \leq
\bigg\{ \frac{s_{K}( d(g(y, t) , a(y)))}{s_{K}( d(g(y,s) , a(y))) } \bigg\}^{N-1}
\end{equation}
\end{theorem}
\begin{proof}
{\it Step 1.}
We first show that the set of initial points has $\eta$-measure zero.
In fact suppose by contradiction that $\eta(a(S))>0$, where $S$ is a section for the ray equivalence relation of $\pi$.
Hence we can assume that $S$ and $a(S)$ are compact and at strictly positive distance.
Applying Lemma \ref{L:fine} to the transport problem with marginals $\eta\llcorner_{a(S)}$ and $f_{\sharp}\eta$,
where $f$ is the quotient map, it follows that $\eta(a(S))=0$.
{\it Step 2.}
Since the initial points have $\eta$-measure zero,
we can disintegrate $\eta \llcorner_{\mathcal T_e}$ w.r.t. the ray equivalence relation obtaining
$\eta \llcorner_{\mathcal T_e} = \int \eta_{y} m(dy)$.
By a standard covering argument, it is enough to prove the statement on the set
\[
D_{\varepsilon}:=\big\{x: d(x,b(x)) \geq \varepsilon\big\}.
\]
For any $0<\delta<\varepsilon$ we can take the section $S$ compact such that $d(f(x),b(x))=\delta$, in particular we have $g(y,\delta)=b(y)$.
For $S'\subset S$ and $t_{1}<t_{2}$ consider $\eta\llcorner_{g^{-1} (S'\times [t_{1},t_{2}])}$.
Applying Lemma \ref{L:fine} with
$$
\mu = \frac{\eta\llcorner_{g^{-1} (S'\times [t_{1},t_{2}])}}{ \eta(g^{-1} (S'\times [t_{1},t_{2}]))}, \quad
\nu = f_{\sharp}\mu
$$
where $f$ is the quotient map for the ray equivalence relation $R$, it holds
\begin{align*}
\int_{S'} \eta_{y} \big( g(y,[t_{1},t_{2}] (1-s)) \big) m(dy) \geq \min_{c\in [t_{1},t_{2}]} \bigg\{ (1-s) \frac{s_{K}((1- s) |c| ) }{s_{K}( |c| )} \bigg\} ^{N-1} \int_{S'} \eta_{y} \big( g(y,[t_{1},t_{2}]) \big) m(dy).
\end{align*}
As in Step 2 of the proof of Lemma \ref{L:punto}, the estimate \eqref{E:evo} holds for $m$-a.e. $y$ and every $t_{1}<t_{2}$
and we deduce
\[
\bigg\{ \frac{s_{K}( d(g(y, t) , b(y)) -\delta ) }{s_{K}( d(g(y,s) , b(y))-\delta) } \bigg\}^{N-1} \leq \frac{q(y,t)}{q(y,s)}.
\]
Letting $\delta \to 0$, we obtain the left hand side of \eqref{E:bigreg}.
{\it Step 3.}
The right hand side of of \eqref{E:bigreg} is obtained by the same procedure taking
$$
F_{\varepsilon}:= \big\{ x : d(d,a(x)) \geq \delta \big\}
$$
and the section $S$ such that $d(y,a(y))=\delta$ for all $y \in S$.
\end{proof}
\begin{figure}
\label{Fi:mongemetric5}
\psfrag{q}{$q(y,t)$}
\psfrag{sa}{$q(y,\bar t) \left( \frac{s_K(t+d(y,a(y))}{s_K(\bar t+d(y,a(y))} \right)^{N-1}$}
\psfrag{sb}{$q(y,\bar t) \left( \frac{s_K(d(y,b(y))-t}{s_K(d(y,b(y))-\bar t} \right)^{N-1}$}
\psfrag{qyt}{$q(y,\bar t)$}
\psfrag{da}{$-d(y,a(y))$}
\psfrag{db}{$d(y,b(y))$}
\psfrag{t}{$\bar t$}
\centerline{\resizebox{12cm}{8cm}{\includegraphics{mongemetric.5.eps}}}
\caption{The region where $q(y,t)$ takes values.}
\end{figure}
Since $\mu \ll \eta$, it follows that also the densities of the conditional probabilities of $\mu$ are absolutely continuous
w.r.t. $\mathcal{H}^{1}$, and therefore we have the following corollary.
\begin{corollary}\label{C:conclu}
Let $(X,d,\eta)$ satisfies $MCP(K,N)$, let $\mu,\nu \in \mathcal{P}(X)$ with $\mu \ll \eta$, then there exists a $\mu$-measurable map $T:X \to X$ such that
$T_{\sharp}\mu =\nu$ and
\[
\int d(x,T(x))\mu(dx) = \min_{\pi \in \Pi(\mu,\nu)} \int d(x,y) \pi(dxdy).
\]
\end{corollary}
We can obtain additional regularity of the conditional probabilities $\eta_y$ under $MCP(K,N)$: in particular we deduce that the conclusion of Corollary \ref{C:regudual} holds and if the support of $\mu$ and $\nu$ are compact sets with empty intersection the statements of Lemma \ref{L:normalcurr} and Remark \ref{R:abscurr} are true.
\begin{lemma}
\label{L:mhdregul}
The marginal densities
\[
\big( -d(a(y),y),d(y,b(y) \big) \ni t \mapsto q(y,t) \in \mathbb{R}^+
\]
are strictly positive Lipschitz continuous for $m$-a.e. $y \in S$, and for some constant $C>0$
\[
\text{\rm Tot.Var.} \big( q(y,\cdot) \big) \leq \frac{C}{d(a(y),b(y))}.
\]
\end{lemma}
\begin{proof}
From \eqref{E:bigreg} it follows immediately that the function $q(y,t) > 0$ and Lipschitz continuous for $t \in (-d(y,a(y)),d(y,b(y)) )$ and
$m$-a.e. $y$.
By differentiating it follows that
\begin{equation}\label{E:der}
-(N-1) \frac{s'_{K}( d(g(y, t) , b(y)))}{s_{K}( d(g(y,t) , b(y))) } \leq \frac{q'(y,t)}{ q(y,t)} \leq
(N-1) \frac{s'_{K}( d(g(y, t) , a(y)))}{s_{K}( d(g(y,t) , a(y))) }.
\end{equation}
In particular $q(y,\cdot)$ is Lipschitz.
For notational convenience let us assume that $d(a(y),y)=d(y,b(y))=l$.
From \eqref{E:bigreg} one can prove that
\[
q(y,t) \geq q(y,0) \cdot
\begin{cases}
\displaystyle \frac{s_{K}(l-t)}{s_{K}(l)}, & t\geq 0 \crcr
\displaystyle \frac{s_{K}(-l+t)}{s_{K}(-l)}, & t\leq 0
\end{cases}
\]
Since $\int q(y,t) dt=1$ it follows that
\[
q(y,0) \leq c_{K}(d(a(y),b(y))),
\]
where
\[
c_{K}(t) := \frac{s_{k}( t/2 )^{N-1}}{2} \bigg( \int_{0}^{t/2} s_{K}(\tau)^{N-1} d\tau \bigg)^{-1} \leq \frac{C}{t},
\]
being $C$ a constant depending only on $K$.
To show that
\[
\int_{-l}^{l} |q' (y,t)| dt < + \infty,
\]
it is enough to prove
\[
\int_{-l}^{0} |q' (y,t)| dt < + \infty,
\]
From \eqref{E:der} it follows
\[
\omega'(y,t) := q'(y,t) +(N-1) \frac{s'_{K}( l-t)}{s_{K}( l) }q(y,0) \geq 0
\]
so that
$$
\text{\rm Tot.Var.} \big( \omega(y,\cdot) ) \leq \bigg(1 + (N-1)\bigg( \frac{s_{K}(2l)}{s_{K}(l)} -1 \bigg) \bigg) q(y,0).
$$
Hence
\begin{align*}
\text{\rm Tot.Var.} \big( q(y,\cdot), (-\ell,0] \big) \leq&~ \text{\rm Tot.Var.} \big( \omega(y,\cdot), (-\ell,0] \big) + \text{\rm Tot.Var.} \bigg( (N-1) \frac{s'_{K}}{s_{K}}, (-\ell,0] \bigg) q(0,y) \crcr
\leq&~ \text{\rm Tot.Var.} \big( \omega(y,\cdot), (-\ell,0] \big) + (N-1) \frac{s'_{K}(2 l)}{s_{K}(l)} q(0,y) \crcr
\leq&~ \bigg( 1 + 2\bigg( \frac{s_{K}(2l)}{s_{K}(l)} - 1 \bigg) q(y,0).
\end{align*}
Collecting all the estimates, we get
\[
\text{\rm Tot.Var.} \big( q(y,\cdot) \big) \leq 2 \bigg(1 + 2\bigg( \frac{s_{K}(2l)}{s_{K}(l)} -1 \bigg) c_{K}(2l).
\]
\end{proof}
In general, the current $\dot g$ is not normal, as one can easily verify in $\mathbb T^2$ with the standard distance.
\section{Examples}
\label{S:examples}
We end this paper with some examples which shows how the different hypotheses of Section \ref{ss:Metric} enter into the analysis. In the following we denote the standard Euclidean scalar product in $\mathbb{R}^d$ as $\cdot$ and the standard distance in $\mathbb{T}^d$ by $|\cdot|$. We will also denote points by $p = (x,y,z,\dots) \in \mathbb{R}^d$, and $\alpha$ a fixed constant in $[0,1] \setminus \mathbb{Q}$.
\begin{example}[Non strongly consistent disintegration along rays]
\label{Ex:nonsection}
Consider the metric space
\[
(X,d) = \big( \mathbb{T}^2, |\cdot| \big)
\]
and the l.s.c. distance in the local chart $X = \{ (x,y) : 0 \leq x,y < 1 \}$
\[
d_L(p_1,p_2) :=
\begin{cases}
|x_1 - x_2 + i| & y_1 - y_2 = \alpha (x_1 - x_2) + i \alpha + n \crcr
+\infty & \text{otherwise}
\end{cases}
\]
for $i,n \in \mathbb{Z}$.
The sets $D_L$ are given by
\[
D_L(p_1) = \Big\{ (x,y) : y = y_1 + \alpha (x - x_1 + i) \mod 1, i \in \mathbb{N} \Big\},
\]
so that it is easy to see that the partition $\{D_L(p)\}_{p \in X}$ does not yield a strongly consistent disintegration. Since $t \mapsto (t \mod 1, \alpha t \mod 1)$ is a continuous not locally compact geodesic, Condition \eqref{Cond:XdL5} is not verified in this system.
Consider the measures $\mu = \mathcal{L}^2 \llcorner_{\mathbb{T}}$ and the map $T : (x,y)\mapsto (x,y + \alpha \mod 1)$: being $\mu$ invariant w.r.t. translations, one has $T_\sharp \mu = \mu$, and moreover
\[
\int d_L(x,T(x)) \mu(dx) = 1.
\]
If we consider points $(p_i,(x_i, y_i + \alpha \mod 1))$, $i = 1,\dots,I$, then the only case for which $d_L(p_{i+1},p_i) < +\infty$ is when $p_{i+1} = (x_i + t \mod 1, y_i + \alpha t \mod 1)$ for some $t \in \mathbb{R}$, i.e. they belong to the geodesic
\[
\mathbb{R} \ni t \mapsto (x_i + t \mod 1, y_i + \alpha t \mod 1) \in X.
\]
Hence, to prove $d_L$-cyclical monotonicity, it is sufficient to consider path which belongs to a single geodesic, where $d_L$ reduces to the the one dimensional length:
\[
d_L \big( (x,y), (x + t \mod 1, y + \alpha t \mod 1) \big) = |t|.
\]
Since translations in $\mathbb{R}$ are cyclically monotone w.r.t. the absolute value, we conclude that $T$ is $d_L$-cyclically monotone.
The fact that the optimal rays coincide with the sets $D_L$ yields that the disintegration is not strongly consistent, in particular there is not a Borel section up to a saturated negligible set. Note that every transference plan which leaves the common mass in the same place has cost $0$, so that this example shows the necessity of Condition \eqref{Cond:XdL5} for Proposition \ref{P:ortho}.
\begin{figure}
\label{Fi:mongem1}
\psfrag{x}{$x$}
\psfrag{y}{$y$}
\psfrag{DL(P)}{$D_L(p)$}
\psfrag{T2}{$\mathbb{T}^2$}
\psfrag{L2}{$\mathcal{L}^2 \llcorner_{\mathbb{T}^2}$}
\psfrag{p}{$p$}
\psfrag{T(p)}{$T(p)$}
\centerline{\resizebox{9cm}{9cm}{\includegraphics{mongemetric1.eps}}}
\caption{The metric space of Example \ref{Ex:nonsection}}
\end{figure}
\end{example}
\begin{example}[Non optimality of transport map]
\label{Ex:nooptir}
Consider countable copies of the manifolds $\mathbb{T}^2$: we denote them in local coordinates by
\[
C := \big\{ (x,y) : 0 \leq x,y < 1 \big\}, \quad C^i := \big\{ (x^i,y^i) : 0 \leq x^i,y^i < 1 \big\},\ i \in \mathbb{Z} \setminus \{0\}.
\]
With this fixed choice of coordinates, identify the points $(x,0) \equiv (x^i,0)$ if $0 \leq x = x^i < 1$. In other words, we glue the sets $C$, $C^i$, $i \in \mathbb{Z} \setminus \{0\}$, along a maximal circle $S$, which will be written in local coordinates by
\[
S = \big\{ \theta : 0 \leq \theta < 1 \big\}.
\]
The space $X$ is the set obtained with this procedure.
In the following points we need to divide $C$ into two parts: with the same coordinates as above, we set
\[
C^- := \big\{ (x,y) \in C : 0 \leq x < 1, 0 \leq y \leq 1/2 \big\}, \quad C^+ := \big\{ (x,y) \in C : 0 \leq x < 1, 1/2 < y < 1 \big\}.
\]
{\it Definition of $d$ and $d_L$.} The distance $d$ is defined as follows:
\[
d(p_1,p_2) = \min \Big\{ |p_1 - p_2|, |(x_1,y_1) - (\theta,0)| + |(x_2,y_2) - (\theta,0)|, \theta \in S^1 \Big\}.
\]
Note that $p_1 - p_2$ can be computed only when the points belong to the same component. It is fairly easy to see that $(X,d)$ is a compact set, in particular Polish.
The distance $d_L$ is defined as follows: if $\gamma : [0,1] \to X$ is a $d$-Lipschitz map, then set
\[
L(\gamma) := \int_0^1 \omega(\gamma(t),\dot \gamma(t)) dt, \quad \omega(p,v) :=
\begin{cases}
|\dot \gamma| & p \in C^-, \dot \gamma \cdot (-1,\alpha) = 0 \crcr
4 |\dot \gamma| & p \in C^+ \setminus S, \dot \gamma \cdot (-1,\alpha) = 0 \crcr
|\dot \gamma| & p \in C^i \setminus S, \dot \gamma \cdot (-1,i\alpha) = 0 \crcr
+\infty & \text{otherwise}
\end{cases}
\]
In other words, the Lipschitz path with finite length are a countable union of segments in $C$ or $C^i$, $i \in \mathbb{Z} \setminus \{0\}$ with slope $(\alpha,1)$, $(i\alpha,1)$, respectively. The distance $d_L$ is defined then by
\[
d_L(p_1,p_2) := \inf \Big\{ L(\gamma): \gamma \in \text{Lip}([0,1],X), \gamma(0)=p_1, \gamma(1)=p_2 \Big\}.
\]
{\it Study of the distance $d_L$.} To study the distance $d_L$, observe that it is enough to analyze the induced distance on $S^1$. We consider the length of the return map on $S$ depending on which sets we are moving on:
\begin{enumerate}
\item if we take the path $\theta \to \theta+\alpha$ along $C$, then its length is $\frac{5}{2} \sqrt{1+\alpha^2}$;
\item if we take the path $\theta \to \theta+i\alpha$ along $C^i$, then its length is $\sqrt{1+(i\alpha)^2}$.
\end{enumerate}
In particular, geodesics starting from $S$ and ending in some $C^i$, $i \in \mathbb{Z} \setminus \{0\}$, never take values in $C \setminus S$. Due to the invariance w.r.t. translations $(x,y) \mapsto (x+\alpha \mod 1,y)$, it is sufficient to study the structure the metric space $(D_L((0,0)),d_L)$.
The set $D_L((0,0))$ is the set $\{y = \alpha x + z \alpha \mod 1, z \in \mathbb{Z}\}$ in each component $C$, $C^i$, $i \in \mathbb{Z} \setminus \{0\}$. The metric $d_L \llcorner_{D_L((0,0))}$ is obtained as follows: given two points $(p_1,p_2)$, we can connect them using a path on the same component or by connecting each of them to points $\theta_1$, $\theta_2$ of $S$, and using one of the $C^i$ to connect these last points.
It follows that $(D_L((0,0)),d_L)$ is geodesic, and a more careful analysis shows that $d_L$ is actually l.s.c.. Moreover, the fact that $\sqrt{\cdot}$ is subadditive yields that there are not geodesic of infinite length: in particular all the assumptions listed on Page \pageref{P:assumpDL} are satisfied.
{\it Transport problem.} Define the sets
\[
A := \bigg\{ (x,y) \in C: 0 \leq x < 1, \frac{1}{2} < y < \frac{5}{8} \bigg\}, \quad B := \bigg\{ (x,y) \in C: 0 \leq x < 1, \frac{7}{8} < y < 1 \bigg\},
\]
and the measures $\mu := \mathcal{L}^2 \llcorner_A$, $\nu := \mathcal{L}^2 \llcorner_B$. Consider the two maps defined on $A$
\[
\begin{array}{ccccc}
T^+ &:& A &\to& B \crcr
&& (x,y) &\mapsto& T^+(x,y) := \Big( x + \frac{3}{8} \alpha \mod 1, y + \frac{3}{8} \Big)
\end{array}
\]
\[
\begin{array}{ccccc}
T^- &:& A &\to& B \crcr
&& (x,y) &\mapsto& T^-(x,y) := \Big( x - \frac{5}{8} \alpha \mod 1, y + \frac{3}{8} \Big)
\end{array}
\]
It is standard to show that $T^\pm_\sharp \mu = \nu$.
Let $(p_i,T^+(p_i)) \in \text{graph}(T^+)$, $i = 1,\dots,I$: from the definition of $d_L$, $d_L(p_{i+1},T^+(p_i))$ can be either equal to $d_L(p_i,T^+(p_i))$ or greater than $3 \sqrt{1+\alpha^2}$ (by taking the path along $C^- \cup C^1$). Since $d_L(p_i,T^+(p_i)) = \frac{3}{2} \sqrt{1+\alpha^2}$, it follows that $T^+$ is $d_L$-cyclically monotone.
However, one has
\[
\int d_L(x,T^-(x)) \mu(dx) = \frac{1}{8}\sqrt{1 + \alpha^{2}} < \frac{3}{2} \frac{1}{8}\sqrt{1 + \alpha^{2}} = \int d_L(x,T^+(x)) \mu(dx).
\]
Hence the $d_L$-cyclical monotonicity is not sufficient for optimality. Note that Assumption \ref{A:NDE} is verified.
\begin{figure}
\label{Fi:mongemetri2}
\psfrag{D}{$C$}
\psfrag{Ci}{$C^i$}
\psfrag{t}{$\theta$}
\psfrag{t+a}{$\theta+\alpha$}
\psfrag{S}{$S$}
\psfrag{t+ia}{$\theta+i\alpha$}
\psfrag{D-}{$C^-$}
\psfrag{D+}{$C^+$}
\psfrag{T-}{$T^-$}
\psfrag{T+}{$T^+$}
\psfrag{mu}{$\mu$}
\psfrag{nu}{$\nu$}
\centerline{\resizebox{14cm}{5cm}{\includegraphics{mongemetric2.eps}}}
\caption{The metric space of Example \ref{Ex:nooptir}.}
\end{figure}
\end{example}
|
\subsection*{Figures}
\begin{figure}[Hhbt]
\centering
\includegraphics[width=12cm]{Fig1_KittelEtAl08.eps}
\caption{
(a) Schematic drawing of the tip in its holder, with blown-up
cross section of the tip's very end, displaying the glass capillary,
the platinum wire, and the gold coating which form the thermocouple.
(b) SEM image of a typical NSThM tip. The platinum wire protrudes from
the center of the glass mantle. Both are covered by a gold film.
}
\end{figure}
\vspace{2cm}
\begin{figure}[Hhbt]
\centering
\includegraphics[width=12cm]{Fig2_KittelEtAl08_small.eps}
\caption{
(a) STM-topography of a gold surface on mica. (b) Spatial
distribution of the local thermovoltage measured by the thermocouple.
The temperature of the sample is 110~K, that of the probe 293~K.
}
\end{figure}
\vspace{2cm}
\begin{figure}[Hhbt]
\centering
\includegraphics[width=12cm]{Fig3_KittelEtAl08_small.eps}
\caption{
(a) Numerically calculated LDOS in $10^6\,{\rm m}^{-3}\,{\rm s}$ at a constant
distance of $9\,{\rm nm}$ above the two-dimensional topography directly extracted from
the STM data in Fig.~2. (b) A plot of the thermovoltage data (in arbitrary units) from Fig.~2 rescaled as described in the text.
}
\end{figure}
\vspace{2cm}
\begin{figure}[Hhbt]
\centering
\includegraphics[width=12cm]{Fig4_KittelEtAl08.eps}
\caption{
(a) LDOS at a constant distance of 9~nm above the surface profile in comparision to the
thermovoltage data, for the linescan indicated by the vertical arrow
in Fig.~3. (b) LDOS at a constant distance of 9~nm above the profile in comparision to the
thermovoltage data, for the line scan indicated by the horizontal arrow in Fig.~3.
\
}
\end{figure}
\end{document}
|
\section{Introduction and Set-Up}
The evolution of the diameter of a bounded set under the action of a stochastic flow has been studied since the 1990's
(see~\cite{CSS99},~\cite{CSS00},~\cite{DKK04},~\cite{LS01},~\cite{LS03}, \cite{SS02}, and the survey article~\cite{SCH09} to name just
a few references). For a large class of flows -- including isotropic Brownian flows (IBFs) with non-negative top Lyapunov exponent --
the diameter is known to be linearly growing in time.
In this paper, we will consider the evolution of the spatial derivative of a flow in time and derive an explicit upper
bound on the supremum of the spatial derivative taken over a bounded set. Our bound depends on the box dimension of the set.
In the case of IBFs such a result has been obtained in~\cite{vB09} with a different (and more technical) proof. We will be much more
general with our set-up but, contrary to \cite{vB09}, will not derive lower bounds for the growth rates.
We point out that there is a close link between this paper and work of Peter Imkeller: in~\cite{ImSch99}, the growth of the spatial
derivative of a flow in the spatial direction was studied over a fixed time horizon $[0,T]$ and we did not care about constants
(even $T$ was regarded as a constant).
Still, the proof of Lemma \ref{le:twopointderivative}, which constitutes the core of our results,
largely follows that of Proposition 2.3 of \cite{ImSch99}. Apart from keeping track of constants, our proof here differs from that
in~\cite{ImSch99} towards the end
when we apply a non-linear Gronwall-type Lemma (the usual Gronwall Lemma will not provide an exponential growth rate in $T$).
Exponential bounds on the growth of spatial derivatives play an important role in the proof of Pesin's formula for stochastic
flows (see~\cite{LQ95}). They can also be used to obtain bounds on the exponential growth rate of e.g.~the length of a curve under a flow.
Even though we try to keep track of constants, we make no claims about optimality (and we conjecture that our bound is far from optimal).
We give explicit formulas for the exponential growth rate only for the first order derivative but indicate how such bounds
can be obtained also for higher order derivatives under additional smoothness assumptions.
The paper is organized as follows: we start by defining a suitable class of stochastic flows. Then, we provide a general result --
Theorem \ref{th:zweites} -- which shows how one can obtain exponential growth rates for a random field $\psi$ indexed by ${\mathbf R}^d$ given moment
bounds on the field and on two-point differences of the field. Afterwards, we apply this theorem to the derivative of a stochastic flow.
Here, the main task is to compute the moment bounds needed in order to apply Theorem \ref{th:zweites}. Then, we specialize to IBFs.\\
Let us introduce our set-up which is essentially the same as in \cite{ImSch99} and is based on \cite{Ku90}.
Let $F(x,t), t\geq0$ be a family of $\ensuremath{\mathbf{R}^{d}}$-valued continuous semimartingales on a filtered probability space
$(\Omega,{\cal F},({\cal F}_t)_{t\geq 0},\P)$ indexed by $x\in\ensuremath{\mathbf{R}^{d}}$, starting at 0.
Let $F(x,t)=M(x,t)+V(x,t)$ be the canonical decomposition into a local martingale $M$ and a process $V$ of locally bounded variation
(both starting at 0). We will assume throughout that both $M$ and $V$ are jointly continuous in $(x,t)$. Furthermore we assume
that there exist $a:\ensuremath{\mathbf{R}^{d}}\times\ensuremath{\mathbf{R}^{d}}\times[0,\infty[\times\Omega\to{\mathbf R}^{d\times d}$ which is continuous in the first two and predictable in the
last two variables and $b:\ensuremath{\mathbf{R}^{d}}\times[0,\infty[\times\Omega\to\ensuremath{\mathbf{R}^{d}}$ which is continuous in the first and predictable in the
last two variables such that
$$\langle M_i(x,.),M_j(y,.)\rangle(t)=\int_0^ta_{ij}(x,y,u)\dd u,\hspace{5mm}V_i(x,t)=\int_0^tb_i(x,u)\dd u. $$
Here, $\langle.,.\rangle$ denotes the joint quadratic variation.
The pair of random fields $(a,b)$ is called the {\em local characteristics} of the semimartingale field $F$.
We will abbreviate $\mathcal{A}(x,y,t):=a(x,x,t)-a(x,y,t)-a(y,x,t)+a(y,y,t)$
(which is the derivative of the quadratic variation of $M(x,t)-M(y,t)$). Throughout, we will assume that the following hypothesis
holds:\\
\newpage
\noindent {\bf Hypothesis (A):}
\begin{align*}
\esssup_{\omega\in\Omega}&\sup_{t \ge 0} \sup_{x,y\in \ensuremath{\mathbf{R}^{d}}}\Big(\frac{\|a(x,y,t)\|}{(1+|x|)(1+|y|)}
+ \sum_{k=1}^d \|D_{x_k} D_{y_k} a(x,y,t)\|\\ &+ \sum_{k=1}^d \| D_{x_k} D_{y_k}a(.,.,t)\|^\sim \Big)<\infty, \mbox{ and }\\
\esssup_{\omega\in\Omega}&\sup_{t \ge 0} \Big( \sup_{x\in \ensuremath{\mathbf{R}^{d}}} \frac{|b(x,t)|}{1+|x|}+\sup_{x\in\ensuremath{\mathbf{R}^{d}}}\|D_x b(x,t)\|+
\sup_{x\neq y \in\ensuremath{\mathbf{R}^{d}}}\frac{ \|D_xb(x,t)-D_y b(y,t)\|}{|x-y|}\Big)<\infty,
\end{align*}
where
$$
\| f(.,.)\|^\sim:=\sup_{x\neq x',y\neq y'}\left\{\frac{\|f(x,y)-f(x',y)-f(x,y')+f(x',y')\|}{|x-x'|\,|y-y'|} \right\}.
$$
Since Hypothesis (A) implies the assumptions of \cite[Theorem 4.6.5]{Ku90} (with $k=1$, $\delta=1$), the stochastic differential equation
\begin{equation}\label{sde} \dd X(t)=F(X(t),\dd t ),\qquad X(s)=x,\;t\ge s \end{equation}
not only admits a unique solution for each fixed $x \in {\mathbf R}^d$ and $s \ge 0$, but even generates a
{\em stochastic flow of diffeomorphisms} $\phi$, i.e.~there exist a random field
$\phi:[0,\infty)^2 \times {\mathbf R}^d \times \Omega \to {\mathbf R}^d$ and a set $\Omega_0$ of full measure such that
\begin{itemize}
\item $t \mapsto \phi_{s,t}(x)$, $t\ge s$ solves \eqref{sde} for all $x \in {\mathbf R}^d$, $s \ge 0$.
\item $\phi_{s,t}(\omega)$ is a diffeomorphism on $\ensuremath{\mathbf{R}^{d}}$ for all $s,t \ge 0$, $\omega \in \Omega_0$.
\item $\phi_{s,u}=\phi_{t,u} \circ \phi_{s,t}$ for all $s,t,u \ge 0$, $\omega \in \Omega_0$.
\item $(s,t,u) \mapsto \phi_{s,t}(x)$ is continuous for all $\omega \in \Omega_0$.
\end{itemize}
We will often write $x_t:=\phi_t(x):=\phi_{0,t}(x,\omega) $.
\section{Exponential Growth Rates: General Results}
In the following lemma and theorem, $o(T)$ stands for a function $g(T)$ which may depend on $q$, but not on $x,y$
such that $\lim_{T \to \infty} g(T)/T=0$.
\begin{lemma}\label{erstes} Let $(E,\rho)$ be a complete, separable metric space and let
$(t,x) \mapsto \psi_t(x)$ be a continuous
random field on $[0,\infty) \times {\mathbf R}^d$ with values in $(E,\rho)$ which satisfies
$$
{\mathbf E} \sup_{0 \le t \le T} \rho(\psi_t(x),\psi_t(y))^q \le |x-y|^q \exp\{ (cq^2+\hat c q)T + o(T)\}
$$
for some $q > d$ and all $x,y \in {\mathbf R}^d$.
Then, for $u>0$, we have
\begin{align*}
\P \Big\{ &\sup_{x,y \in [0,1]^d} \sup_{0 \le t \le T} \rho(\psi_t(x),\psi_t(y)) \ge u \Big\}
\le \exp\{(cq^2+\hat c q)T + o(T)\} u^{-q},
\end{align*}
\end{lemma}
\begin{proof} This follows from Kolmogorov's continuity theorem (e.g.~in the version of~\cite[Lemma 2.1]{SCH09}).
\end{proof}
\begin{theorem}\label{th:zweites} Let $(E,\|.\|)$ be a separable real Banach space and let $(t,x) \mapsto \psi_t(x)$
be a continuous random field on $[0,\infty) \times {\mathbf R}^d$ with values in $(E,\|.\|)$ which satisfies
\begin{equation}\label{oans}
{\mathbf E} \sup_{0 \le t \le T} \|\psi_t(x)-\psi_t(y)\|^q \le |x-y|^q \exp\{ (cq^2+\hat c q)T + o(T)\}
\end{equation}
for some $c>0$, $\hat c \in {\mathbf R}$, all $q > d$ and all $x,y \in {\mathbf R}^d$.
Assume further that
\begin{equation}\label{zwo}
\sup_{x}{\mathbf E} \sup_{0 \le t \le T} \|\psi_t(x)\|^q \le \exp\{ (kq^2+\hat k q)T + o(T)\}
\end{equation}
for some $k>0$, $\hat k \in {\mathbf R}$ and for all $q \ge 0$. If ${\mathbb X}$ is any compact subset of ${\mathbf R}^d$ with box
dimension $\Delta$, then
\begin{equation}\label{toshow}
\limsup_{T \to \infty} \frac 1T \log \sup_{x \in {\mathbb X}} \sup_{0 \le t \le T} \|\psi_t(x)\| \le \xi \;\mbox{ a.s.},
\end{equation}
where $$\xi=
\left\{\begin{array}{ll}
\hat k & \mbox{ if }\; \hat k \ge cd+\hat c\\
\hat k + 2\sqrt{k\Delta \gamma_1}&\mbox{ if }\; \hat k \le cd+\hat c \;\mbox{ and }\;
2\sqrt{ck}\Delta d+cd^2-2c\Delta d +\Delta(\hat k-\hat c)\ge 0\\
\hat k + 2\sqrt{k\Delta \gamma_2} &\mbox{ if }\; \hat k \le cd+\hat c \;\mbox{ and } \;
2\sqrt{ck}\Delta d+cd^2-2c\Delta d +\Delta(\hat k-\hat c)\le 0,
\end{array}\right.
$$
where
\begin{align*}
\gamma_1:&=
\left\{\begin{array}{ll}\frac{cd + \hat c - \hat k}{2\sqrt{kd}}& \mbox{ if }\;\Delta = d\\
\Big( \Big(1-\frac {\Delta}{d}\Big)^{-1} \Big(-\sqrt{k\Delta}
+\sqrt{k\Delta+\big(1-\frac{\Delta}d\big)(cd+\hat c-\hat k)}\Big) \Big)^2& \mbox{ if }\;\Delta < d
\end{array}\right.\\
\gamma_2:&=\Big(\sqrt{c\Delta}+\sqrt{k\Delta}+\sqrt{(\sqrt{c}-\sqrt{k})^2 \Delta - \hat k + \hat c}\Big)^2.
\end{align*}
\end{theorem}
\begin{proof}
Let ${\varepsilon} >0$. For each $\gamma >0$, we can cover the set ${\mathbb X}$ with $N\le \ee^{\gamma T(\Delta + {\varepsilon})}$ balls of diameter
$\ee^{-\gamma T}$ in case $T$ is large enough. For given such $\gamma,T$ we denote these balls by ${\mathbb X}_1,...{\mathbb X}_N$ and
their centers by $x_1,...,x_N$. Let $r >0$.
Then, using Lemma~\ref{erstes}, we obtain
$$
\P\big\{\exists i \in \{1,...,N\} \mbox{ s.t. } \sup_{0 \le t \le T} \diam (\psi_t({\mathbb X}_i)) \ge \ee^{rT}\big\}
\le \ee^{\gamma T(\Delta + {\varepsilon})}\ee^{(cq^2+(\hat c - \gamma -r)q)T + o(T)},
$$
and therefore
\begin{align*}
B(r):=&\limsup_{T \to \infty} \frac 1T \log \P\big\{\exists i \in \{1,...,N\}
\mbox{ s.t. } \sup_{0 \le t \le T} \diam (\psi_t({\mathbb X}_i)) \ge \ee^{rT}\big\}\\ \le&\gamma \Delta + cq^2+(\hat c-\gamma -r)q.
\end{align*}
Optimizing over $q > d$ yields
$$
B(r) \le \left\{
\begin{array}{cc}
\gamma\Delta-\frac{(r-\hat c+\gamma)^2}{4c}&\mbox {if } r\ge 2cd+\hat c-\gamma\\
\gamma\Delta+(cd+\hat c-\gamma-r)d&\mbox {if } r< 2cd+\hat c-\gamma\end{array}
\right..
$$
Further,
$$
\P\big\{\exists i \in \{1,...,N\} \mbox{ s.t. } \sup_{0 \le t \le T}\|\psi_t(x_i)\|\ge \ee^{rT}\big\}
\le \ee^{\gamma T(\Delta + {\varepsilon})}\ee^{(kq^2+(\hat k -r)q + o(1))T}
$$
and therefore
$$
C(r):=\limsup_{T \to \infty} \frac 1T \log \P\big\{\exists i \in \{1,...,N\} \mbox{ s.t. }
\sup_{0 \le t \le T}\|\psi_t(x_i)\|\ge \ee^{rT}\big\} \le \gamma \Delta + kq^2 + (\hat k - r)q.
$$
Optimizing over $q \ge 0$, we get
$$
C(r) \le \gamma \Delta - \frac{(r-\hat k)^2}{4k}
$$
provided that $r \ge \hat k$ which we will assume to hold from now on.
Once we know that for a particular value of $r>0$
$$
A(r):=\limsup_{T \to \infty} \frac 1T \log \P\{ \sup_{0\le t\le T}\sup_{x \in {\mathbb X}} \|\psi_t(x)\|\ge 2e^{rT}\} <0,
$$
then a simple Borel-Cantelli argument (using the fact that
$T \mapsto \sup_{0\le t\le T}\sup_{x \in {\mathbb X}} \|\psi_t(x)\|$
is non-decreasing) shows that \eqref{toshow} holds with $\xi$ replaced by $r$. Since
$A(r) \le B(r) \vee C(r)$, we have $A(r)<0$ whenever there exists some $\gamma>0$ such that both upper bounds of
$B(r)$ and $C(r)$ are negative.
Defining $\xi$ as the infimum over all such $r$, we obtain
\begin{equation}\label{ximax}
\xi=\inf_{\gamma>0}\Big(
(\hat k +2\sqrt{k\gamma\Delta}) \vee \left\{\begin{array}{cc}
2\sqrt{c\gamma\Delta}+\hat c-\gamma&:\gamma \Delta \ge c d^2\\
\gamma\Delta d^{-1}+cd+\hat c-\gamma &:\gamma \Delta \le c d^2
\end{array}\right.\Big).
\end{equation}
Computing the infimum, we obtain the result in the theorem.
\end{proof}
\begin{remark} It follows from \eqref{ximax} that
$$
\hat k \le \xi \le (cd + \hat c)\vee \hat k.
$$
Note that the lower bound is attained in case $\Delta=0$.
\end{remark}
\begin{remark}
Our assumptions on the range of admissible values of $q$ in Theorem \ref{th:zweites} are a bit arbitrary (but motivated by
applications to IBFs). It is clear from the proof that \eqref{toshow} still holds with a larger value of $\xi$ if
we only assume that \eqref{oans} holds for {\em some} $q>d$ and that \eqref{zwo} holds for {\em some} $q>0$ (the values of
$q$ can be different).
\end{remark}
\section{Application to the Derivative of a Stochastic Flow}
Next, we want to use the results in the previous section to obtain bounds on the exponential growth rate of
the supremum of the derivative of a stochastic flow taken over all initial points in a compact set
of box dimension $\Delta$. In order to apply Theorem \ref{th:zweites}, we have to estimate moments of the difference
of derivatives of a stochastic flow. We start by introducing some more notation (as in~\cite{ImSch99}).\\
Let $\phi=(\phi^1,\ldots,\phi^d)$ be a stochastic flow of diffeomorphisms generated by~\eqref{sde}
satisfying Hypothesis (A) and let $p\geq1$. Fix $i\in\{1,\ldots,d\}$ and define
$Y_j(t):=D_i\phi^j_t(x)$, $Z(t):=\sup_{0\leq s\leq t}\big(\sum_{j=1}^d Y_j^2(s)\big)^{1/2}$,
$f_p(t)=\max_i\big({\mathbf E} Z(t)^p \big)^{1/p} $, and $\tilde f_p(t)=\max_i\big({\mathbf E}\big(\sum_{j=1}^d Y_j^2(t) \big)^{p/2}\big)^{1/p}$.
Here and in the following, we write $D_k$ instead of $D_{x_k}$ or $D_{y_k}$.
We will also need two two-point versions
$V_j(t):=D_i \phi^j_{0,t}(x) - D_i \phi^j_{0,t}(y)$, $W(t):=\sup_{0\le s\le t} \big(\sum_{j=1}^d V_j^2(s) \big)^{1/2} $
and $g_p(t):=\max_i\big({\mathbf E}\big(W^p(t)\big)^{1/p}$. Let $C_p$ denote the constant in Burkholder's inequality. It is well-known, that
there exists a constant $k_5>0$ such that $C_p \le (k_5 p^{1/2})^p$ for all $p \ge 2$ (see \cite{BY82}) (one can choose $k_5=2\sqrt{5}$).
We point out that $D_k D_k$ in front of a function of two spatial arguments means that we differentiate with respect to the $k-$th
component of {\em both} arguments. Note that we trivially have $\tilde f_p(t) \le f_p(t)$, so the terms $\tilde f_p(t)$
which appear in the upper bound of $H$ in the following lemma can be replaced by $f_p(t)$ (but one may get better bounds by not doing this).
\begin{lemma}\label{le:twopointderivative}
Let $\phi$ be the flow generated by $F$ satisfying Hypothesis (A) and denote
\begin{enumerate}
\item $k_1:=\mathrm{esssup}_{\omega \in \Omega}\sup_{\bar x\in\ensuremath{\mathbf{R}^{d}},t\geq 0,1\le j,k\le d}|D_kD_ka_{jj}(\bar x,\bar x,t)|$,
\item $k_2:=\mathrm{esssup}_{\omega \in \Omega} \sup_{\bar x,\bar y\in\mathbf{R}^d,t\geq0,1\le j,k \le d}
\frac{|D_k D_k \mathcal{A}_{jj}(\bar x,\bar y,t)|}{|\bar x-\bar y|^2}$,
\item $k_3:=\mathrm{esssup}_{\omega \in \Omega}\sup_{\bar x\in\ensuremath{\mathbf{R}^{d}},t\geq 0,1\le j,k\le d}|D_kb_j(\bar x,t)|$,
\item $k_4:=\mathrm{esssup}_{\omega \in \Omega}\sup_{\bar x,\bar y\in\mathbf{R}^d,t\geq0,1\le j,k\le d}
\frac{ |D_k b_j(\bar x,t)- D_k b_j(\bar y,t)|}{|\bar x-\bar y|}$.
\end{enumerate}
Then there exist constants $\Lambda$, $\bar{c}$ and $\sigma \ge 0$ such that for all $p\geq2$, $t \ge0$ and $\bar x,\bar y \in {\mathbf R}^d$,
\begin{equation}\label{esti}
\Big({\mathbf E} |\phi_t(\bar x)-\phi_t(\bar y)|^p\Big)^{1/p}\le \bar{c}|\bar x-\bar y|\ee^{(\Lambda +p\sigma^2/2)t}.
\end{equation}
Further, for all $p \ge 2$, $x \in {\mathbf R}^d$, and all $\alpha_1,\,\alpha_2,\,\alpha_3>1$ whose reciprocals sum up to 1,
$f_p$ satisfies
\begin{equation}\label{eq:fgleichung}
f_p^2(t)\leq \alpha_1 + \alpha_1 \frac{\alpha_2\bar{d}^2k_1C_p^{2/p}+\sqrt{\alpha_3}\bar{d}k_3}{\alpha_2\bar{d}^2k_1C_p^{2/p}+2\sqrt{\alpha_3}\bar{d}k_3}
\Big(\exp\{(\alpha_2\bar{d}^2k_1C_p^{2/p}+2\sqrt{\alpha_3}\bar{d}k_3)t\}-1\Big),
\end{equation}
where $\bar{d} = d^{2-1/p}$.
Further, for all $p \ge 2$, $x,y \in {\mathbf R}^d$, and all $\beta_m>1$, $m=1,2,3,4$ whose reciprocals sum up to 1, we have
\begin{equation}\label{eq:ggleichung}
g_p^2(t)\leq H(t) + H(t)\frac{\bar C_1+\sqrt{\bar C_2}}{\bar C_1+2\sqrt{\bar C_2}} \Big(\exp\{(\bar C_1+2\sqrt{\bar C_2})t\}-1\Big),
\end{equation}
where $\bar C_1:=\beta_1d^{3-2/p} C_p^{2/p} k_1$, $\bar C_2:=\beta_3d^{3-2/p}k_3^2 $ and
$$
H(t):=d^3\bar{c}^2|x-y|^2\Big(C_p^{2/p}\beta_2 k_2 \int_0^t \tilde f^2_{2p}(s)\ee^{2(\Lambda+p\sigma^2)s} \dd s
+ \beta_4 k_4^2\Big(\int_0^t\tilde f_{2p}(s)\ee^{(\Lambda+\sigma^2p)s}\dd s \Big)^2\Big).
$$
\end{lemma}
\begin{proof}
First note that $k_1,...,k_4<\infty$ since $F$ satisfies Hypothesis (A). Assertion \eqref{esti} follows from \cite[Lemma 2.6]{SCH09}
and the fact that $F$ satisfies (A) (one can choose $\sigma=\tilde a$ and $\Lambda=\tilde b+(d-1)\tilde a^2/2$,
where $\tilde b$ is a deterministic upper bound of the Lipschitz constant of $b$ and $\tilde a\ge 0$ is chosen such that
$\|\mathcal{A}(x,y,t,\omega)\|\le \tilde a^2|x-y|^2$ for all $x,y \in {\mathbf R}^d$ and almost all $\omega \in \Omega$). Fix $i \in \{1,...,d\}$.
We have
$$
Y_j(s)= \delta_{i,j}+\sum_{n=1}^d \int_0^s Y_n(u)D_n F_j(x_u,\dd u)
$$
(see \cite[p.174 (21)]{Ku90} or~\cite[(18)]{ImSch99}).
Applying Burkholder's inequality, we get
\begin{align}
\Big({\mathbf E}&\sup_{0\leq s\leq t}|Y_j(s)|^p \Big)^{1/p}
\le \delta_{i,j}+\sum_{n=1}^d \Big({\mathbf E}\big(\sup_{0\leq s\leq t}\int_0^s Y_n(u)D_n F_j(x_u,\dd u)\big)^p \Big)^{1/p}\nonumber\\
&\le\delta_{i,j}+ C_p^{1/p}\sum_{n=1}^d\Big({\mathbf E}\big(\int_0^t Y^2_n(u)D_nD_na_{jj}(x_u,x_u,u)\dd u\big)^{p/2}\big)^{1/p}\nonumber\\
& \hspace{.5cm} + \sum_{n=1}^d\Big({\mathbf E}\big(\int_0^t |Y_n(u) D_nb_j(x_u,u)|\dd u\big)^p\Big)^{1/p}\nonumber\\
&\le\delta_{i,j}+C_p^{1/p}\sqrt{k_1} \sum_{n=1}^d \Big( {\mathbf E} \big(\int_0^t Y_n^2(u)\dd u\big)^{p/2}\Big)^{1/p}\nonumber\\
& \hspace{.5cm} + k_3 \sum_{n=1}^d\Big({\mathbf E}\big(\int_0^t |Y_n(u)|\dd u\big)^p\Big)^{1/p}.\label{zweites}
\end{align}
Since $p \ge 2$, Jensen's inequality implies
\begin{align*}
\sum_{n=1}^d &\Big( {\mathbf E} \big(\int_0^t Y_n^2(u)\dd u\big)^{p/2}\Big)^{1/p}
\le \sqrt{d} \Big(\sum_{n=1}^d \Big( {\mathbf E} \big(\int_0^t Y_n^2(u)\dd u\big)^{p/2}\Big)^{2/p}\Big)^{1/2} \\
&\le \sqrt{d} \Big(\int_0^t \sum_{n=1}^d\big({\mathbf E} |Y_n(u)|^p\big)^{2/p}\dd u\Big)^{1/2}
\le \sqrt{d} \Big(\int_0^t d^{1-\frac 2p}\big({\mathbf E} \sum_{n=1}^d |Y_n(u)|^p\big)^{2/p}\dd u\Big)^{1/2}
\\
&\le d^{1-1/p}\Big(\int_0^t \Big({\mathbf E} \big(\sum_{n=1}^d Y_n^2(u)\big)^{p/2}\Big)^{2/p}\dd u\Big)^{1/2} \le d^{1-1/p} \Big(\int_0^tf_p^2(u)\dd u\Big)^{1/2}.
\end{align*}
The term \eqref{zweites} can be estimated similarly:
\begin{align*}
\sum_{n=1}^d&\Big({\mathbf E}\big(\int_0^t |Y_n(u)|\dd u\big)^p\Big)^{1/p} \le \int_0^t \sum_{n=1}^d \big( {\mathbf E} |Y_n(u)|^p\big)^{1/p} \dd u\\
&\le d^{1-1/p} \int_0^t \Big({\mathbf E} \sum_{n=1}^d |Y_n(u)|^p \Big)^{1/p} \dd u \le d^{1-1/p}
\int_0^t \Big( {\mathbf E} \Big(\sum_{n=1}^d Y_n^2(u)\Big)^{p/2}\Big)^{1/p}\dd u\\
&\le d^{1-1/p} \Big(\int_0^tf_p(u)\dd u\Big).
\end{align*}
Therefore we obtain
\begin{align*}
f_p(t)&=\max_i ({\mathbf E} Z^p(t))^{1/p}\leq \max_i\sum_{j=1}^d \Big({\mathbf E}\sup_{0\leq s\leq t}|Y_j(s)|^p \Big)^{1/p} \\
&\le 1+d^{2-1/p}\sqrt{k_1}C_p^{1/p}\Big(\int_0^tf_p^2(s)\dd s\Big)^{1/2}+d^{2-1/p}k_3\int_0^tf_p(s) \dd s.
\end{align*}
Taking squares and using the formula $(A+B+C)^2 \le \alpha_1 A^2 + \alpha_2 B^2 + \alpha_3 C^2$ for $A,B,C \ge 0$, we obtain
\begin{align*}
f^2_p(t)\leq \alpha_1+\alpha_2 d^{4-2/p}k_1C_p^{2/p}\int_0^tf_p^2(s)\dd s+\alpha_3 d^{4-2/p}k_3^2\big(\int_0^t f_p(s) \dd s\big)^2 ,
\end{align*}
and hence~\eqref{eq:fgleichung} by Lemma~\ref{le:gronwall}.
Let us now treat the two-point differences and recall from~\cite[p. 123]{ImSch99} that
$$
V_j(t)=\sum_{n=1}^d \Big( \int_0^t V_n(s) D_n F_j(x_s,\dd s)
+ \int_0^t D_i\phi_s^n(y)\big( D_n F_j(x_s,\dd s)- D_n F_j(y_s,\dd s)\big) \Big).
$$
Then for $p \ge 2$ we have
\begin{align*}
\Big( {\mathbf E}&\big( \sum_{j=1}^d \sup_{0 \le t \le T} V_j^2(t) \big)^{p/2} \Big)^{2/p}
\le \sum_{j=1}^d \big( {\mathbf E} \sup_{0 \le t \le T} V_j^p(t)\big)^{2/p}\\
&\le \sum_{j=1}^d \Big( \sum_{n=1}^d \Big({\mathbf E} \sup_{0 \le t \le T} \Big| \int_0^t V_n(s) D_n M_j(x_s,\dd s) \Big|^p \Big)^{1/p}\\
&\hspace{.4cm}+ \sum_{n=1}^d \Big({\mathbf E} \sup_{0 \le t \le T} \Big| \int_0^t D_i\phi_s^n(y)
\big( D_n M_j(x_s,\dd s)- D_n M_j(y_s,\dd s)\big) \Big|^p \Big)^{1/p} \\
&\hspace{.4cm}+ \sum_{n=1}^d \Big({\mathbf E} \sup_{0 \le t \le T} \Big| \int_0^t V_n(s) D_n b_j(x_s,s) \dd s \Big|^p \Big)^{1/p}\\
&\hspace{.4cm}+ \sum_{n=1}^d\Big( {\mathbf E} \sup_{0 \le t \le T} \Big| \int_0^t D_i\phi_s^n(y)
\big( D_n b_j(x_s,s)\dd s- D_n b_j(y_s,s)\big) \dd s\big) \Big|^p \Big)^{1/p} \Big)^2.
\end{align*}
We have by Burkholder's inequality
\begin{align*}
{\mathbf E} \sup_{0 \le t \le T} \Big| \int_0^t V_n(s) D_n M_j(x_s,\dd s) \Big|^p
\le& C_p {\mathbf E} \Big( \Big| \int_0^T V_n^2(t) D_n D_n a_{jj}(x_t,x_t,t) \dd t\Big|^{p/2}\Big)\\
\le& C_p k_1^{p/2} {\mathbf E} \Big( \Big| \int_0^T V_n^2(t) \dd t\Big|^{p/2}\Big)
\end{align*}
and
\begin{align*}
\Big({\mathbf E} &\sup_{0 \le t \le T} \Big|\int_0^t D_i\phi_s^n(y)
\big( D_n M_j(x_s,\dd s)- D_n M_j(y_s,\dd s)\big) \Big|^p\Big)^{1/p} \\
\le& C_p^{1/p} \Big({\mathbf E} \Big| \int_0^T (D_i\phi^n_s(y))^2 D_n D_n \mathcal{A}_{jj}(x_s,y_s,s) \dd s\Big|^{p/2}\Big)^{1/p}\\
\le& C_p^{1/p} k_2^{1/2} \Big({\mathbf E} \Big| \int_0^T (D_i\phi^n_s(y))^2 |\phi_s(x)-\phi_s(y) |^2 \dd s\Big|^{p/2} \Big)^{1/p}\\
\le& C_p^{1/p} k_2^{1/2} \Big| \int_0^T \big({\mathbf E}| D_i\phi^n_s(y)|^p |\phi_s(x)-\phi_s(y) |^p\big)^{2/p} \dd s\Big|^{1/2} \\
\le& C_p^{1/p} k_2^{1/2} \Big| \int_0^T \big({\mathbf E}| D_i\phi^n_s(y)|^{2p}\big)^{1/p} \big({\mathbf E} |\phi_s(x)-\phi_s(y) |^{2p}\big)^{1/p} \dd s\Big|^{1/2} \\
\le& C_p^{1/p} k_2^{1/2} \Big| \int_0^T \tilde f^2_{2p}(s)\bar{c}^2|x-y|^2\ee^{2(\Lambda+p\sigma^2)s} \dd s\Big|^{1/2}
\end{align*}
and
\begin{align*}
{\mathbf E} \sup_{0 \le t \le T} \Big| \int_0^t V_n(s) D_n b_j(x_s,s)\dd s \Big|^p \le k_3^p {\mathbf E} \Big( \int_0^T |V_n(s)| \dd s \Big)^p \\
\end{align*}
and
\begin{align*}
&\Big({\mathbf E} \sup_{0 \le t \le T} \Big| \int_0^t D_i\phi_s^n(y)
\big( D_n b_j(x_s,s)- D_n b_j(y_s,s)\big)\dd s \Big|^p \Big)^{1/p}\\
\le&k_4\Big({\mathbf E}\big(\int_0^T |D_i\phi^n_s(y) | |\phi_s(x)-\phi_s(y)| \dd s \big)^p \Big)^{1/p}
\leq k_4\bar{c}|x-y|\int_0^T\tilde f_{2p}(s)\ee^{(\Lambda+\sigma^2p)s}\dd s.
\end{align*}
Therefore, using the same estimates as in the first part of the proof, we get
\begin{align*}
g_p^2(T)\le&d\Big( d^{1-1/p}C_p^{1/p} k_1^{1/2} \Big( \int_0^T g_p^2(t) \dd t\Big)^{1/2}\\
& +dC_p^{1/p}\bar{c}|x-y| k_2^{1/2} \Big( \int_0^T \tilde f^2_{2p}(s)\ee^{2(\Lambda+p\sigma^2)s} \dd s\Big)^{1/2}\\
& +d^{1-1/p}k_3 \int_0^T g_p(s) \dd s +
dk_4\bar{c}|x-y|\int_0^T\tilde f_{2p}(s)\ee^{(\Lambda+\sigma^2p)s}\dd s \Big)^2\\
&\le \beta_1d^{3-2/p} C_p^{2/p} k_1 \int_0^T g^2_p(t) \dd t+ \beta_3 d^{3-2/p}k_3^2 \Big(\int_0^T g_p(t) \dd t\Big)^2 +H(T),
\end{align*}
where $H$ is as in the lemma. Therefore, \eqref{eq:ggleichung} follows from Lemma \ref{le:gronwall} and the proof is complete.
\end{proof}
\begin{theorem}\label{main}
Let $d \ge 2$ and let $\phi$ be a stochastic flow satisfying the assumptions of Lemma~\ref{le:twopointderivative}.
For any (deterministic, compact) subset ${\mathbb X}$ of $\ensuremath{\mathbf{R}^{d}}$ with
box dimension $\Delta \ge 0$, we have
$$
\limsup_{T \to \infty} \frac 1T \log \sup_{x \in {\mathbb X}} \sup_{0 \le t \le T} \|D\phi_{0,t}(x)\| \le \xi,
$$
where
$$\xi=
\left\{\begin{array}{ll}
\hat k & \mbox{ if }\; \hat k \ge cd+\hat c\\
\hat k + 2\sqrt{k\Delta \gamma_1}&\mbox{ if }\; \hat k \le cd+\hat c \;\mbox{ and }\;
2\sqrt{ck}\Delta d+cd^2-2c\Delta d +\Delta(\hat k-\hat c)\ge 0\\
\hat k + 2\sqrt{k\Delta \gamma_2} &\mbox{ if }\; \hat k \le cd+\hat c \;\mbox{ and } \;
2\sqrt{ck}\Delta d+cd^2-2c\Delta d +\Delta(\hat k-\hat c)\le 0,
\end{array}\right.
$$
where $\gamma_1$ and $\gamma_2$ are defined as in Theorem~\ref{th:zweites}
and
\begin{align*}
k&=\alpha_2d^4k_1k_5^2/2,\qquad \hat k= \sqrt{\alpha_3} d^2 k_3 + 2k\\
c&=\alpha_2 d^4 k_1 k_5^2 + \frac 12 \beta_1 d^3 k_1 k_5^2 + \sigma^2 ,\qquad \hat c= \sqrt{\alpha_3}d^2 k_3
+ \sqrt{\beta_3 d^3 k_3^2} +\Lambda.
\end{align*}
\end{theorem}
\begin{proof}
The proof is just a combination of Lemma~\ref{le:twopointderivative} and Theorem~\ref{th:zweites}.
Lemma~\ref{le:twopointderivative} tells us that
$$
\big({\mathbf E} \sup_{0 \le t \le T} \|D\phi_t(x)\|^p\big)^{1/p} \le \exp\{(kp+\tilde k)T + o(T)\}
$$
for all $p \ge 2$, where
$$
k=\alpha_2d^4k_1k_5^2/2,\qquad \tilde k= \sqrt{\alpha_3} d^2 k_3.
$$
In order to obtain an estimate for all $p \ge 0$, we define $\hat k:=2k+\tilde k$ and get
$$
\big({\mathbf E} \sup_{0 \le t \le T} \|D\phi_t(x)\|^p\big)^{1/p} \le \exp\{(kp+\hat k)T + o(T)\}
$$
for all $p \ge 0$.
Lemma~\ref{le:twopointderivative} tells us further that for $p\ge 2$ (and hence for $p>d$)
$$
{\mathbf E} \sup_{0 \le t \le T} \|D\phi_t(x)-D\phi_t(y)\|^p\le
|x-y|^p \exp\{ (cp^2+\hat c p)T+o(T)\},
$$
with $c,\bar c$ as in the theorem. Therefore, the assumptions of Theorem~\ref{th:zweites} hold
and the assertion follows.
\end{proof}
\begin{remark}
The formulas in Theorem \ref{main} still contain the numbers $\alpha_2,\,\alpha_3,\,\beta_1$ and $\beta_3$. Since
$\alpha_1, \, \beta_2$ and $\beta_4$ do not appear in the formulas, it is possible to choose
$\alpha_2=\alpha_3=\beta_1=\beta_3=2$ but a different choice may result in a sharper bound.
\end{remark}
\begin{remark}
One can establish a corresponding result also in case $d=1$ by adjusting $\hat c$ just like we adjusted $\hat k$ in order to extend the
moment bound on the differences of derivatives from $p \ge 2$ to $p > d=1$.
\end{remark}
\begin{remark}
It is not hard to establish a version of both Lemma~\ref{le:twopointderivative} and Theorem~\ref{main} for higher derivatives of
a stochastic flow using an induction proof along the lines of~\cite[Proposition 2.3]{ImSch99}.
\end{remark}
\section{Isotropic Brownian Flows}\label{se:IBF}
In this section we will specialize the results of Theorem \ref{main} to isotropic Brownian flows (IBFs). We will be able to
establish somewhat better upper bounds by exploiting -- for example -- an explicit representation of the growth
of the derivative of an IBF. We start by defining an IBF. We assume that $d \ge 2$.
\begin{defi}
A stochastic flow $\phi$ on ${\mathbf R}^d$ is called {\em isotropic Brownian flow} if it is generated by a semimartingale field
$F=M+V$ with $V \equiv 0$ and martingale field $M$ with
quadratic variation $a(x,y,t,\omega)=b(x-y)$, where $b:{\mathbf R}^d \to {\mathbf R}^{d \times d}$ is a deterministic function satisfying
\begin{itemize}
\item $x \mapsto b(x)$ is $C^4$.
\item $b(0)=\mathrm{id}_{{\mathbf R}^d}$.
\item $x \mapsto b(x)$ is not constant.
\item $b(x)=O^*b(Ox)O$ for any $x \in {\mathbf R}^d$ and any orthogonal matrix $O \in O(d)$.
\end{itemize}
\end{defi}
For an IBF, we define its {\em longitudinal} resp.~{\em normal} correlation functions by
$$
B_L(r):=b_{11}(r e_1),\qquad B_N(r):= b_{11}(r e_2), \quad r \ge 0,
$$
where $e_i$ denotes the $i^{\mathrm{th}}$ unit vector in ${\mathbf R}^d$ (1 and 2 can be replaced by any $i\neq j$ by isotropy).
We will need the following facts about IBFs.
\begin{itemize}
\item $\beta_L:=-B_L''(0)>0$, $\beta_N:=-B_N''(0)>0$.
\item For each $x \in {\mathbf R}^d$, we have $\lambda_1:=\frac 12 \big( (d-1)\beta_N-\beta_L \big)=\lim_{t \to \infty}\frac 1t \log \|D\phi_t(x)\|$
almost surely. The number $\lambda_1$ is called the {\em top Lyapunov exponent} of the IBF.
\item $b,B_L$ and $B_N$ are bounded with bounded derivatives up to order 2.
\item $k_1:=\max_{i,j}|D_{x_i}D_{y_i}b_{jj}(x-y)|_{x=y}=\beta_L \vee \beta_N$ (independently of $x$)
\item $F$ satisfies Hypothesis (A).
\end{itemize}
The first two of these facts can be found in \cite{BaH86}, $b$ is bounded since $b$ is a covariance function and
boundedness of the second derivatives (and therefore also of the first) follows from equation \eqref{cov}. The fourth item
follows from the definition of $\beta_L$ and $\beta_N$ and the final one from the boundedness of the second derivatives of $b$.
\begin{theorem}\label{iso}
Let $d \ge 2$ and let $\phi$ be an IBF. For any (deterministic, compact) subset ${\mathbb X}$ of $\ensuremath{\mathbf{R}^{d}}$ with
box dimension $\Delta \ge 0$, we have
$$
\limsup_{T \to \infty} \frac 1T \log \sup_{x \in {\mathbb X}} \sup_{0 \le t \le T} \|D\phi_{0,t}(x)\| \le \xi,
$$
where $\xi$ is defined as in Theorem \ref{th:zweites} with
$$
c=2 \beta_L + 10d^3(\beta_L \vee \beta_N), \quad \hat c=2\lambda_1,\quad k=\frac{\beta_L}2,\quad \hat k= \lambda_1^+.
$$
\end{theorem}
\begin{proof}
Defining $f_p$ and $\tilde f_p$ as in the previous section, we obtain by Lemma \ref{Ableitung}:
$$
f_p(t) \le \exp\Big\{\Big(\lambda_1^+ + \frac{\beta_L}2 p\Big) t + o(t)\Big\},
\qquad \tilde f_p(t) \le \exp\Big\{\Big(\lambda_1 + \frac{\beta_L}2 p\Big) t + o(t)\Big\}
$$
for all $p >0$. Next, we estimate $g_p$ according to formula \eqref{eq:ggleichung}. Observing that $\bar C_2=0$, $\Lambda=\lambda_1$
and
$\sigma=\sqrt{\beta_L}$ (by Lemma \ref{IBFesti}), we obtain
$$
g_p(t) \le |x-y| \exp\big\{\big(2\lambda_1+2\beta_Lp+ \frac 12 d^3(\beta_L \vee \beta_N) C_p^{2/p} \big)t +o(t) \big\}.
$$
Noting that $C_p^{1/p}\le 2\sqrt{5}p^{1/2} $ for $p \ge 2$ \cite[Proposition 4.2]{BY82}, the assertion follows from
Theorem \ref{th:zweites}.
\end{proof}
\begin{remark}
Even though we have no reason to believe that the bound in Theorem \ref{iso} is optimal in general, it is at least optimal in case
$\Delta=0$ (by Lemma \ref{Ableitung}).
\end{remark}
\section*{Appendix A: A Gronwall-type Lemma}
We include in this section an elementary lemma which we conjecture to be essentially well-known but which we could not
find in the literature in a version suitable for our needs.
\begin{lemma}\label{le:gronwall}
Let $f:[0,\infty) \to [0,\infty)$ be a locally integrable function and $H:[0,\infty) \to [0,\infty)$
non-decreasing such that
$$
f(t) \le C_1 \int_0^t f(s) \dd s + C_2 \Big( \int_0^t \sqrt{f(s)} \dd s \Big)^2 + H(t)
$$
for some $C_1,C_2\ge 0$ and all $t \ge 0$. Then
$$
f(t) \le H(t) + H(t)\frac{C_1+\sqrt{C_2}}{C_1+2\sqrt{C_2}} \Big(\exp\{(C_1+2\sqrt{C_2})t\}-1\Big)
$$
for all $t \ge 0$.
\end{lemma}
\begin{proof} Let $\lambda>0$ and $\gamma_t:=\lambda (1-\ee^{-\lambda t})^{-1}$ for $t>0$. Then, by Jensen's inequality, for $t>0$
\begin{align*}
f(t) &\le C_1 \int_0^t f(s) \dd s +
C_2 \Big( \int_0^t \sqrt{f(s)} \frac{ \ee^{\lambda (t-s)}}{\gamma_t} \gamma_t \ee^{-\lambda (t-s)} \dd s \Big)^2 + H(t)\\
&\le C_1 \int_0^t f(s) \dd s + C_2 \int_0^t f(s) \frac{ \ee^{\lambda (t-s)}}{\gamma_t} \dd s + H(t).
\end{align*}
Therefore,
\begin{align*}
\ee^{-\lambda t}f(t)&\le C_1 \ee^{-\lambda t} \int_0^t f(s) \dd s + \frac{C_2}{\lambda}\int_0^t f(s)
\ee^{-\lambda s} \dd s + H(t) \ee^{-\lambda t}\\
&\le C_1 \int_0^t \ee^{-\lambda s} f(s) \dd s + \frac{C_2}{\lambda}\int_0^t f(s)
\ee^{-\lambda s} \dd s + H(t) \ee^{-\lambda t}.
\end{align*}
Gronwall's Lemma implies
\begin{align*}
f(t) &\le \ee^{\lambda t} \Big(H(t) \ee^{-\lambda t} + \int_0^t H(s) \ee^{-\lambda s}\Big(C_1+\frac{C_2}{\lambda}\Big)
\ee^{\Big(C_1+\frac{C_2}{\lambda}\Big)(t-s)} \dd s \Big)\\
&= H(t) + \ee^{\lambda t} \int_0^t H(s) \ee^{-\lambda s}\Big(C_1+\frac{C_2}{\lambda}\Big) \ee^{\Big(C_1+\frac{C_2}{\lambda}\Big)(t-s)} \dd s.
\end{align*}
Choosing $\lambda= \sqrt{C_2}$ (which minimizes $\lambda + C_1 + \frac{C_2}{\lambda}$) and using the monotonicity of $H$, we obtain
$$
f(t) \le H(t) + H(t) \frac {C_1 + \sqrt{C_2}}{C_1 + 2\sqrt{C_2}} \Big( \ee^{(C_1 + 2\sqrt{C_2})t}-1\Big)
$$
as claimed in the lemma.
\end{proof}
\section*{Appendix B: Some Estimates for IBFs}
In this appendix, we collect three basic properties of IBFs which are used in Section \ref{se:IBF} and which
do not seem to have appeared in the literature so far.
\begin{lemma}\label{beta}
Let $\phi$ be an IBF with covariance tensor $b$. Then we obtain for the correlation functions $B_{N/L}$ defined in
Section \ref{se:IBF}
$$
1-B_L(r) \le \frac{\beta_L}{2} r^2 \mbox{ and } 1-B_N(r) \le \frac{\beta_N}{2} r^2
$$
for all $r \ge 0$.
\end{lemma}
\begin{proof}
Let $U(x)$, $x \in {\mathbf R}^d$ be an ${\mathbf R}^d$-valued centered Gaussian process with
$\mathrm{cov}(U_i(x),U_j(y)) = b_{ij}(x-y)$. Denoting the $i^{\mathrm{th}}$ unit coordinate vector by $\ee_i$ and using Schwarz' inequality,
we get
\begin{align}
B_L''(r)&=\lim_{h \to 0}\lim_{\delta \to 0} {\mathbf E} \Big(\frac{U_1(h\ee_1)-U_1(0)}{h} \frac{U_1(-(r+\delta)\ee_1)-U(-r\ee_1)}{\delta}\Big)\nonumber\\
&=-{\mathbf E}\Big( U_1'(r\ee_1)U_1'(0)\Big) \ge - {\mathbf E}\Big( U_1'(0)^2\Big) = B_L''(0).\label{cov}
\end{align}
Therefore, for each $r>0$ there exists some $\theta \in (0,r)$ such that
$$
B_L(r) = B_L(0) + \frac{1}{2} B_L''(\theta) r^2 \ge 1 + \frac 12 B_L''(0) r^2= 1-\frac 12 \beta_L r^2.
$$
The estimate for $B_N$ follows in the same way, so the assertion of the lemma follows.\\
\end{proof}
Observe that the following lemma holds for every IBF -- even if the top exponent $\lambda_1$ is negative.
\begin{lemma}\label{IBFesti} Let $\phi$ be an IBF. Let $x,y \in {\mathbf R}^d$, $x \neq y$
and let $\rho(t):=|\phi_t(x)-\phi_t(y)|$. Then
$$
{\mathbf E} \rho_t^q \le |x-y|^q \exp\Big\{ \Big(q\lambda_1 + q^2\frac{\beta_L}{2}\Big) t\Big\}
$$
for all $t\ge 0$ and all $q \ge 1$.
\end{lemma}
\begin{proof}
We know that
$$
\dd \rho_t = (d-1) \frac{1-B_N(\rho_t)}{\rho_t} \dd t + \sqrt{2(1-B_L(\rho_t))} \dd W_t.
$$
Let $q \ge 1$. It\^o's formula implies
\begin{align*}
\dd \rho_t^q &= q\rho_t^{q-1}\dd \rho_t + \frac{q (q-1)}{2} \rho_t^{q-2} \dd \langle \rho \rangle_t \\
&=q \rho_t^{q-2} \big( (d-1)(1-B_N(\rho_t))+(q-1)(1-B_L(\rho_t))\big) \dd t
+ q \rho_t^{q-1} \sqrt{2(1-B_L(\rho_t))} \dd W_t.
\end{align*}
By Lemma \ref{beta}, we get
$$
{\mathbf E} \rho_t^q \le |x-y|^q + q \big(\frac{\beta_N}{2}(d-1) + \frac{\beta_L}{2}(q-1)\big) \int_0^t {\mathbf E} \rho_s^q \dd s.
$$
Gronwall's Lemma, together with the formula for the top exponent $\lambda_1=(d-1)\beta_N/2-\beta_L/2$ imply the assertion.
\end{proof}
Next, we look at the derivative of an IBF. The formula in the next lemma becomes particularly nice if we use the following
{\em Schatten norm} of a $d\times d$-matrix $A=(a_{ij})$:
$$
\| A \|_S := \left( \sum_{i,j=1}^{d} \left( \sum_{k=1}^d a_{ik} a_{jk} \right)^2 \right)^{1/4}
= \left(\sum_{i=1}^d \sigma_i^4\right)^{1/4},
$$
where $\sigma_1,...,\sigma_d$ are the singular values of $A$.
\begin{lemma}\label{Ableitung}
Let $\phi$ be an IBF. Then, for each $x \in {\mathbf R}^d$, there exists a one-dimensional Wiener process $W$
such that for all $t \ge 0$,
$$
\|D\phi_t(x)\|_S = d^{1/4} \exp\{ \lambda_1t + \sqrt{\beta_L}W_t\}.
$$
\end{lemma}
\begin{proof}
Fix $x \in {\mathbf R}^d$ and define
$$
N_t:=\int_0^t \sum_{i,j,k} D_k M^i(\phi_s(x),\dd s) \frac{D_j\phi_s^j(x) D_j\phi_s^k(x) }{\|D\phi_s(x)\|_S}.
$$
Then it is easy to see (cf.~\cite[p. 101]{HvB10}) that
$$
\langle N \rangle_t = \beta_L t.
$$
Therefore, $W_t:=(\beta_L)^{-1/2} N_t$, $t \ge 0$ is a standard Brownian motion and applying It\^o's formula, we get
$$
\log\|D\phi_t(x)\|_S= \frac 14 \log d + N_t + \lambda_1 t = \frac 14 \log d + \sqrt{\beta_L} W_t + \lambda_1 t.
$$
Exponentiating this expression, the lemma follows.
\end{proof}
\bibliographystyle{abbrv}
|
Subsets and Splits
No community queries yet
The top public SQL queries from the community will appear here once available.