cond-mat0311355/md2.tex
1: \documentclass[aps,floats,twocolumn]{revtex4}
2: %\documentclass[aps,prl,preprint]{revtex4}
3: 
4: %\usepackage[T1]{fontenc}R
5: \usepackage{amssymb}
6: \usepackage{amsbsy}
7: \usepackage{epsfig}
8: \usepackage{amsmath}
9: \hyphenation{whe-ther ferm-ion-ic}
10: 
11: \newcommand{\be}{\begin{equation}}
12: \newcommand{\ee}{\end{equation}}
13: \newcommand{\bea}{\begin{eqnarray}}
14: \newcommand{\eea}{\end{eqnarray}}
15: 
16: \newcommand{\Ham}{H}
17: 
18: %----------------------------------------------------------------------------
19: %---------------------------------------------------------------------------
20: 
21: \begin{document}
22: 
23: \title{Quench dynamics across quantum critical points}
24: \author{K. Sengupta, Stephen Powell, and Subir Sachdev}
25: \affiliation{Department of Physics, Yale University, P.O. Box
26: 208120, New Haven, Connecticut 06520-8120}
27: \date{\today}
28: 
29: %--------------------------------------------------------------------------
30: 
31: \begin{abstract}
32: We study the quantum dynamics of a number of model systems as
33: their coupling constants are changed rapidly across a quantum
34: critical point. The primary motivation is provided by the recent
35: experiments of Greiner {\em et al.} (Nature {\bf 415}, 39 (2002))
36: who studied the response of a Mott insulator of ultracold atoms in
37: an optical lattice to a strong potential gradient. In a previous
38: work, it had been argued that the resonant response observed at a
39: critical potential gradient could be understood by proximity to an
40: Ising quantum critical point describing the onset of density wave
41: order. Here we obtain numerical results on the evolution of the
42: density wave order as the potential gradient is scanned across the
43: quantum critical point. This is supplemented by studies of the
44: integrable quantum Ising spin chain in a transverse field, where
45: we obtain exact results for the evolution of the Ising order
46: correlations under a time-dependent transverse field. We also
47: study the evolution of transverse superfluid order in the three
48: dimensional case. In all cases, the order parameter is best
49: enhanced in the vicinity of the quantum critical point.
50: \end{abstract}
51: 
52: \pacs{}
53: 
54: \maketitle
55: 
56: \section{Introduction}
57: \label{sec:intro}
58: 
59: Recent experiments with ultracold atoms have achieved reversible
60: tuning of bosonic atoms between superfluid and Mott insulating
61: states by varying the strength of periodic potential produced by
62: standing laser light \cite{Bloch,Kasevich}. The physics of such
63: ultracold atoms in the Mott insulating state can be described by
64: bosonic Hubbard model, well known in context of other condensed
65: matter systems \cite{fwgf,Sachdev1}. However, ultracold atoms in
66: optical lattices offer much better control over microscopic
67: parameters of the model. Consequently, it is possible to explore
68: parameter regimes which are not available in other analogous
69: condensed matter systems.
70: 
71: This paper will focus on a particular experiment reported by
72: Greiner {\em et al.} \cite{Bloch}. With the boson system in the
73: Mott insulating state, they applied a steep potential gradient to
74: the lattice, and observed its response. In a typical condensed
75: matter system, one might have expected a response analogous to
76: that of a sliding charge density wave: no motion of atoms until a
77: critical tilt was applied, and a sliding motion at all tilts above
78: the critical tilt. However, the experiment observed strikingly
79: different behavior: there was a strong {\em resonant} response in
80: the vicinity of tilts where the potential energy drop between
81: nearest neighbor optical lattice sites ($E$) equaled the repulsion
82: between two atoms on the same site ($U$). For $E\sim U$, applying
83: the tilt produced a noticeable change in the ground state, but (in
84: contrast to sliding charge density wave systems), there was little
85: change in the ground state for larger $E$ until a second resonant
86: peak at $E \sim 2U$. This resonant response is a clear indication
87: that the atoms experience little extrinsic dissipation, and their
88: dynamics should be described by an energy-conserving quantum
89: Hamiltonian.
90: 
91: A framework for describing the experiments of Greiner {\em et al.}
92: \cite{Bloch} was proposed in Ref.~\onlinecite{Sachdev2} (hereafter
93: referred to as I). (We also note here the numerical studies of
94: Braun-Munzinger {\em et al.} \cite{burnett} which addressed these
95: experiments by studying the time evolution of the underlying
96: Bose-Hubbard model.) For $w, |E-U| \ll E,U$, where $w$ is the
97: tunnelling matrix element between nearest neighbor lattice sites,
98: it was argued that we need only focus on a set of states which
99: were {\em resonantly coupled} to the original Mott insulating
100: state. In one dimension, the resonant subspace could be described
101: simply in terms of nearest neighbor {\em dipole} states,
102: consisting of a particle and a hole excitation about the Mott
103: insulator on nearest neighbor states; in higher-dimensions, the
104: particle and hole were no longer constrained to be on
105: nearest-neighbor sites but could reside anywhere on planes
106: orthogonal to the potential gradient, but separated by a single
107: lattice spacing. An effective Hamiltonian on such resonant
108: subspaces was proposed in I, and its phase diagram was presented.
109: In the regime of large potential gradient $E - U > w$, this
110: effective Hamiltonian possessed ground states with density wave
111: order with a period of 2 lattice spacings (see also
112: Ref.~\onlinecite{fendley} for conditions under which other periods
113: may obtain). It was argued in I that the proximity of the quantum
114: critical point, associated with the onset of this density wave
115: order, was responsible for the resonant response observed by
116: Greiner {\em et al.}.
117: 
118: The tilt experiments of Greiner {\em et al.} were carried out in
119: highly non-equilibrium situations, and the approach of I was to
120: describe these, to the extent possible, by an equilibrium analysis
121: of an effective Hamiltonian describing the primary states accessed
122: over the experimental time scale. The purpose of the present paper
123: is to directly address the non-equilibrium dynamics of the tilted
124: Mott insulator. We will mainly do this using the effective
125: Hamiltonian of I. The specific question we shall address is the
126: following. Begin with the system in the ground state in a regime
127: of small $E=E_i$ where there is no density wave order. Then,
128: suddenly change the value of $E$ to a $E=E_f$, including values
129: such that the ground state has density wave order at $E_f$. Allow
130: the system to evolve under the resulting Hamiltonian. What is the
131: nature of the state to which the system evolves at long times? We
132: will find, as conjectured in I, that the density wave order that
133: develops under this dynamic evolution is most robust when $E_f$ is
134: near the quantum critical point.
135: 
136: We will also address a similar question for the Ising chain in a
137: transverse field, $g$. Like the models of I, this model also has a
138: regime, $g< g_c$, where the ground state has spontaneous Ising
139: order. However, this much simpler model is completely integrable,
140: and so offers an opportunity to analyze the non-equilibrium
141: dynamics exactly. We initialize the Ising model in the ground
142: state in a transverse field $g_i > g_c$. The transverse field is
143: then changed rapidly to $g=g_f$, and the wavefunction evolves at
144: this field. We will compute equal-time correlations in this
145: wavefunction as a function of the time $t$, including in the $t
146: \rightarrow \infty$ limit. In some cases, exact closed-form
147: results will be obtained. The structure of these correlations as a
148: function of $g_f$ bear some similarity to the results of the model
149: of I as a function of $E_f$; however, there are some interesting
150: differences which, we suspect, are related to the integrability of
151: the Ising chain.
152: 
153: We now outline the remainder of the paper. In Section~\ref{oned}
154: we present numerical results on  the dynamics of the
155: one-dimensional dipole model of I. Section~\ref{sec:ising} will
156: address the non-equilibrium dynamics of the Ising chain in a
157: transverse field: this analysis uses the Jordan-Wigner
158: transformation, and obtains the required dynamic correlation
159: functions in the form of Toeplitz determinants.
160: Section~\ref{threed} returns to the model of I, but turns to the
161: dynamics in three dimensions; here, we use a combination of
162: mean-field theory and exact diagonalization to obtain results
163: similar to those in Section~\ref{oned}, but with the order
164: parameter now being a `transverse superfluid' order. We review the
165: results and discuss implications for experiments in
166: Section~\ref{conc}.
167: 
168: \section{Dipole dynamics in one dimension}
169: \label{oned}
170: 
171: This section will describe our numerical results on the quantum
172: dynamics of the one-dimensional dipole model of the Mott insulator
173: in a potential gradient.
174: 
175: Starting from a parent Mott state with $n_0$ bosons per site, we
176: identified the set of states which are resonantly coupled to the
177: parent Mott state when $U \sim E$ (recall that $U$ is the
178: repulsive energy between two bosons on the same site, and $E$, the
179: `electric field', is the potential drop between two nearest
180: neighbor sites). In one dimension, the resonant subspace involves
181: dipole states consisting of quasihole-quasiparticle pairs at
182: adjacent sites, and the low energy behavior of the system can be
183: described by the effective dipole Hamiltonian obtained in I:
184: \begin{eqnarray}
185: H_{\rm 1D}[E] &=& -w \sqrt{n_0(n_0+1)}\sum_\ell (d_\ell^\dagger +
186: d_\ell)
187: \nonumber\\
188: && +(U-E) \sum_{\ell} d_\ell^{\dagger} d_\ell . \label{ham1d}
189: \end{eqnarray}
190: The dipoles are subject to hardcore constraints that there is
191: never more than a single dipole on any pair of nearest neighbor
192: sites
193: \begin{equation}
194: d_\ell^{\dagger} d_\ell \le
195: 1~~~;~~~d_{\ell+1}^{\dagger}d_{\ell+1}d_\ell^{\dagger} d_\ell = 0.
196: \end{equation}
197: When the electric field $E$ is adiabatically tuned through $U$,
198: the ground state of the system changes from one with no dipoles
199: ($U\gg E$) to one with maximum possible number of dipoles ($E\gg
200: U$). At an intermediate critical electric field
201: \begin{equation}
202: E_c = U + 1.310 w \sqrt{n_0 (n_0 + 1)},
203: \end{equation}
204: the system undergoes a quantum phase transition in the Ising
205: universality class.
206: 
207: As discussed in Section~\ref{sec:intro}, we study the dynamics of
208: the ultracold atoms when the potential gradient is changed
209: suddenly. Such a situation can be very easily achieved
210: experimentally in these systems by rapidly shifting the center of
211: the confining magnetic trap. We shall specifically consider the
212: situation where the change in the potential gradient is fast
213: enough for the sudden perturbation assumption to be valid but slow
214: enough to restrict the dynamics within the resonant subspaces so
215: that the Hamiltonians (\ref{ham1d}) (and (\ref{ham3d}) in
216: Section~\ref{threed}) are still valid.
217: 
218: We assume that the atoms in the 1D lattice are initially in the
219: ground state $|\Psi_G\rangle $ of the dipole Hamiltonian
220: (\ref{ham1d}) with $E=E_{i} \ll E_c$. This ground state
221: corresponds to dipole vacuum. Consider shifting the center of the
222: magnetic trap so that the new potential gradient is $E_{f}$. If
223: this change is done suddenly, the system initially remains in the
224: old ground state. The state of the system at time $t$ is therefore
225: given by
226: \begin{eqnarray}
227: |\Psi(t)\rangle  =\sum_n c_n \exp(-i\epsilon_n t/\hbar) |n\rangle
228: , \label{state}
229: \end{eqnarray}
230: where $|n\rangle $ denotes the complete set of energy eigenstates
231: of the Hamiltonian $H_{\rm 1D}[E_{f}]$ in (\ref{ham1d}),
232: $\epsilon_n = \,\,\langle n|H_{\rm 1D}[E_{f}]|n\rangle $ is the
233: energy eigenvalue corresponding to state $|n\rangle $, and $c_n
234: =\,\,\langle n|\Psi(t=0)\rangle  = \langle n|\Psi_G\rangle $
235: denotes the overlap of the old ground state with the state
236: $|n\rangle $. Notice that the state $|\Psi(t)\rangle $ is no
237: longer the ground state of the new Hamiltonian. Furthermore, in
238: the absence of any dissipative mechanism, which is the case for
239: ultracold atoms in optical lattices, $|\Psi(t)\rangle $ will never
240: reach the ground state of the new Hamiltonian. Rather, in general,
241: we expect the system to thermalize at long enough times, so that
242: the correlations are similar to those of $H_{\rm 1D} [E_f]$ at
243: some finite temperature.
244: 
245: We are now in a position to study the dynamics of the Ising
246: density wave order parameter
247: \begin{equation}
248: O = \frac{1}{N}\langle \Psi|\sum_\ell (-1)^\ell d_\ell^{\dagger}
249: d_\ell|\Psi\rangle, \label{isingO}
250: \end{equation}
251: where $N$ is the number of sites. The time evolution of $O$ is
252: given by
253: \begin{eqnarray}
254: O(t) &=& \frac{1}{N} \sum_{m,n} c_m  c_n \cos\left[\left( E_m-E_n
255: \right)t/\hbar \right] \nonumber\\
256: && \times  \langle m|\sum_\ell (-1)^\ell d_\ell^{\dagger}
257: d_\ell|n\rangle \label{ising}
258: \end{eqnarray}
259: Eq.\ \ref{ising} is solved numerically using exact diagonalization
260: to obtain the eigenstates and eigenvalues of the Hamiltonian
261: $H_{\rm 1D}[E_{f}]$. Before resorting to numerics, it is however
262: useful to discuss the behavior of $O(t)$ qualitatively. We note
263: that if $E_f$ is close to $E_i$, the old ground state will have a
264: large overlap with new one $\it i.e.$ $c_m \sim \delta_{m1}$.
265: Hence in this case we expect $O(t)$ to have small oscillations
266: about $O(t=0)$. On the other hand, if $E_f \gg E_c$, the two
267: ground states will have very little overlap, and we again expect
268: $O(t)$ to have a small oscillation amplitude. This situation is in
269: stark contrast with the adiabatic turning on of the potential
270: gradient, where the systems always remain in the ground state of
271: the new Hamiltonian $H_{\rm 1D}[E_f]$, and therefore has a maximal
272: value of $\langle O\rangle $ for $E_f \gg E_c$. In between, for
273: $E_f \sim E_c$, the ground state $|\Psi\rangle $ has a finite
274: overlap with many states $|m\rangle $, and hence we expect $O(t)$
275: to display significant oscillations. Furthermore, if the symmetry
276: between the two Ising ordered states is broken slightly (as is the
277: case in our studies below), the time average value of $O(t)$ will
278: be non-zero.
279: 
280: This qualitative discussion is supported by numerical calculations
281: on finite size systems for system size $N=9,11,13$. For numerical
282: computations with finite systems, we choose systems with an odd
283: number of sites and open boundary conditions, so that dipole
284: formation on odd sites is favored, thus breaking the $Z_2$
285: symmetry. The results are shown in Figs.\ \ref{fig1}-\
286: \ref{fig3a}.
287: % ************************************************************
288: \begin{figure}
289:         \centerline{\epsfig{file=fig1.eps,width=8cm,angle=0}}
290:         \caption{Evolution of the Ising order parameter in (\protect\ref{isingO})
291:         under the Hamiltonian $H_{\rm 1D} [E_f]$ for $n_0 = 1$.
292:         The initial state is the ground state of $H_{\rm 1D} [E_i
293:         ]$. All the plots in this section have
294:         $U=40$, $w=1$, and $E_i=32$, and consequently the
295:         equilibrium quantum critical point is at $E_c = 41.85$.}
296: \label{fig1}
297: \end{figure}
298: % **********************************************************
299: % ************************************************************
300: \begin{figure}
301:         \centerline{\epsfig{file=fig2.eps,width=8cm,angle=0}}
302:         \caption{System size ($N$) dependence of the results of Fig~\protect\ref{fig1} for
303:         $E_f=40$. The curves are labelled by the value of $N$. }
304:          \label{fig2}
305: \end{figure}
306: % ************************************************************
307: % ************************************************************
308: \begin{figure}
309:         \centerline{\epsfig{file=fig3.eps,width=8cm,angle=0}}
310:         \caption{The curve labelled `dynamic' is the long time
311:         limit, $\langle O\rangle_t$
312:         of the Ising order in (\protect\ref{ising}) as a function
313:         of $E_f$ (for $N=11$), with other parameters as in
314:         Fig~\protect\ref{fig1}. This long time limit can be obtained
315:         simply by setting $m=n$ in (\protect\ref{ising}).
316:         For comparison, in the curve
317:         labelled `adiabatic', we show the expectation value of the
318:         Ising order $O$ in the ground state of $H_{\rm 1D} [E_f]$;
319:         such an order would be observed if the value of $E$ was
320:         changed adiabatically. Note that the dynamic curve has its
321:         maximal value near (but not exactly at) the equilibrium quantum critical point
322:         $E_c = 41.85$, where the system is able to respond most easily to
323:         the change in value of $E$; this dynamic curve is our theory of
324:         the `resonant' response in the experiments of Ref.~\protect\onlinecite{Bloch}
325:         discussed in Section~\ref{sec:intro}. In contrast the adiabatic
326:         result {\em increases monotonically} with $E_f$ into the $E>E_c$ phase where
327:         the Ising symmetry is spontaneously broken.}
328:         \label{fig3}
329: \end{figure}
330: % ************************************************************
331: % ************************************************************
332: \begin{figure}
333:         \centerline{\epsfig{file=ltop2.eps,width=8cm,angle=0}}
334:         \caption{Size dependence of the `dynamic' results in
335:         Fig~\protect\ref{fig3}. The sizes range from $N=7$ to $N=15$ (as labeled),
336:         with the intermediate values $N=9,11,13$:
337:         $\langle O\rangle_t$ decreases monotonically with $N$}
338:         \label{fig3a}
339: \end{figure}
340: % ************************************************************
341: Fig.\ \ref{fig1} shows the oscillation of the order parameter
342: $O(t)$ for different values of $E_f$ for $N=13$. In agreement with
343: our qualitative expectations, the oscillations have maximum
344: amplitude when $E_f \approx 40$ is near the critical value $E_c =
345: 41.85$. For either $E_f \ll E_c$ or $E_f \gg E_c$, the
346: oscillations have a small amplitude around $O(t=0)$. Furthermore,
347: it is only for $E_f \approx E_c$ that the time-average value of
348: $O(t)$ is appreciable. Fig.\ \ref{fig2} shows the system size
349: dependence of the time evolution for $E_f = U=40$. We find that
350: the oscillations remain visible as we go to higher system sizes,
351: although they do weaken somewhat. More significantly, the time
352: average value of $O(t)$ remains non-zero, and has a weaker
353: decrease with system size. In Fig.\ \ref{fig3}, we plot the long
354: time limit of the Ising order parameter, $\langle O\rangle _t$, as
355: a function of $E_f$, and compare it with the $O_{\rm ad}$, the
356: value of the order parameter when $E$ reaches $E_f$ adiabatically
357: and the wavefunction is that of the ground state at $E=E_f$. We
358: find that $\langle O\rangle _t$ stays close to $O_{\rm ad}$ as
359: long as there is a large overlap with between the old and the new
360: ground states. However, as we approach the adiabatic phase
361: transition point, this overlap decreases and $\langle O\rangle _t$
362: can not follow $O_{\rm ad}$ any more. The deviation of $\langle
363: O\rangle _t$ is therefore a signature that the system is now in a
364: different phase for the new value of the electric field.
365: 
366: The `dynamic' curve in Fig~\ref{fig3} should be compared with Figs
367: 5e,f in Ref.~\onlinecite{Bloch}. The latter show that the Mott
368: insulator has a resonantly strong response to an applied potential
369: gradient $E \sim U$. Here, we have found a similar resonant
370: enhancement in a simple model system in one dimension, induced by
371: the proximity of a quantum critical point.
372: 
373: We comment briefly on the nature of the thermodynamic limit, $N
374: \rightarrow \infty$ for the results in Fig~\ref{fig2},~\ref{fig3}.
375: For $O_{\rm ad}$ it is clear that there is a non-zero limit only
376: for $E > E_c$, when it equals the order parameter of the
377: spontaneously broken Ising symmetry. If we assume that the system
378: thermalizes at long times for the dynamic case, then $\langle O
379: \rangle_t$ corresponds to the expectation value of the equilibrium
380: order parameter in $H_{1D} [E_f]$ at some finite temperature. In
381: one dimension, it is not possible to break a discrete symmetry at
382: finite temperatures, and so the thermodynamic limit of the order
383: parameter must always vanish. By this reasoning, we expect
384: $\langle O \rangle_t$ to also vanish in the thermodynamic limit.
385: This is consistent with the results in Fig.\ \ref{fig3a}, where we
386: show the $N$ dependence of the long time limit $\langle O
387: \rangle_t$. Our data are at present not extensive enough to
388: definitely characterize the dependence of $\langle O \rangle_t$ on
389: $N$.
390: 
391: 
392: \section{Dynamics of the quantum Ising chain}
393: \label{sec:ising}
394: 
395: As a complement to the physically relevant, but numerical,
396: computations in Section~\ref{oned}, this section will describe
397: similar results in a simpler, analytically tractable model. We
398: will consider the integrable Ising chain in a transverse field,
399: which also has a zero temperature, quantum phase transition
400: between a phase with a broken $Z_2$ symmetry and a symmetric
401: phase. We will address questions on the evolution of the
402: wavefunction under a time-dependent change in the transverse
403: field.
404: 
405: The model of interest in this section is
406: \begin{equation}
407: H_I = -J \sum_{j} \left( \sigma^z_j \sigma^z_{j+1} + g(t)
408: \sigma^x_i \right), \label{HI}
409: \end{equation}
410: where $\sigma^{x,z}_j$ are Pauli matrices acting on a `spin' on
411: the sites, $j$, of an infinite chain. We have allowed the
412: transverse field to acquire an arbitrary time dependence $g(t)$.
413: We will mainly consider here the case of a sudden change at time
414: $t=0$ from an initial value $g(0^-) = g_i$ to a final value
415: $g(0^+)=g_f$, but our methods easily generalize to the arbitrary
416: time dependence in $g(t)$.
417: 
418: For time-independent $g(t)$, $H_I$ has a quantum critical point at
419: $g=g_c = 1$, with two equivalent ground states for $g< g_c$
420: related by a global $Z_2$ spin-flip. However, unlike
421: Section~\ref{oned} we will not introduce any external perturbation
422: which introduces a preference between these two states: all such
423: perturbations destroy the integrability of $H_I$. Consequently, we
424: do not obtain any useful information from the analog of the
425: time-dependence of the order parameter in (\ref{isingO}),
426: (\ref{ising}), as these quantities will be identically zero at all
427: times. Rather, we will compute here the two-point correlation
428: function of the order parameter in an infinite chain, which is
429: \begin{equation}
430: G_n (t) = \langle \psi (t) | \sigma^z_j \sigma^z_{j+n} | \psi (t)
431: \rangle. \label{defgn}
432: \end{equation}
433: Here $|\psi (t) \rangle$ is the state of the system at time $t$,
434: evolving under the Sch\"odinger equation specified by the
435: time-dependent Hamiltonian $H_I$. In equilibrium, the information
436: contained in a correlation function like (\ref{defgn}) is related
437: to an observable like that in (\ref{ising}) (which is the response
438: in the Ising order parameter to perturbations in the boundary
439: condition) by the fluctuation-dissipation theorem. However, we are
440: not aware of any analog of such a theorem for the non-equilibrium
441: case under consideration here, and so are not able to directly
442: relate the results of the present section to those of
443: Section~\ref{oned}.
444: 
445: Our analysis of $H_I$ proceeds with the standard Jordan-Wigner
446: transformation, and we follow the notation and methods of Chapter
447: 4 of Ref.~\onlinecite{Sachdev1}. We express the $S=1/2$ states in
448: terms of those of the spinless Jordan-Wigner fermion $c_j$, and
449: after transforming to momentum space fermions $c_k$, the
450: Hamiltonian becomes
451: \begin{eqnarray}
452: H_I &=& J \sum_k\left[ 2\left(g-\cos k \right) c_k^{\dagger} c_k
453: \right. \nonumber \\ &~&~~~~~~~~~~\left. - i \sin k
454: \left(c^{\dagger}_{-k} c_{k}^{\dagger} + c_{-k} c_{k} \right) -g
455: \right].
456: \end{eqnarray}
457: Now, transforming to the Heisenberg picture, we can follow the
458: evolution of the system by solving the equations of motion
459: \begin{equation}
460: \frac{d c_k}{dt} = i \left[ H_I , c_k \right].
461: \end{equation}
462: These equations are easily solved by a Bogoliubov transformation.
463: Finally, the correlator in (\ref{defgn}) is computed by a simple
464: generalization of the methods appropriate for the equilibrium
465: case. A few details of such a computation appear in the Appendix.
466: 
467: Here, we discuss the results for $G_n (t)$ for the case of a
468: sudden change from $g (0^-) = g_i$ to $g(0^+ ) = g_f$. For $t< 0$,
469: we assume the system is in the ground state appropriate for
470: $g=g_i$, and consequently $G_n (t<0)$ is independent of $t$ and
471: equal to the well-known equilibrium result at $g=g_i$. For $t>0$,
472: there is a non-trivial time dependence, and it is possible to
473: obtain the general expression for $G_n (t)$ as described in the
474: Appendix. We will restrict our attention here to the simpler
475: expression of the long time limit $G_n (t \rightarrow \infty)$,
476: which is the primary quantity of physical interest. For this, we
477: obtain the Toeplitz determinant
478: \begin{equation} G_n (\infty) =
479: \begin{vmatrix}
480: a_0 & a_{-1} & \ldots & a_{-n+1}\\
481: a_1 & a_0 & \ldots & a_{-n+2}\\
482: \vdots & \vdots & \ddots & \vdots\\
483: a_{n-1} & a_{n-2} & \ldots & a_0
484: \end{vmatrix},
485: \label{toep}
486: \end{equation}
487: where
488: \begin{equation}
489: a_r = \frac{1}{2\pi}\int_{-\pi}^{\pi} e^{-ikr} \tilde{a}(k),
490: \end{equation}
491: with
492: \begin{eqnarray}
493: \tilde{a}(k) &=& \frac{2(g_f g_i + 1)z - (g_f+g_i)(z^2+1)}{2(z-g_f)} \nonumber \\
494: &~&~~~~~~~~~~~\times \sqrt{\frac{z}{(z-g_i)(1-z g_i)}},
495: \label{tak}
496: \end{eqnarray}
497: where $z = e^{ik}$.
498: 
499: We now need to evaluate the $n \times n$ Toeplitz determinant in
500: (\ref{toep}), especially for the case of large $n$. In the
501: equilibrium situation, this is aided by Szeg\"o's lemma, and its
502: generalization in the Fisher-Hartwig formula \cite{fh}. For the
503: present situation, the expression in (\ref{tak}) does not obey the
504: winding number constraint required for application of the
505: Fisher-Hartwig formula, and so we are unable to take advantage of
506: this result. However, we shall show that an exact evaluation of
507: (\ref{toep}) is possible for two important special cases ($g_i =
508: \infty$ and $g_i = 0$), and supplement these by numerical
509: evaluation of (\ref{toep}) for other values of $g_i$.
510: 
511: In the case $g_i = 0$, we have
512: \begin{equation}
513: \tilde{a}(k) = \frac{2z - g_f (z^2+1)}{2(z-g_f)}
514: \end{equation}
515: and it is straightforward to evaluate $a_r$ by contour
516: integration. This gives
517: \begin{equation}
518: \begin{array}{c|c|c}
519:  & g_f < 1 & g_f > 1\\ \hline \rule[-4mm]{0mm}{11mm}
520: r \le -1 &  \displaystyle \frac{g_f^{-r}}{2} (1-g_f^2) & 0\\
521: \hline \rule[-4mm]{0mm}{11mm} r = 0 & 1 -  \displaystyle
522: \frac{g_f^2}{2} &  \displaystyle \frac{1}{2}\\ \hline
523: \rule[-4mm]{0mm}{11mm} r = 1 & -
524: \displaystyle \frac{g_f}{2} & - \displaystyle \frac{1}{2g_f}\\
525: \hline \rule[-4mm]{0mm}{11mm} r \ge 2 & 0 &  \displaystyle
526: \frac{g_f^{-r}}{2} (g_f^2 - 1)
527: \end{array}~~.
528: \end{equation}
529: 
530: For the case $g_i = +\infty$, we have
531: \begin{equation}
532: \tilde{a}(k) = \frac{2g_f z - (z^2+1)}{2(z-g_f)}
533: \end{equation}
534: and $a_r$ is given by
535: \begin{equation}
536: \begin{array}{c|c|c}
537:  & g_f < 1 & g_f > 1\\ \hline \rule[-4mm]{0mm}{11mm}
538: r \le -1 &  \displaystyle -\frac{g_f^{-r-1}}{2} (1-g_f^2) & 0\\
539: \hline \rule[-5mm]{0mm}{11mm} r = 0 &  \displaystyle \frac{g_f}{2}
540: &  \displaystyle \frac{1}{2g_f}\\ \hline \rule[-5mm]{0mm}{11mm} r
541: = 1 &  \displaystyle -\frac{1}{2} & \displaystyle \frac{1}{2g_f^2}
542: - 1\\ \hline \rule[-5mm]{0mm}{11mm} r \ge 2 & 0 & \displaystyle
543: -\frac{1}{2g_f^{r+1}}(g_f^2 - 1)
544: \end{array}~~.
545: \end{equation}
546: 
547: \newtheorem{cond}{Condition}
548: In both of these two cases, the following conditions are met:
549: \begin{cond}
550: For $g_f > 1$, $a_r = 0$ for $r \le -1$.
551: \end{cond}
552: \begin{cond}
553: For $g_f < 1$, $a_r = 0$ for $r \ge 2$.
554: \end{cond}
555: \begin{cond}
556: For $g_f < 1$, $a_r = ga_{r+1}$ for $r \le -2$.
557: \end{cond}
558: 
559: Using Condition 1, we can immediately write $G_n (\infty ) =
560: a_0^n$ for $g_f
561: > 1$. For $g_f < 1$, define \be D_n^r =
562: \begin{vmatrix}
563: a_{-r} & a_{-r-1} & \ldots & a_{-r-n+1}\\
564: a_{1} & a_0 & \ldots & a_{-n+2}\\
565: \vdots & \vdots & \ddots & \vdots\\
566: a_{n-1} & a_{n-2} & \ldots & a_0
567: \end{vmatrix},
568: \ee so that $G_n (\infty ) = D_n^0$. Condition 2 gives $D_n^r =
569: a_-r D_{n-1}^0 - a_1 D_{n-1}^{r+1}$ and Condition 3 gives $D_n^r =
570: g_f D_n^{r-1}$ for $r \ge 2$. Also, $D_1^r = a_{-r}$. We can
571: therefore write \bea
572: \begin{pmatrix}
573: D_n^0\\
574: D_n^1
575: \end{pmatrix}
576: &=&
577: \begin{pmatrix}
578: a_0 & -a_1\\
579: a_{-1} & -a_1 g_f
580: \end{pmatrix}
581: \begin{pmatrix}
582: D_{n-1}^0\\
583: D_{n-1}^1
584: \end{pmatrix}\\
585: &=& {\begin{pmatrix}
586: a_0 & -a_1\\
587: a_{-1} & -a_1 g_f
588: \end{pmatrix}}^{n-1}
589: \begin{pmatrix}
590: D_1^0\\
591: D_1^1
592: \end{pmatrix}\\
593: &=& {\begin{pmatrix}
594: a_0 & -a_1\\
595: a_{-1} & -a_1 g_f
596: \end{pmatrix}}^n
597: \begin{pmatrix}
598: 1\\
599: 0
600: \end{pmatrix}.
601: \eea This can be evaluated by diagonalizing the matrix.
602: 
603: Collecting all the analytic results above, we have for the case
604: $g_i = 0$:
605: \begin{equation}
606: G_n (\infty )= \left\{
607: \begin{array}{r} \displaystyle
608: \frac{g_f^{n+1}}{2^n} \cosh \left[ (n+1) \ln \left(
609: \frac{1+\sqrt{1-g_f^2}}{g_f} \right) \right], \\
610: \mbox{for $g_f \leq 1$} \\~\\ \displaystyle
611: \left(\frac{1}{2}\right)^n,~~~~~~~~~~~\mbox{for $g_f \geq 1$}
612: \end{array}
613: \right.
614: \end{equation}
615: In the limit that $n \rightarrow \infty$, the result for $g_f < 1$
616: becomes \be G_n (\infty ) \rightarrow {\left(
617: \frac{1+\sqrt{1-g_f^2}}{2} \right)}^{n+1}. \ee
618: 
619: In the case $g_i = +\infty$, the corresponding results are
620: \begin{equation}
621: G_n (\infty )= \left\{
622: \begin{array}{cc} \displaystyle
623: \left(\frac{1}{2}\right)^n \cos \left( n \arccos (g_f) \right), &
624: \mbox{for $g_f \leq 1$} \\~\\ \displaystyle
625: \left(\frac{1}{2g_f}\right)^n, & \mbox{for $g_f \geq 1$}
626: \end{array}
627: \right. \label{e1}
628: \end{equation}
629: Note that there are spatial oscillations in the correlator for the
630: case where the field is reduced from a large positive value ($g_i
631: = +\infty$) to a value below the critical point ($g_f<1$).
632: 
633: Of these exact results, the case $g_i = \infty$ is the one that
634: corresponds most closely to the physical situation discussed in
635: Section~\ref{oned}. Here we start from an fully `disordered'
636: initial state, and then suddenly change parameters to values with
637: increasing order (this is the analog of increasing $E$ in
638: Section~\ref{oned}). For final parameter values $g_f > g_c = 1$,
639: we find here a result quite similar to that found in
640: Section~\ref{oned}: from (\ref{e1}) we see that the order
641: parameter correlations decay with the a correlation length $\xi_f$
642: given by
643: \begin{equation}
644: \xi_f = \frac{1}{\ln(2 g_f )}.
645: \end{equation}
646: This increases monotonically with decreasing $g_f$, and is thus
647: similar to the increase in the value of $\langle O \rangle_t$ with
648: increasing $E_f$ for $E_f < E_c$ in Fig~\ref{fig3}. By the analogy
649: with Fig~\ref{fig3}, we would expect here that there is a maximum
650: in $\xi_f$ at $g=g_c$. However, we find a somewhat different
651: behavior for $g_f < g_c$ in (\ref{e1}): the correlations do not
652: decay in a simple exponential, but now oscillate, with the period
653: of oscillation becoming smaller with decreasing $g_f$. So the
654: correlations of the Ising ordered state are indeed best formed at
655: $g_f = g_c$, but we find an unusual oscillatory decay of
656: correlations for $g_f < g_c$. The oscillations are a clear
657: indication of the absence of thermalization in the present model,
658: and we expect they are special consequence of its integrability.
659: 
660: We extended these analytic results by numerical evaluation of
661: (\ref{toep}) for other values of $g_i$, and found closely related
662: behavior. Our results for $g_i = 2$ are shown in Fig~\ref{fig3b},
663: and these are the analog here of the results in Fig~\ref{fig3}
664: and~\ref{fig3a}.
665: % ************************************************************
666: \begin{figure}
667:         \centerline{\epsfig{file=gi_2.eps,width=8cm,angle=0}}
668:         \caption{Ising order correlations defined in (\protect\ref{defgn}). The
669:         system is in the ground state of $H_I$ for $t < 0$ with $g=g_i = 2$.
670:         At $t=0^+$, the value of $g$ is changed suddenly to $g=g_f$, and remains
671:         at this value for all $t > 0$. Note that at long times, the order is best
672:         developed for $g_f = 1$, which is the location of the equilibrium quantum
673:         critical point. This result is the analog of Figs~\protect\ref{fig3} and~\ref{fig3a}
674:         for the dipole model of Section~\protect\ref{oned}.}
675:         \label{fig3b}
676: \end{figure}
677: % ************************************************************
678: As $g_f$ is decreased, the correlations become longer-ranged,
679: until they reach a maximum range at $g_f = g_c = 1$. At smaller
680: values of $g_f$, the correlations acquire an oscillatory behavior,
681: but are also clearly shorter ranged. So the Ising order is best
682: developed for $g_f$ near the quantum critical point.
683: 
684: \section{Dynamics in three dimensions}
685: \label{threed}
686: 
687: We now return to the `tilted' Mott insulator problem addressed in
688: Section~\ref{oned} and in I. Here we will address questions of
689: quench dynamics for the three dimensional case. As discussed at
690: length in I, the resonant subspace in 3D is described by
691: quasiparticles and quasiholes which are free to move in the
692: directions transverse to the applied electric field. Consequently,
693: the dipoles of Section~\ref{oned}, which are bound
694: quasihole-quasiparticle pairs in adjacent sites, constitute only a
695: small part of the resonant subspace, and an effective Hamiltonian
696: for unbound quasiparticle and quasihole states is necessary. A
697: mean-field theory of this effective Hamiltonian was examined in I,
698: and a fairly complex phase diagram was found. In addition to the
699: Ising density wave order that appeared in one dimension, states
700: with a {\em transverse superfluid} order were present. The latter
701: states correspond to delocalization of the quasiholes and
702: quasiparticles in the direction transverse to the applied electric
703: field.
704: 
705: In this section, we will address the quench dynamics across the
706: transition associated with the onset of transverse superfluid
707: order. This was found to be a second-order transition in the
708: mean-field theory of I, and here we will extend the mean-field
709: theory to an analysis of the non-equilibrium dynamics across the
710: superfluid-insulator transition. We will not examine here the
711: onset of Ising order, already studied in Section~\ref{oned}; the
712: present mean-field theory found a strong first-order transition
713: for the onset of Ising order. Our analysis will be restricted to
714: the regime where both the superfluid and insulating states have no
715: Ising density wave order.
716: 
717: The effective mean-field Hamiltonian describing the dynamics of
718: these quasiparticles and quasiholes can be written as in I:
719: \begin{eqnarray}
720: \lefteqn{ H_{3D}[\langle p_\ell\rangle ,\langle h_\ell\rangle ;E]=}  \nonumber\\
721: && \sum_\ell \Bigg [ -w Z \Big(n_0 h_\ell \langle h_\ell\rangle ^*
722: + (n_0+1)p_\ell \langle p_\ell\rangle ^*
723: \nonumber\\
724: &&+ {\rm h.c} \Big) -w \sqrt{n_0(n_0+1)} \Big( p_\ell h_{\ell-1}
725: +{\rm h.c} \Big)
726: \nonumber\\
727: && + \frac{(U-E)}{2} \left(p_\ell^{\dagger} p_\ell +
728: h_\ell^{\dagger} h_\ell \right)  -\mu_\ell
729: \left(p_{\ell+1}^{\dagger} p_{\ell+1} - h_\ell^{\dagger}
730: h_\ell \right) \Bigg]. \nonumber\\
731: \label{ham3d}
732: \end{eqnarray}
733: Here $\ell$ is a one-dimensional site index labelling sites along
734: the longitudinal direction of the applied potential gradient (the
735: transverse degrees of freedom are treated in a mean-field
736: approximation and so there is no dependence on the transverse site
737: label), $p$ and $h$ are quasiparticle and quasihole annihilation
738: operators, $Z$ is the number of nearest neighbors in the
739: transverse directions and $\mu_\ell$ denotes chemical potential
740: which enforces the constraints
741: \begin{equation} \langle
742: p_{\ell+1}^{\dagger} p_{\ell+1}\rangle  = \langle h_\ell^{\dagger}
743: h_\ell\rangle. \label{const1}
744: \end{equation}
745: Although the Hamiltonian (\ref{ham3d}) has no non-linear terms,
746: its diagonalization is non-trivial because of the hard-core
747: constraint on all sites
748: \begin{equation}p_\ell^{\dagger}p_\ell
749: \le 1,~~~~h_\ell^{\dagger} h_\ell \le 1,~~~~p_\ell^{\dagger}
750: p_\ell h_\ell^{\dagger} h_\ell =0.
751: \label{const2}
752: \end{equation}
753: The mean fields $\langle p_\ell\rangle $ and $\langle
754: h_\ell\rangle $ correspond to transverse particle/hole superfluid
755: order and were self consistently determined by diagonalizing the
756: 3D Hamiltonian (\ref{ham3d}) while maintaining (\ref{const2}).
757: 
758: We now consider the evolution of the ground state under a sudden
759: shift in the value of $E$ from $E=E_i$ to $E=E_f$ at $t=0^+$. We
760: place $E_i$ in a regime where the ground state preserves all
761: symmetry, and there is neither Ising or transverse superfluid
762: order. The initial ground state $|\Psi^{\rm 3D}\rangle $ will
763: evolve according to the new Hamiltonian $H_{\rm 3D}[\langle
764: p\rangle ,\langle h\rangle ;E_f]$. However, in contrast to the 1D
765: case, here the evolutions of the mean fields $\langle p\rangle $
766: and $\langle h\rangle $ have to be self-consistently determined.
767: Within time-dependent Hartree approximation, we obtain
768: \begin{eqnarray}
769: |\Psi^{\rm 3D}(t)\rangle  &=& \sum_m c_m(t) |m\rangle  \nonumber\\
770: i \hbar \frac{d c_m (t)}{dt} &=&\sum_n c_n(t) \nonumber\\
771: && \times \langle n|H_{\rm 3D}[\langle p_\ell(t)\rangle ,\langle h_\ell(t)\rangle ;E_f]|m\rangle  \nonumber\\
772: \langle p_\ell(t)\rangle  &=& \sum_{m,n} c_m^*(t) c_n(t) \langle m|p|n\rangle  \nonumber\\
773: \langle h_\ell(t)\rangle  &=& \sum_{m,n} c_m^*(t) c_n(t) \langle
774: m|h|n\rangle .
775:  \label{scd}
776: \end{eqnarray}
777: We used a basis of states $|n\rangle $ (the final results are, of
778: course, independent of the choice of this basis) which are the
779: complete set of eigenkets of the Hamiltonian $H_{\rm 3D}[\langle
780: p_\ell^f\rangle ,\langle h_\ell^f\rangle ;E_f]$, where $\langle
781: p_\ell^f\rangle $ and $\langle h_\ell^f\rangle $ are the ground
782: states values of the particle and hole order condensates for
783: $E=E_f$. All the states $|n \rangle$ maintain (\ref{const2})
784: exactly, and so these hard-core constraints are fully respected by
785: our calculation: this is what makes diagonalization of the
786: Hamiltonian time consuming and numerically intensive. We note that
787: these equations also maintain the constraints (\ref{const1}) at
788: all $\ell$ and $t$.
789: 
790: We examined the above equations for the transverse superfluid
791: order using the same protocol used in Section~\ref{oned} for the
792: Ising order. The set of Eqs.\ \ref{scd} were solved
793: self-consistently for longitudinal system size $N=4$. We consider
794: the starting potential gradient $E_i$ to be in the insulator phase
795: with neither superfluid or Ising order, and ramp up the potential
796: gradient to enter the superfluid phase. The gauge symmetry of the
797: superfluid order parameter is broken by adding a small symmetry
798: breaking term $H_{\rm sym} = -\sum_\ell \eta [(p_\ell + h_\ell) +
799: {\rm h.c}]$, where $\eta$ is an infinitesimal positive constant.
800: In the absence of such a symmetry breaking term we would have
801: $\langle p_{\ell} \rangle$ and $\langle h_{\ell} \rangle$ vanish
802: identically for all $t$ by gauge invariance. However, even an
803: infinitesimal symmetry breaking is sufficient, in the suitable
804: parameter regime, to induce appreciable values of the superfluid
805: order. In practice, this symmetry breaking is provided by the
806: coupling of the system to its environment.
807: 
808: With the symmetry breaking term present, the superfluid order
809: parameters $\langle p_\ell \rangle$ and $\langle h_\ell \rangle$
810: are initially real. As seen in Fig.\ \ref{fig4a}, $\langle
811: h(t)\rangle = \sum_\ell \langle h_\ell(t)\rangle $ develop
812: coherent oscillations once the potential gradient takes the value
813: $E_f$. (These results are the analog of
814: Figs~\ref{fig1},\ref{fig2}.)
815: % ************************************************************
816: \begin{figure}
817:         \centerline{\epsfig{file=oscdata1.eps,width=8cm,angle=0}}
818:         \caption{Oscillations of the hole superfluid order parameter.
819:         The plot is shown for $U=40$, $w=1$, $n_0 = 1$, $Z=4$, $E_i=20$ and $E_f=30$. For
820:         these parameters, the quantum critical point is at $E_c = 26.4$.}
821:         \label{fig4a}
822: \end{figure}
823: % ************************************************************
824: The oscillations in $\langle p(t)\rangle $ are similar, but occur
825: with a different period due to the inherent particle-hole
826: asymmetry in Eq.\ \ref{ham3d}. The long time average of the
827: oscillations is purely real, and this is, of course determined by
828: the symmetry breaking term. Such oscillations of the superfluid
829: order parameter were also obtained recently in a different context
830: in Refs.~\onlinecite{aa,psg} and by Levitov\ \cite{Levitov}.
831: 
832: We also examined the $E_f$ dependence of the long time average of
833: the superfluid order, and the analog of Fig~\ref{fig3} appears in
834: Fig~\ref{fig4b}.
835: % ************************************************************
836: \begin{figure}
837:         \centerline{\epsfig{file=longtime3d.eps,width=8cm,angle=0}}
838:         \caption{Analog of Fig~\protect\ref{fig3}, but for the
839:         transverse superfluid order, using the same parameters
840:         (apart from $E_f$) as Fig~\protect\ref{fig3a}. The
841:         "static" curve is the equilibrium superfluid order
842:         parameter determined in I. The "dynamic" curve is the
843:         long time average of the real part of $\langle h(t)
844:         \rangle$.}
845:         \label{fig4b}
846: \end{figure}
847: % ************************************************************
848: Again the superfluid order is most strongly enhanced in the
849: vicinity of the quantum critical point. However, unlike
850: Fig~\ref{fig3}, we do not observe a precursor to the superfluid
851: order in the insulating phase: this is surely an artifact of the
852: mean-field treatment of the transverse degrees of freedom.
853: 
854: \section{Conclusions}
855: \label{conc}
856: 
857: With advent of the study of quantum phase transitions in trapped
858: atomic systems, there is a clear need for theoretical studies in
859: the highly non-equilibrium situations that experiments are often
860: in. In particular, experiments can easily explore the change in
861: the state of the system upon a sudden change in a parameter in the
862: Hamiltonian. There are few general principles in such cases ({\em
863: e.g.\/} there is no fluctuation-dissipation theorem which controls
864: correlations of the final state), and theory is clearly still in
865: its infancy. Two recent studies in this class \cite{aa,psg},
866: examined the evolution of superfluid order under a sudden change
867: in the optical lattice potential exerted on trapped bosons.
868: 
869: It is clear that exact results on simple solvable models in
870: non-equilibrium situations would be valuable. We have provided
871: such an example here in Section~\ref{sec:ising}, where we examined
872: the Ising chain in a transverse field, $g$. This model has a
873: quantum critical point at $g=g_c$, with spontaneous ferromagnetic
874: order in the ground state for $g<g_c$. We started the Ising model
875: in the paramagnetic state ($g_i \gg g_c$), suddenly at $t=0$
876: changed $g$ to a final value $g_f$, and examined the long time
877: development of correlations of the ferromagnetic order. (Our
878: formalism also provided results for all $t>0$, but we have not
879: examined the detailed time evolution here.) The results are
880: summarized in Fig~\ref{fig3b}. True long-range order does not
881: develop at any value of $g_f$; however, significant order
882: parameter correlations do appear, and these are best formed for
883: $g_f \approx g_c$. In a general non-integrable system we may
884: expect thermalization at long times, at a temperature such that
885: the average energy equals that of the state at $t=0^+$. Such
886: thermalization does not occur for the present integrable system,
887: and the results have certain artifacts associated with this: the
888: long-time correlations have an oscillatory spatial dependence for
889: $g_f < g_c$.
890: 
891: In the remainder of the paper we studied the non-equilibrium
892: dynamics of models introduced in a previous paper \cite{Sachdev2}
893: which addressed the response of a bosonic Mott insulator to a
894: change in a strong potential gradient \cite{Bloch}. These models
895: exhibit a number of quantum critical points associated with the
896: onset of Ising density wave and superfluid order. Our numerical
897: studies here found a feature similar to that also obtained for the
898: solvable Ising model: the order was best formed when the final
899: parameter value was in the vicinity of the associated quantum
900: critical point, as illustrated in Fig~\ref{fig3}. Here, and in
901: Ref.~\onlinecite{Sachdev2}, we have proposed this feature as the
902: explanation for the resonant response observed by Greiner {\em et
903: al.} \cite{Bloch} upon `tilting' a Mott insulator of bosons in an
904: optical lattice.
905: 
906: \begin{acknowledgments}
907: We thank A. G. Abanov, S.~M.~Girvin, L.~Levitov, and
908: A.~Polkovnikov for useful discussions. This research was supported
909: by US NSF grants DMR-0098226 and DMR-0342157.
910: \end{acknowledgments}
911: 
912: \appendix*
913: 
914: \section{Computations for the Ising chain}
915: 
916: The Jordan-Wigner transformation allows the Hamiltonian of an
917: Ising chain in a transverse field $g$ to be written as \be
918: \label{Hamiltonian} \Ham_I = \sum_k \epsilon_k \gamma_k^\dag
919: \gamma_k, \ee where $\gamma_k$ is a fermionic annihilation
920: operator (see Chapter 4 of Ref.~\onlinecite{Sachdev1}). These are
921: related to the Jordan-Wigner fermions $c_k$ by a Bogoliubov
922: transformation, parametrized by angle $\theta_k$, where
923: \be\label{DefineThetaK} \tan\theta_k = \frac{\sin k}{g - \cos k}.
924: \ee In the present case, we define the $\gamma$ fermions as those
925: that diagonalize the Hamiltonian for $t > 0$, with field $g_f$.
926: 
927: Since the Hamiltonian is throughout translationally symmetric,
928: only fermionic states with opposite pseudomomentum $k$ and $-k$
929: are mixed. We may therefore write the two component column vector
930: \be\label{DefineGammaK}\Gamma_k =
931: \begin{pmatrix}
932: \gamma_k\\
933: \gamma_{-k}^\dag
934: \end{pmatrix}
935: \ee and similarly $c_k$ for the Jordan-Wigner fermions. The
936: Bogoliubov transformation relating $c_k$ and $\Gamma_k$ is
937: expressed as $c_k = R^x(\theta_k) \Gamma_k$, where \be R^x(\alpha)
938: = \cos\frac{\alpha}{2} + i\sigma^x\sin\frac{\alpha}{2} \ee and
939: here $\sigma^x$ is a $2\times 2$ Pauli matrix. (These are used for
940: conciseness of notation and should not be confused with the
941: operators representing the `spins' of the Ising chain.)
942: 
943: For $t < 0$, the field is $g_i$ and the system is taken to be in
944: its ground state. We define the $\gamma'$ fermions as those which
945: diagonalize the Hamiltonian in the form (\ref{Hamiltonian}) with
946: this field. (Similarly, $\theta_k'$ and $\Gamma_k'$ are given by
947: analogy with (\ref{DefineThetaK}) and (\ref{DefineGammaK}).)
948: 
949: The state $|\psi\rangle$ is therefore the vacuum of $\gamma'$
950: fermions: in matrix notation, \be \langle\psi |
951: \Gamma_k'\Gamma_k'^\dag|\psi\rangle =
952: \begin{pmatrix}
953: 1 & 0\\
954: 0 & 0
955: \end{pmatrix}
956: = \frac{1}{2}(\sigma^z + 1). \ee Applying the Bogoliubov
957: transformation to $c_k$ and then $\Gamma_k$ gives \be \langle\psi
958: | \Gamma_k\Gamma_k^\dag|\psi\rangle = R^{x\dag}(\theta_k -
959: \theta_k') \frac{1}{2}(\sigma^z + 1) R^x(\theta_k - \theta_k').
960: \ee Using $R^{x\dag}(\alpha)\sigma^z R^x(\alpha) = \sigma^z \cos
961: \alpha - \sigma^y \sin \alpha$ with $\phi_k = \theta_k -
962: \theta_k'$ gives \be \langle\psi |
963: \Gamma_k\Gamma_k^\dag|\psi\rangle = \frac{1}{2}(1 + \sigma^z \cos
964: \phi_k - \sigma^y \sin \phi_k), \ee as the set of matrix elements
965: for the initial state.
966: 
967: The time evolution of the operators now proceeds (using the
968: Heisenberg picture) according to the Hamiltonian,
969: (\ref{Hamiltonian}), so that $\Gamma_k(t) = U_k(t) \Gamma_k(0)$,
970: where \be U_k(t) =
971: \begin{pmatrix}
972: e^{-i \epsilon_k t} & 0\\
973: 0 & e^{i \epsilon_k t}
974: \end{pmatrix}
975: = R^{z\dag}(2\epsilon_k t). \ee The expectation values at any time
976: can therefore be evaluated using the algebra of $SU(2)$ matrices.
977: 
978: The $n$-site correlator can be written as $\langle G_n \rangle$ =
979: $\langle B_0 A_1 B_1 \ldots B_{n-1} A_n \rangle$, where $A_i =
980: c_i^\dag + c_i$ and $B_i = c_i^\dag - c_i$. Wick's theorem can
981: then be used to write this expression in terms of the expectation
982: values of expressions bilinear in $A_i$ and $B_i$. We therefore
983: let \be \Omega_k(t) =
984: \begin{pmatrix}
985: A_k(t)\\
986: B_k(t)
987: \end{pmatrix}
988: = \sqrt{2} R^{y}\left(\frac{\pi}{2}\right) C_k(t), \ee where $A_k$
989: ($B_k$) is the Fourier transform of $A_i$ ($B_i$) so that
990: \begin{widetext}
991: \bea \langle\psi | \Omega_k \Omega_k^\dag|\psi\rangle &=&
992: \langle\psi |
993: \begin{pmatrix}
994: A_k A_{-k} & -A_k B_{-k} \\
995: B_k A_{-k} & -B_k B_{-k}
996: \end{pmatrix}
997: |\psi\rangle = 2 R^{y}\left(\frac{\pi}{2}\right)\langle\psi | C_k C_k^\dag|\psi\rangle R^{y\dag}\left(\frac{\pi}{2}\right)\\
998: &=& 1 - \sigma^x (\cos \theta_k \cos \phi_k - \sin\phi_k\sin\theta_k\cos 2\epsilon_k t)\nonumber\\
999: & &+ \sigma^y (\sin \theta_k \cos \phi_k +
1000: \sin\phi_k\cos\theta_k\cos 2\epsilon_k t)
1001: + \sigma^z (\sin\phi_k \sin 2\epsilon_k t)\\
1002: &=& \begin{pmatrix}
1003: 1 - \sin\phi_k\sin 2\epsilon_k t & -e^{i \theta_k} (\cos \phi_k + i \sin\phi_k\cos 2\epsilon_k t)\\
1004: -e^{-i \theta_k} (\cos \phi_k - i \sin\phi_k \cos 2\epsilon_k t) &
1005: 1 + \sin\phi_k\sin 2\epsilon_k t
1006: \end{pmatrix}.
1007: \eea
1008: \end{widetext}
1009: 
1010: Transforming to real space and using the conservation of
1011: pseudomomentum gives \bea
1012: \langle A_lA_j\rangle &=& \frac{1}{M}\sum_k e^{ik(l-j)}(1-\sin\phi_k\sin 2\epsilon_k t)\nonumber\\
1013: \langle B_lB_j\rangle &=& \frac{1}{M}\sum_k e^{ik(l-j)}(-1-\sin\phi_k\sin 2\epsilon_k t)\nonumber\\
1014: \langle B_lA_j\rangle &=& \frac{1}{M}\sum_k e^{ik(l-j)}e^{-i\theta_k}(-\cos\phi_k\nonumber\\
1015: & &+ i\sin\phi_k\cos 2\epsilon_k t). \eea The long time averages
1016: of these expressions are \bea
1017: \langle A_l A_j \rangle &=& \delta_{lj}\nonumber\\
1018: \langle B_l B_j \rangle &=& -\delta_{lj}\nonumber\\
1019: \langle B_l A_j \rangle &=& a_{l-j+1}, \eea where \bea
1020: a_r &=& \frac{1}{M}\sum e^{-ikr} \tilde{a}(k)\\
1021: &=& \frac{1}{2\pi}\int_{-\pi}^{\pi} e^{-ikr} \tilde{a}(k), \eea in
1022: the limit where the number of sites $M$ becomes infinite. Here,
1023: $\tilde{a}(k) = -e^{i(\theta_k + k)}\cos (\theta_k - \theta_k')$.
1024: Wick's theorem then allows the expression for $\langle G_n
1025: \rangle$ to be written as (\ref{toep}).
1026: 
1027: 
1028: \begin{thebibliography}{99}
1029: 
1030: \vspace{-0.7cm}
1031: 
1032: \bibitem{Bloch} M.~Greiner, O.~Mandel, T.~Esslinger, T.~W.~H\"ansch, and I.~Bloch,
1033: Nature (London) {\bf 415}, 39 (2002).
1034: 
1035: \bibitem{Kasevich} C.~Orzel, A.~K.~Tuchman, M.~L.~Fenselau, M.~Yasuda, and
1036: M.~A.~Kasevich, Science {\bf 291}, 2386 (2001).
1037: 
1038: \bibitem{fwgf} M.~P.~A.~Fisher, P.~B.~Weichman, G.~Grinstein, D.~S.~Fisher,
1039: Phys. Rev. B {\bf 40}, 546 (1989).
1040: 
1041: \bibitem{Sachdev1} For a review, see Chapter 11 in {\it Quantum Phase
1042: Transitions}, S. Sachdev,(Cambridge University Press, Cambridge,
1043: England, 1999).
1044: 
1045: \bibitem{Sachdev2} S.~Sachdev, K.~Sengupta and S.~M.~Girvin, Phys.
1046: Rev. B, {\bf 66} 075128 (2002).
1047: 
1048: \bibitem{burnett} K.~Braun-Munzinger, J.~A.~Dunningham, and K. Burnett,
1049: cond-mat/0211701.
1050: 
1051: \bibitem{fendley} P.~Fendley, K.~Sengupta, and S.~Sachdev,
1052: cond-mat/0309438.
1053: 
1054: \bibitem{fh} M.~E.~Fisher and R.~E.~Hartwig, Adv. Chem. Phys. {\bf 15}, 333
1055: (1968); E.~L.~Basor and C.~A.~Tracy, J. Phys. A {\bf 177}, 167
1056: (1991); T. Ehrhardt, Operator Th: Advances and App. {\bf 124}, 217
1057: (2001).
1058: 
1059: \bibitem{aa} E. Altman and A. Auerbach,
1060: Phys. Rev. Lett. {\bf 89}, 250404 (2002).
1061: 
1062: \bibitem{psg} A. Polkovnikov, S. Sachdev, and S. M. Girvin,
1063: Phys. Rev. A {\bf 66}, 053607 (2002)
1064: 
1065: \bibitem{Levitov} L.~S.~Levitov, private communication. Note that our Hamiltonian
1066: (\protect\ref{ham3d}) has a certain resemblance to the BCS theory,
1067: but we are considering bosons subject to the hard-core constraints
1068: (\ref{const2}), rather than fermions.
1069: 
1070: 
1071: \end{thebibliography}
1072: 
1073: \end{document}
1074: