nlin0512060/r05.tex
1: 
2: \documentclass[12pt]{iopart}
3: 
4: \usepackage{iopams}  
5: \usepackage{bm}
6: \usepackage{graphicx}
7: %\usepackage{amsmath}
8: \usepackage{amssymb}
9: \usepackage{hyperref}
10: 
11: \def\la{\label}
12: %\def\be{\begin{equation}}
13: \def\beq{\begin{equation}}
14: \def\eeq{\end{equation}}
15: %\def\ee{\end{equation}}
16: \def\bea{\begin{eqnarray}}
17: \def\eea{\end{eqnarray}}
18: \def\p{\partial}
19: 
20: \begin{document}
21: 
22: \title{Relaxation of nonlinear oscillations in BCS superconductivity}
23: 
24: \author{Razvan Teodorescu}
25: 
26: \address{Physics Department, Columbia University, 538 West 120th Street, Mail Code 5293, New York, NY 10027}
27: \ead{rteodore@phys.columbia.edu}
28: 
29: \begin{abstract}
30: The diagonal case of the $sl(2)$ Richardson-Gaudin quantum pairing model \cite{Richardson1,Richardson2,Richardson3,Richardson4,Richardson5,Richardson6,Gaudin76} is known to be solvable as an Abel-Jacobi inversion problem \cite{SOV,Kuznetzov,Kuz1,Kuz2,Kuz3,Kuz4,Kuz5,YAKE04}. This is an isospectral (stationary) solution to a more general integrable hierarchy, in which the full time evolution can be written as isomonodromic deformations. Physically, the more general solution is appropriate when
31: the single-particle electronic spectrum is subject to external perturbations. The asymptotic behavior of the nonlinear oscillations in the case of elliptic solutions is derived.
32: 
33: \end{abstract}
34: 
35: 
36: %\pacs{05.30, 05.40, 05.45}
37: %\submitto{\JPA}
38: 
39: %\maketitle
40: 
41: \section{Introduction}
42:       \label{sec:intro}
43: 
44: The integrable system described by the pairing hamiltonian introduced by Richardson and Sherman \cite{Richardson1,Richardson2,Richardson3,Richardson4,Richardson5,Richardson6} in the context
45: of nuclear physics has received revived interest in recent years, after being applied to metallic  superconducting grains 
46: \cite{Sierra} and cold fermionic systems 
47: \cite{Levitov,YAKE04}. The model is intimately related \cite{CRS97,ALO21,DES01} to a class of integrable systems 
48: generally referred to as Gaudin magnets \cite{Gaudin76}. These systems have been 
49: studied both at quantum and classical level \cite{Dukelsky,AFS,Samtleben,Hikami,Gawedzki1,Gawedzki2,Ortiz,Korotkin,Harnad,Dilorenzo}, in the elliptic case as well as trigonometric and rational degenerations, using various methods  from integrable vertex models to singular limits of Chern-Simons theory. 
50: 
51: In \cite{Levitov}, an interesting regime of the pairing problem was considered, which may be relevant to recent experiments with cold fermionic
52: gases exhibiting the paired BCS state \cite{Fermi1,Fermi2,Fermi3,Fermi4,Fermi5,Fermi6,Fermi7,Fermi8}. It was shown 
53: that for such systems, the time scales of the order parameter $\tau_{\Delta}
54: \sim |\Delta|^{-1}$, and the quasiparticle energy relaxation time $\tau_{\epsilon}$ are both much larger than typical time for switching
55: on the pairing interaction $\tau_0$, essentially given by the variation of external parameters, such as detuning from the Feshbach resonance. It was 
56: argued that in this regime, for times $t \ll \tau_{\epsilon}$, the dynamics
57: of the system is given by non-linear, non-dissipative equations describing the coherent BCS fluctuations for the system out of equilibrium. In this limit, the system is integrable, and features non-perturbative behavior, such as soliton-type solutions.
58: 
59: In the mean-field limit, such non-trivial solutions describing the collective mode of the Anderson spins \cite{Anderson} were derived in \cite{Levitov}, for a two-level effective system. This work was generalized \cite{YAKE04} in algebro-geometrical terms. 
60: 
61: In \cite{Gurarie,Leggett,YTA05}, the long-time behavior of the solution has been considered, under various conditions. An issue not addressed so far is the relaxation of the nonlinear oscillatory solution induced by perturbations of the spectral curve, physically justified by coupling to the environment. Several possible kinds of perturbations may be considered, which may lead to different 
62: types of relaxation.
63:  
64: In this paper, we consider the effect of fluctuations of single-particle energy levels, which amount to slow deformations of the Liouville tori, and can be described by hydrodynamic-type equations in phase space \cite{Krichever,K1}. These equations  describe the evolution of the moduli for the complex curve of the system. 
65: 
66: \section{The Richardson-Gaudin Model}
67: 
68: \subsection{The quantum pairing hamiltonian}
69: 
70: 
71: Following \cite{Dukelsky}, we briefly review the Richardson pairing model.  
72: It describes a system of $n$ fermions characterized by a set of independent one-particle states of energies $\epsilon_l$, where the label $l$ takes values from a set $\Lambda$. The labels may refer, for instance, to orbital angular  momentum eigenstates. Each state $l$ has a total degeneracy $d_l$, and the states within the subspace corresponding to  $l$ are further labeled by an internal quantum number $s$. For instance, if the quantum number $l$ labels orbital momentum eigenstates, then $d_l = 2l+1$ and $s = -l, \ldots, l$. However, the internal degrees of freedom can be defined independently of $l$. We will assume
73: that $d_l$ is even, so for every state $(ls)$, there is another one related by time reversal symmetry $(l\bar s)$. For simplicity, we specialize to the case $d_l = 2, s= \uparrow, \downarrow $. Let $\hat c^{\dag}_{ls}$ represent the fermionic creation operator for the state $(ls)$. Using the Anderson pseudo-spin operators \cite{Anderson}  (quadratic pairing operators), satisfying the $su(2)$ algebra
74: \beq
75: [t^3_i, \, t_j^{\pm}] = \pm \delta_{ij} t_j^{\pm}, \,\,\,\, [t^+_i, \, t^-_j] = 2\delta_{ij}t^3_j,
76: \eeq
77: the Richardson pairing hamiltonian is given by
78: \beq \label{pairing}
79: H_P = 
80: \sum_{l \in \Lambda} 2\epsilon_l t^3_l - g\sum_{l, l'}t^+_l t^-_{l'} = 
81: \sum_{l \in \Lambda} 2\epsilon_l t^3_l - g{\bf{t}}^+\cdot {\bf{t}}^-, 
82: \eeq
83: where ${\bf {t}} = \sum_{l}{\bf {t}}_l$ is the total spin operator. 
84: It maps to the  reduced BCS model 
85: \beq \label{hamiltonian}
86: \hat H = \sum_{{\bf{p}}, \sigma}\epsilon_{\bf{p}}\hat{c}_{{\bf{p}}, \sigma}^{\dag}\hat{c}_{{\bf{p}}, \sigma}
87: -g\sum_{{\bf{p}}, {\bf{k}}}\hat{c}_{{\bf{p}} \, \uparrow}^{\dag}\hat{c}_{{-\bf{p}}\, \downarrow}^{\dag}
88: \hat{c}_{{-\bf{k}}\, \downarrow}\hat{c}_{{\bf{k}}\, \uparrow}
89: \eeq
90: by replacing the translational degrees of freedom by rotational ones, where $l  \in \Lambda = \{ 1, \ldots n \}$ ennumerates the one-particle orbital degrees of freedom, while $s = \, \uparrow, \, \downarrow$ indicates the two internal spin states per orbital ($d_l=2$). The pairing hamiltonian can be decomposed into the linear combination
91: \beq
92: H_P=2\sum_{l \in \Lambda} \epsilon_l R_l + g \left [ \left ( \sum_{l \in \Lambda} t^3_l \right )^2 - \frac{1}{4} \sum_{l \in \Lambda} (d_l^2-1) \right ].
93: \eeq
94: At a fixed value of the component $t^3$ of the total angular momentum, the last term becomes a constant and is dropped from the 
95: hamiltonian. The operators $R_l$ (generalized Gaudin magnets \cite{Gaudin76}) are given by
96: \beq \label{gaudin}
97: R_l = t^3_l - \frac{g}{2}\sum_{l' \ne l}\frac{{\bf {t}}_l \cdot {\bf {t}}_{l'}}{\epsilon_l - \epsilon_{l'}}.
98: \eeq
99: These operators solve the Richardson pairing hamiltonian because \cite{CRS97} they are independent, commute with
100: each other, and span all the degrees of freedom of the system. Richardson showed \cite{Richardson1,Richardson2} that the exact $N-$pair wavefunction 
101: of his hamiltonian is given by application of operators 
102: %\beq
103: $
104: b^{\dag}_{k} = \sum_{l}\frac{t_l^{\dag}}{2\epsilon_l - e_k}
105: $
106: %\eeq 
107: to vacuum (zero pairs state). The unnormalized $N-$pair wavefunction reads
108: %\beq
109: $
110: \Psi_R(\epsilon_i) = \prod_{k=1}^Nb_k^{\dag}|0\rangle.
111: $
112: %\eeq 
113: The eigenvalues $e_k$ satisfy the self-consistent algebraic equations 
114: \beq \la{ec}
115: \frac{1}{g} = \sum_{p \ne k}\frac{2}{e_k - e_p} +\sum_l \frac{1}{2\epsilon_l-e_k},
116: \eeq which can be given a 2D electrostatic interpretation \cite{Dukelsky}
117: with energy
118: \begin{eqnarray} \label{electrostatic}
119: U(\epsilon_l, e_k) = \frac{2}{g} \left [ \sum_{k=1}^N 
120: \mathcal{R}e (e_k) -
121: \sum_{l=1}^n \mathcal{R}e (\epsilon_l) \right ] + \\
122: 2\sum_{l=1}^n\sum_{k=1}^N \log|e_k -2\epsilon_l| - 
123: 4\sum_{k < p}\log|e_k-e_p|- 
124: \sum_{i<j}\log|2\epsilon_i-2\epsilon_j|
125: \end{eqnarray}
126: Equations (\ref{ec}) appear as equilibrium conditions for a set of charges of strength $q=2$ placed at points $e_k$, in the presence of fixed charges of strength $q=-1$ at points 
127: $2\epsilon_l$, and uniform electric field of strength $\frac{1}{g}$, pointing along the real axis. This interpretation proves to be very useful for the conformal field theory (CFT) description of the Richardson problem. The electrostatic energy 
128: (\ref{electrostatic}) is minimized for values $\{ e_k\}$ corresponding to pair energies. In (\ref{electrostatic}), $n, N$ represent the number of single-particle levels and the number of pair energies, respectively. 
129: For large interaction constant $g$, the equilibrium positions $\{ e_k \}$ form a set of complex conjugated pairs defining a curve $\gamma$ in the complex plane of energies. We note that the eigenvalues $r_i$ of Gaudin hamiltonia are proportional in this language to the values of the electric field at positions $2\epsilon_i$, $2r_i = g\frac{\p U}{\p \epsilon_i}. $
130: 
131: For a set of single-particle energies $\{\epsilon_i \}$, the BCS ground state is obtained by minimizing the electrostatic energy (\ref{electrostatic}) with respect to positions of
132: the free charges at $\{ e_k \}$. Once found,
133: they also determine exactly the values of the electric field at positions
134: $\{ 2\epsilon_i \}$ on the real axis, which are proportional to 
135: the ground-state eigenvalues  $\{ r^{GS}_i\}$. For any other values of $\{ r_i \}$, the electrostatic energy 
136: (\ref{electrostatic}) is not minimized. This indicates
137: that for arbitrary values $r_i \ne r_i^{GS}$, the system is not in equilibrium.
138: 
139: \subsection{The mean field limit of Richardson-Gaudin models}
140: 
141: \paragraph*{General description of the classical model} 
142: In  the mean-field limit,
143: the spin operators ${\bf {t}}_l$ are replaced by their quantum mechanical averages. Written in terms of the classical vectors ${\bf {S}}_l = 2 \langle {\bf {t}}_l \rangle$, the semiclassical approximation for the pairing 
144: hamiltonian becomes
145: \beq \label{meanfield}
146: H_{MF} = \sum_{l \in \Lambda}\epsilon_l S^3_l - \frac{g}{4}|J^-|^2,
147: \eeq 
148: where ${\bf {J}} = \sum_{l \in \Lambda} {\bf {S}}_l$ and the BCS gap function is given by $\Delta = gJ^-/2$. Replacing commutators  by canonical Poisson brackets, 
149: \beq \label{poisson}
150: \{ S^{\alpha}_i, \, S^{\beta}_j \} = 2\epsilon^{\alpha \beta \gamma}S^{\gamma}_i \delta_{ij},
151: \eeq
152: variables $S^{\alpha}_i$ become smooth functions of time. In this limit, 
153: the problem can analyzed with tools of classical integrable systems, and the 
154: solution is known to be exact as $n \to \infty$. 
155: 
156: The Poisson brakets (\ref{poisson}) and hamiltonian 
157: (\ref{meanfield}) lead to the equations of motion
158: \beq \label{bloch}
159: \dot{\vec{{S}}}_i = 2(-\vec{\Delta} + \epsilon_i \hat{z}) \times \vec{S}_i,
160: \eeq
161: where $2\vec{\Delta} = (gJ_x, gJ_y, 0)$ and $\vec{J}$ is the total spin. The 
162: semiclassical limit of Gaudin hamiltonia are independent constants of motion,
163: \beq
164: r_i = \frac{1}{2}\left [S_i^z - \frac{g}{2}\sum_{j \ne i}
165: \frac{\vec{S}_i \cdot \vec{S}_j}{\epsilon_i -\epsilon_j} \right ],
166: \quad
167: %\{r_i, \, r_j \} =0, \quad
168: %\eeq
169: %\beq
170: \dot r_i 
171: %= \{ 2\sum_j \epsilon_j r_j, \, r_i \} 
172: = 0.
173: \eeq
174: Equations (\ref{bloch}) describe a set of strongly interacting spins and have
175: generic non-linear oscillatory solutions. The exact solution may be obtained through the Abel-Jacobi inverse map \cite{Kuznetzov,Dickey,YAKE04}. 
176: 
177: This solution can be described exactly in the language of hyperelliptic Riemann surfaces (see Appendix A for details). At this  point, it is useful to make use of the intuitively clear features of this construction (Figure~\ref{oscilat}). For a given set of initial conditions for the spins 
178: $\{ \vec{S}_i \}$, $i.e.$ also of the constants of motion $\{ r_i \}$, a
179: polynomial $Q(u)$ of degree $2n$ and with $n$ pairs of complex conjugated
180: roots $E_{2k+2}=\overline{E}_{2k+1}, \, k=0, \ldots, n-1$, is constructed. 
181: A schematic representation of these roots is given in Figure~\ref{oscilat}. 
182: Between each pair of roots, we place a simple cut 
183: $\mathcal{C}_k = [E_{2k+1}, \, E_{2k+2}]$ 
184: on the complex plane of energies. The surface thus obtained is a representation 
185: of a torus of smooth genus $g=n-1$.  
186: 
187: Variables $u_k$ are introduced for $n-1$ of these cuts, with respect to which the equations of motion separate. The variables $u_k$ evolve in time in a complicated fashion, solving a system of nonlinear coupled differential equations (\ref{dubrovin1}). Up to a constant, the time dependence of the gap parameter amplitude is given by 
188: \beq
189: \log |\Delta(t)| = \mathcal{I}m \, \int u(t) dt, \quad u(t) = 
190: \sum_k u_k (t). 
191: \eeq
192: The widths of the cuts $\mathcal{C}_k$ and the 
193: periods of the non-linear oscillators $u_k$ are determined by the values of
194: constants of motion $\{ r_i \}$. For the particular choice $r_i =r_i^{GS}$, all
195: the cuts $\mathcal{C}_k, \, k=1, \ldots , n-1$ vanish, and the width of the remaining cut $\mathcal{C}_n$ equals the equilibrium value of the gap function:
196: $|E^{GS}_{2n-1}-E^{GS}_{2n}| = 2 |\Delta|^{GS}$. In that case, the oscillators
197: $u_k=E_{2k-1}=E_{2k}$ are at rest, and the only time dependence left in the system is the uniform precession of the parallel planar spins $S^{-}_i$, with 
198: frequency $\omega = 2\sum_{k=1}^n\epsilon_k - 2\sum_{p=1}^{n-1}E_{2p-1}-\sum_{i=1}^nS^z_i$. In the case of particle-hole symmetry, $\omega$ vanishes as well. 
199: \begin{figure} \begin{center}
200:  \includegraphics*[width=8cm]     {oscillators.eps}
201: \caption{\label{oscilat}
202: Schematic representation of the Liouville torus for the integrable
203: system (\ref{bloch}).
204: }
205: \end{center} \end{figure}
206: 
207: \subsection{The onset of pairing interaction and long-time behavior of oscillations}
208: Equations (\ref{bloch}) do not impose any particular 
209: constraints on the integrals of motion. These constants are known for the metallic state $t < 0$, but change abruptly   
210: during the short interval when the electron-electron interaction 
211: is turned on. The onset of pairing interaction is a delicate problem in itself \cite{Gurarie,Leggett,YTA05}, which deserves further investigation. There are several reasons which make the issue non-trivial. We review here several of them. 
212: \begin{itemize} 
213: \item[a)] In the case of KdV and KP2 hierarchies, it is known that finite-gap solutions (for which only a few cuts $\mathcal{C}_k$ are non-degenerate) are dense in the space of all periodic solutions. A conjecture of Krichever \cite{K1} extends this fact to arbitrary algebraic curves. Therefore, solutions where an infinite number of independent frequencies contribute to the mean-field approximation require very special initial conditions. The system is most likely to be described either by 
214: (i) a finite-gap solution, or (ii) a non-oscillatory function. The case (ii)
215: has not been investigated in previous studies. 
216: \item[b)] Another issue related to solutions described by only a few frequencies is 
217: that the $quantum$ integrability of the system may impose additional constraints and provide a selection criterion, non-existent in the mean-field approximation. The exact solution of quantum $XXX$ Gaudin system under a sudden variation of the 
218: spectral curve is a subject of active research and will be addressed elsewhere 
219: \cite{us}.  
220: \item[c)] Concerning the finite-gap solutions of the mean-field approximation, we note that the onset of attractive interaction is realized through some mechanism 
221: which effectively changes the parameters of the spectral curve. It is reasonable 
222: to assume that this variation is not instantaneous and in fact may continue as a
223: perturbation for the rest of the evolution of the system. This mechanism describes 
224: coupling of the system to the environment (the external magnetic field and optical trap, for degenerate Fermi gases) and therefore may induce fluctuations in the parameters describing the spectral curve. We therefore consider the effect of such 
225: perturbations which do not change the monodromy of the solution. The concept 
226: of isomonodromic deformations of integrable systems was introduced in \cite{JM1,JM2,JM3} and used extensively in the theory of Painlev\'e transcendents. We give here a brief review of the method. 
227: \end{itemize}
228: 
229: We consider the unperturbed problem given by a nonlinear scalar differential equation $R(u, \dot u, \ddot u, \ldots) = 0$, where $R$ is a rational function of the solution 
230: $u(t)$ and its derivatives $\dot u, \ddot u, \ldots$. Specializing 
231: to the cases described by hyperelliptic spectral curves, we express it through the Lax pair of $2 \times 2$ operators $L, \, A$, solving the linear vector problem
232: \beq \la{isospectral}
233: L\Psi = \mu \Psi, \,\, \p_t \Psi = A\Psi, \,\, \p_t L = [A, \, L],
234: \eeq  
235: where $\lambda, \mu$ are two auxiliary complex  variables, $A(\lambda, u)$ a $2 \times 2$ matrix, and $\Psi(\lambda, \mu, t)$ 
236: is a column vector. The operators are chosen such that elements of the identity $\dot L = [A, L]$ are equivalent with $R(u, \dot u, \ddot u, \ldots) = 0$. A trivial calculation gives the isospectral property $\dot \mu = 0$. 
237: 
238: Specializing the operator $L$ to have the form \cite{Mumford} (the $u$ functional dependence is implicit throughout) 
239: \beq \la{mumford}
240: L(\lambda) = \left [
241: \begin{array}{cc}
242: a(\lambda) & \quad b(\lambda) \\
243: c(\lambda) & -a(\lambda) 
244: \end{array}
245: \right ],
246: \eeq 
247: the eigenvalue equation for $\Psi$ becomes
248: \beq \la{crv}
249: \mu^2 + \det L = 0, \,\, 
250: \mu(\lambda) = \pm i \sqrt{\det L (\lambda)}.
251: \eeq 
252: When functions $a,b,c$ in (\ref{mumford}) are rational, equation 
253: (\ref{crv}) defines the hyperelliptic Riemann surface $\mu(\lambda)$
254: called $spectral$ $curve$ of the system (\ref{isospectral}). As noted before, the time evolution leaves the spectral curve invariant, which is why the problem is sometimes called {\it{isospectral}}.
255:  
256: The isomonodromic deformation \cite{K1} is introduced through
257: \beq \la{deform}
258: L\Psi = \mu\Psi + \epsilon \p_\lambda \Psi, \quad
259: [\epsilon\p_\lambda - L, \, \p_t -A] = 0,
260: \eeq 
261: where the second equation is the isomonodromy requirement, and $0 \le \epsilon \ll 1$. We recover
262: the unperturbed problem by setting $\epsilon = 0$. The first equation
263: is easily integrated and gives the formal solutions
264: \beq
265: \log \Psi_{\pm}(\lambda, \mu, t) = 
266: \frac{1}{\epsilon} \left [-
267: \mu \lambda \pm i \int ^\lambda \sqrt{\det L(\sigma)} d\sigma
268: \right ], 
269: \eeq
270: up to constants in $\lambda$. Let us now impose the saddle-point 
271: (or turning-point) condition $\p_\lambda \log \Psi = 0$. This will
272: give $\mu = \pm i \sqrt{\det L}$, $i.e.$ the spectral curve. Therefore, we may see the isospectral problem as a saddle-point (turning point) approximation. The compatibility equations
273: \beq
274: [\epsilon\p_\lambda - L, \, \p_t -A] = 0
275: \eeq
276: can be recast in the form 
277: \beq \la{split}
278: \epsilon(\p_\nu L - \p_\lambda A) + \p_\tau L -[A, L] = 0,
279: \eeq
280: where we have split the time dependence $\p_t$ into a fast time scale
281: $\p_\tau$ and a slow one $\epsilon \p_\nu$. At zero order  in $\epsilon$, (\ref{split}) is simply the unperturbed problem. The first order correction gives the slow-time scale dependence of the modulated solution $u_\epsilon(\tau, \nu)$. Moreover, from the matrix elements of (\ref{split}), we get after averaging over the fast motion in $\tau$ \cite{Takasaki},
282: \beq \la{det}
283: \p_\nu \det L = -(\overline{2aA_{11}+bA_{21}+cA_{12}}),
284: \eeq
285: where the bar signifies $\tau-$averaging. Since for the unperturbed problem, $\p_\tau \det L = 0$, the averaging is justified.
286: 
287: Equation (\ref{det}) is simply the result of Bogoliubov-Whitham \cite{Whitham} averaging for the problem (\ref{deform}). It tells us how the previously invariant spectral curve now changes slowly in time over the large time scale $\nu$. It also gives us the form of the
288: deformed solution $u_{\epsilon}(\tau, k_i(\nu))$, as a {\it{modulation}} of the original solution.  
289: 
290: \paragraph*{Note} The fact that such perturbations generalize the autonomous  Garnier system of \cite{YAKE04} to a non-autonomous system of Schlesinger type was indicated in \cite{Takasaki}. Extensions of these systems to include a constant matrix were discussed in \cite{Beauville}. .
291: 
292: \section{The effect of weak perturbations}
293: 
294: \subsection{Topological classification of finite-gap solutions} \la{whith}
295:  
296: In the presence of spectrum symmetry $\epsilon_k = - \epsilon_{-k}, S_k^z = -S_{-k}^z$, the 
297: distribution of cuts $\mathcal{C}_k$ obeys the same symmetry. The number of non-degenerate 
298: cuts is therefore even. This analysis uses the fact that at $t=0$ all Anderson pseudo-spins are aligned along the $z$ axis. In the simplest non-trivial case, there are only two non-degenerate cuts as 
299: shown in Figure~\ref{genus1} and $g=1$, while the corresponding variable $u_1$ is given by elliptic functions. Other non-trivial cuts $\mathcal{C}_k$ may exist in general, associated with dynamics of variables $u_k$. In the limit of ``small" cuts \cite{Bobenko}, their contributions 
300: separate are given by trigonometric functions and the behavior of the gap parameter takes the 
301: simplified form
302: \beq \la{expansion}
303: \log \frac{|\Delta(t)|}{|\Delta(0)|} = \mathcal{I}m \, \int u(t) dt = \mathcal{U}_{ell}(t) + \sum_k
304: \mathcal{U}^k_{trig}.
305: \eeq
306: As we shall see, the most relevant contribution is due to the elliptic part $\mathcal{U}_{ell}$, studied in the next section.
307: 
308: \subsection{Modulations of elliptic solutions}
309: 
310: For the root distribution shown in Figure~\ref{genus1}, there is only one
311: variable $u_k$, taking imaginary values \cite{YAKE04}. It 
312: solves an equation of the type \cite{YAKE04}
313: \beq
314: (\dot u)^2 + (u^2 + m^2)(u^2 + M^2) = 0.
315: \eeq
316: For simplicity, we therefore make the transformation $u \to iu$, keeping the time variable real. The new function is a real quartic oscillator which satisfies
317: \beq \la{osc}
318: (\dot u)^2 = (u^2 - M^2)(u^2 - m^2).
319: \eeq 
320: The solutions corresponding to this distribution of roots are 
321: \begin{eqnarray} \label{u1}
322: u_1(t) = m\cdot sn (Mt + \phi_1, k_1), \quad
323: k_1 = m/M, \\
324: u_2(t) = M\cdot sn (mt + \phi_2, k_2), \quad \,
325: k_2 = M/m, 
326: \end{eqnarray}
327: where $sn$ is the Jacobi sine function, and $\phi_{1,2}$ 
328: are arbitrary phases. In the degenerate case $m=M$, the 
329: solutions become hyperbolic functions. Solution $u_2$ is non-physical in our case. 
330: 
331: 
332: 
333: \begin{figure} \begin{center}
334:  \includegraphics*[width=8cm]     {genus1.eps}
335: \caption{\label{genus1}
336: Distribution of complex roots $E_i$ around the origin, 
337: for the elliptic approximation. The mean level spacing is $d$.}
338: \end{center} \end{figure}
339: 
340: 
341: In order to set up the isomonodromic deformation method, consider the Lax pair \cite{Its}:
342: \begin{eqnarray}
343: \la{L1}
344: L = -\left [ 
345: 2\dot u \sigma_1 +
346: 4u\lambda \sigma_2 +
347: (4\lambda^2 - \xi  + 2u^2)i\sigma_3
348: \right ], \\
349: \la{L2}
350: A = i\lambda \sigma_3 - u\sigma_2,
351: \,\,\,\,\,\,\,
352: \epsilon \p_\lambda \Psi = L \Psi, \,\,\,\,\,\,\,
353: \p_t  \Psi = A \Psi, 
354: \end{eqnarray}
355: where $\sigma_\alpha$, $\alpha = 1, 2, 3$ are the Pauli 
356: matrices, $\epsilon$ is a small real number, 
357: and $\Psi = (\psi_1, \, \psi_2)^t$ is the Baker-Akhiezer
358: function. Choosing the initial condition 
359: $\psi_1 = 0, \, \psi_2 = 1$, we can extract the amplitude of the ratio $\Delta(t)/\Delta(0)$ from $\Psi(t)$ as
360: \beq
361: \left | \frac{\Delta(t)}{\Delta(0)} \right | = \frac{
362: \left [ \psi_1(t) + \overline{\psi}_1(t)\right ]
363: }{2}.
364: \eeq
365: The compatibility (zero-curvature) conditions
366: \beq \label{compat}
367: [\p_t - A , \,  \epsilon \p_\lambda - L] = 0
368: \eeq
369: yield the system of equations
370: \beq
371: \p_t \xi = \epsilon, \quad \p_t^2 u = 2u^3-\xi u.
372: \eeq 
373: We note that
374: \beq
375: \det L = -4 \{ (\dot u)^2 - [ u^4 - \xi u^2 + 
376: (2\lambda^2 - \xi/2)^2] \}.  
377: \eeq
378: Setting $\epsilon = 0$ gives the equation $(\dot u)^2 - u^4 + \xi u^2 = $ constant, while the limit $\epsilon = 1$ yields the Painlev\'e II equation. In fact, the unperturbed case $\epsilon = 0$ allows
379: to retrieve the full elliptic solution, from the equation
380: \beq
381: L\Psi =0, \quad \det(L) = 0,
382: \eeq  
383: which gives the elliptical function $u$ satisfying
384: \beq
385: (\dot u)^2 - [ u^4 - \xi u^2 + (2\lambda^2 - \xi/2)^2] = 0. 
386: \eeq
387: The physical solution $u_1$ is obtained identifying
388: \beq
389: \xi = m^2+M^2, \quad 2 \lambda = M + m.  
390: \eeq
391: Now restore $\epsilon \ne 0$, write $\p_t = \p_{\tau} 
392: +\epsilon \p_{\xi}$, and retain terms of order $\epsilon$
393: from (\ref{compat}). Averaging over the fast variable $\tau$ 
394: gives
395: \beq
396: \p_\xi \det L = 
397: \overline{L_{22}\p_\lambda A_{11}} +
398: \overline{L_{11}\p_\lambda A_{22}}
399: ,
400: \eeq
401: or equivalently,
402: \beq
403: \p_{\xi} \det L 
404: = -(4\overline{u^2} - 2\xi +8\lambda^2).
405: \eeq
406: Performing the computations, we obtain
407: \beq
408: \p_\xi [u^4 -\xi u^2 -(\dot u)^2] = - \overline{u^2}.   
409: \eeq
410: Writing the elliptic equation as
411: \beq \la{c1}
412: (\dot u)^2 = u^4 -\xi u^2 + \mu^2,
413: \eeq
414: the physical solution takes the form
415: %\beq \label{sol}
416: $$
417: u_1(t) = \sqrt{\frac{\xi - \sqrt{\xi^2 - 4\mu^2}}{2}}
418: sn \left [\sqrt{\frac{\xi + \sqrt{\xi^2 - 4\mu^2}}{2}}t \right ],
419: $$
420: %\eeq
421: up to an arbitrary initial phase $\phi$, and elliptic modulus
422: \beq
423: k^2 = \frac{1 - \sqrt{1-(\frac{2\mu}{\xi})^2}}
424: {1 + \sqrt{1-(\frac{2\mu}{\xi})^2}}.
425: \eeq
426: The Whitham averaging equation has the form
427: \beq \label{whitham}
428: 4\frac{\p \mu^2}{\p \xi^2} =1-
429: \frac{2}{2-k^2}\frac{\mathcal{E}(k^2)}{\mathcal{K}(k^2)},
430: \eeq
431: where $\mathcal{E, K}$ are the complete elliptic integrals of the
432: first and second kind, respectively. 
433: Equation (\ref{whitham}) has a fixed point at $k = 0$, $\frac{\mu}{\xi} \to 0$. This shows that, on the slow time scale,
434: the parameter $\mu/\xi = m/M + O(m^3/M^3)$ goes to zero, as $\xi = m^2 + M^2$ increases. Expanding the solution 
435: in this limit, and integrating under the separation of time scales 
436: asumption, we obtain for the elliptic contribution to the gap parameter, the approximation
437: \beq
438: \Delta_{ell}(t) = \Delta(0) e^{k \int sn(\tau, k^2 ) d\tau},
439: \eeq
440: and $k^2 = \frac{\mu^2}{\xi^2} + O(\mu^4/\xi^4) \to 0$ as $t \to \infty$. 
441: 
442: \paragraph*{Asymptotic behavior of modulated elliptic solutions}
443: Starting from the Lax pair (\ref{L1}, \ref{L2}), we can obtain 
444: the asymptotic behavior of $\Psi(\lambda, t)$ as $t \to \infty$, 
445: in the whole complex $\lambda$ plane \cite{Its}. The analysis is
446: simpler when working in the variable $z = \frac{\lambda}{\sqrt{\xi}}$. There are six Stokes sectors at $z \to \infty$, with canonical
447: asymptotes for $\Psi$, but the region of interest is $z = 1/2$, 
448: where 
449: \beq
450: \frac{4\lambda^2}{\xi} \to 1, \quad \frac{\mu^2}{\xi^2} \to 0.
451: \eeq
452: Using the Whitham equation, separation of scales, and identifying the strength of the fluctuations $\epsilon$ with an effective temperature $T$, we obtain for the elliptic contribution 
453: at large times (\ref{expansion})
454: \beq
455: U_{ell}(t) = \frac{\mathcal{F}(t)}{tT},
456: \eeq
457: where $\mathcal{F}$ is a bounded oscillatory function 
458: $|\mathcal{F}| = O(1)$. 
459: 
460: \paragraph*{Modulations of trigonometric solutions}
461: In order to analyze the slow dynamics of the small cuts, 
462: let $\xi \to \infty$ in (\ref{c1}) and write the solution as
463: \beq \la{trg}
464: u = \frac{\mu}{\sqrt{\xi}} \cos(\sqrt{\xi}t + \phi), 
465: \quad \mu^2/\xi \to 0.
466: \eeq
467: This assumption is consistent with (\ref{c1}). Formally, this is simply the trigonometric 
468: approximation of the general elliptic solution, in
469: the limit of small modulus $k^2$. However, the parameter $\xi$ in (\ref{trg}) must be sent to $\infty$ mush faster than the physical $\xi = M^2 + m^2$, in order to give the correct spherical limit.
470:  The contributions from the small cuts therefore vanish faster
471: than the elliptic component.
472: 
473: \paragraph*{Higher-genus contributions}
474: In the case where there are several non-degenerate cuts at $t=0$ corresponding to a high 
475: genus $g>1$, the isomodromic deformation method can be applied in the same way as for $g=1$, 
476: leading to modulated hyper-elliptic solutions of Painlev\'e equations degree higher than 2. There
477: are few systematic results in this field, and a complete classification does not exist at this time. 
478: Specific high-genus solutions may have interesting topological properties with physical 
479: interpretations in terms of the collective excitations of the Gaudin magnet \cite{us}.  
480: 
481: 
482: \appendix{} \la{Lax}
483: 
484: \section{Isospectral case and the Abel-Jacobi inversion problem}
485: 
486: In \cite{Kuznetzov,YAKE04}, the system (\ref{bloch}) with fixed spectral curve was solved through inverse Abel-Jacobi mapping, by using Sklyanin separation of variables techniques \cite{SOV,Smirnov}. 
487: Interesting connections to generalized Neumann systems and Hitchin systems were discovered in \cite{Hikami}.
488: The solution starts from the Lax operator 
489: \beq \label{lax} 
490: \mathcal{L}(\lambda) = \frac{2}{g}\sigma_3 +\sum_{i=0}^n \frac{\vec S_i \cdot\vec \sigma}{\lambda - \epsilon_i} = 
491: \left [
492: \begin{array}{cr}
493: a(\lambda) & \,\, \, b(\lambda) \\
494: c(\lambda) & -a(\lambda)
495: \end{array}
496: \right ],
497: \eeq
498: where $\sigma_\alpha, \, \alpha = 1,2,3$ are the Pauli matrices, and $\lambda$ is an additional complex variable, the spectral 
499: parameter. Let $u_k, \, k=1, \ldots, n-1$ be the roots of the coefficient $c(\lambda)$. 
500: Poisson brackets for variables $S_i^{\alpha}$ read 
501: \beq
502: \{S_j^\alpha , \, S_k^{\beta} \}=2\epsilon_{\alpha \beta \gamma}S_k^\gamma \delta_{jk}
503: \eeq
504: The Lax operator (\ref{lax}) defines 
505: a Riemann surface (the spectral curve) $\Gamma (y, \lambda)$ of genus 
506: $g = n-1$, through 
507: \beq \label{curve}
508: y^2 = Q(\lambda) = \det \mathcal{L}(\lambda)\left [ g\frac{P(\lambda)}{2} \right ]^2 , 
509: \eeq
510: where $P(\lambda) = \prod_{i=1}^n (\lambda -  \epsilon_i).$
511: 
512: The equations of motion for the hamiltonian (\ref{meanfield}) become
513: \beq \label{dubrovin1}
514: \dot  u_i = \frac{2i y( u_i)}{\prod_{j \ne i}( u_i - u_j)},
515: \quad
516: i\dot J^- = J^{-} \left [ gJ^3 + 2\sum_{k=1}^{n} \epsilon_k - u \right ].
517: \eeq
518: In (\ref{dubrovin1}), 
519: $u = -2\sum_{i=1}^{n-1}  u_i, \quad b(u_i) = 0$. 
520: 
521: From the equations of motion, it is clear that knowledge of the initial amplitude of $J^-$ and of the roots $ \{  u_i \}$
522: is enough to specify the $n$ unit vectors $\{ {\bf {S}}_i \},$ for a given set of constants of motion $\{ R_l \}$ given by the classical limit of Gaudin hamiltonia. The Dubrovin equations (\ref{dubrovin1}) are 
523: solved by the inverse of the Abel-Jacobi map, as we explain in the following. We begin by noting that the polynomial 
524: $Q(\lambda)$ has degree $2n$, and is positively defined on the real $\lambda$ axis. Therefore, the curve $\Gamma (y, \lambda)$
525: has $n$ cuts between the pairs of complex roots $[E_{2i-1}, \, E_{2i}], i = 1, 2, \ldots, n$, perpendicular to the real $\lambda$
526: axis. The points $u_i$ belong to $n-1$ of these cuts, $u_i \in [E_{2i-1}, \, E_{2i}], i =1, \ldots, n-1$. These 
527: $g = n-1$ cuts allow to define a canonical homology basis of $\Gamma$, consisting of cycles $\{\alpha_i, \beta_i \}, i = 1, \ldots, g$. With respect to these cycles, a basis of normalized holomorphic differentials $\{ \omega_i \}$ can be defined, through
528: \beq
529: \mu_i = \lambda^{g-i}\frac{d\lambda}{y}, \,\,\, M_{ij} = \int_{\alpha_j} \mu_i, \,\,\, {\bm {\omega}} = M^{-1}{\bm{\mu}}. 
530: \eeq 
531: The period matrix $B_{ij} = \int_{\beta_j} \omega_i$
532: is symmetric and has positively defined imaginary part. The Riemann $\theta$ function is defined with the help of the period matrix as
533: \beq
534: \theta({\bm{z}} | B) = \sum_{{\bm{n}} \in {\bm{z}}^g} e^{2\pi i ({\bm {n}}^t {\bm{z}} + \frac{1}{2} {\bm {n}}^t B {\bm {n}} )}.
535: \eeq
536: The $g$ vectors $\bm{B}_k$ consisting of columns of $B$ and the basic
537: vectors $\bm{e}_k$  define a lattice in ${\mathbb{C}}^g$. The $Jacobian$ variety of the curve $\Gamma$, is then the $g-$dimensional torus defined as the quotien 
538: $J(\Gamma)  = \mathbb{C}^g /(\mathbb{Z}^g + B\mathbb{Z}^g).$
539: The Abel-Jacobi map associates to any point $P$ on $\Gamma$, a point ($g-$ dimensional complex vector) on the Jacobian variety, through
540: ${\bm {A}} (P) = \int_{\infty}^P {\bm {\omega}}.$
541: Considering now a $g-$dimensional complex vector of points $\{P_k \}, 
542: k = 1, \ldots, g$ on $\Gamma$, defined up to a permutation, we can associate to it the point on the Jacobian
543: \beq  \label{a}
544: {\bm{z}} = {\bm{a}}({\bm {P}}) = \sum_{k=1}^g {\bm {A}} (P_k) + {\bm {K}},
545: \eeq
546: where ${\bm {K}}$ is the Riemann characteristic vector for $\Gamma$. 
547: 
548: The map (\ref{a}) suggests that we now have a 
549: way to describe the dynamics on $\Gamma$ by following the image point on the Jacobian. 
550: Given a point on the $g-$dimensional Jacobian  ${\bm{z}} = (\zeta_1, \ldots , \zeta_{n-1})$, we can find an unique set of points $\{ \lambda_k\}, k = 1, \ldots, g$ on $\Gamma$, such that ${\bm{z}} = {\bm {a}}(
551: {\bm {\lambda}}),$ and $\theta ({\bm{a}(\bm{P}) - \bm{z}} | B) = 0$. The system evolves in time according to the point ${\bm{z}}(t)$ 
552: \beq \label{solution1}
553: \zeta_k = ic_k, \,\, 1 \le i \le g-1, \,\, \zeta_{n-1} = i(c_{n-1} + t),
554: \eeq
555: where $\{ c_k\}$ is a set of initial conditions, such that  $
556: {\bm{z}}_0 = {\bm{z}}(t=0) = {\bm {a}}({\bm{c}})$, and ${\bm{c}}$ is the set of initial
557: conditions for positions of ${\bm {\lambda}}$ on $\Gamma$. Together with the initial condition which determines the
558: initial amplitude of $J^-$, this set will determine entirely the evolution of the functions $ u_i(t), J^-(t)$. 
559: 
560: 
561: \subsection*{Acknowledgments}
562: 
563: The author is grateful to I~Aleiner, I~Gruzberg, I~Krichever and P~Wiegmann for suggestions and contributions. Useful discussions with A~G~Abanov, B~Altshuler, E~Bettelheim, and E~Yuzbashyan are acknowledged. The author also thanks I~Aleiner and A~Millis for support. 
564: 
565: 
566: \subsection*{References}
567: 
568: \bibliography{references}
569: \bibliographystyle{unsrt}
570: 
571: \end{document}
572: 
573: 
574: 
575: