1: \documentclass[aps,epsfig,showpacs,twocolumn]{revtex4}
2:
3: %\documentclass[aps,epsfig,showpacs,doublespace]{revtex4}
4: %\renewcommand{\baselinestretch}{3.0}
5:
6: \usepackage{graphicx}
7:
8: \begin{document}
9: \title{Adiabatic-Impulse approximation for avoided level crossings:
10: from phase transition dynamics to Landau-Zener evolutions and back again
11: }
12: \author{Bogdan Damski and Wojciech H. Zurek}
13: \affiliation{
14: Theory Division, Los Alamos National Laboratory, MS-B213, Los Alamos, NM 87545, USA
15: }
16: \begin{abstract}
17: We show that a simple approximation based on concepts underlying the
18: Kibble-Zurek theory of second order phase transition dynamics can be used to
19: treat avoided level crossing problems. The approach discussed in this paper
20: provides an intuitive insight into quantum dynamics of two level
21: systems, and may serve as a link between the theory of dynamics of
22: classical and quantum phase transitions. To illustrate these ideas we
23: analyze dynamics of a paramagnet-ferromagnet quantum phase transition in the Ising model.
24: We also present exact unpublished solutions of the Landau-Zener like problems.
25: \end{abstract}
26: \pacs{03.65.-w,32.80.Bx,05.70.Fh}
27: \maketitle
28:
29: \section{Introduction}
30:
31: Two level quantum systems that undergo avoided level crossing
32: play an important role in physics. Often they provide not
33: only a qualitative description of system properties but also
34: a quantitative one. The possibility for a successful two level approximation arises
35: in different physical systems, e.g.,
36: a single one-half spin; an atom placed in a resonant light;
37: smallest quantum magnets; etc.
38:
39: In this paper we focus on avoided level crossing dynamics in two level quantum
40: systems. Two interesting related extensions of above mentioned long list
41: are the critical dynamics of quantum phase transitions
42: \cite{subir,dorner,jacek,polkovnikov,ralf}
43: and adiabatic quantum computing \cite{ad_q_com}. In both of these cases (and,
44: indeed, in all of the other applications of the avoided level
45: crossing scenario) there are many more than two levels, but the
46: essence of the problem can be still captured by the two level, Landau-Zener type,
47: calculation.
48:
49: Our interest in Landau-Zener model originates also from its well known numerous
50: applications to different physical systems.
51: In many cases there is a possibility of more general dependence
52: of the relevant parameters (i.e. gap between the two levels on time) than in
53: the original Landau-Zener treatment. This motivates our extensions in this
54: paper of the level crossing dynamics to asymmetric level crossings and various power-law
55: dependences. Appropriate variations of external parameters
56: driving Landau-Zener transition can allow for an experimental realization of
57: these generalized Landau-Zener like models in quantum magnets \cite{magnet}
58: and optical lattices \cite{bloch}.
59:
60:
61: We focus on evolutions that include both adiabatic and diabatic (impulse) regime
62: during a single sweep of a system parameter.
63: For simplicity, we call these evolutions diabatic
64: since we assume that they include a period of
65: fast change of a system parameter. This is in contrast to adiabatic time
66: evolutions induced by very slow parameter changes when the system never leaves
67: adiabatic regime.
68: We will use a
69: formalism proposed recently by one of us \cite{bodzio}. It
70: originates from the so-called Kibble-Zurek (KZ) theory of topological defect
71: production in the course of classical phase transitions \cite{kibble,zurek},
72: and works best for two level systems.
73: The two level approximation
74: was recently shown by Zurek {\it et al} \cite{dorner} and
75: Dziarmaga \cite{jacek} to be useful in studies of quantum phase transitions
76: confirming earlier expectations -- see Chap. 1.1 of \cite{subir}
77: and \cite{bodzio}.
78: Therefore, we expect that the
79: formalism presented here will provide a link between dynamical
80: studies of classical and quantum phase transitions.
81:
82: The present
83: contribution extends the ideas presented in \cite{bodzio}. In particular,
84: we have succeeded in replacing the fit to numerics used
85: there for getting a free parameter of the theory,
86: with a simple analytic calculation.
87: We also show that the method of
88: \cite{bodzio} can be successfully applied to a large class of two
89: level systems. Moreover, we present nontrivial
90: exact analytic solutions for Landau-Zener model and the relevant
91: adiabatic-impulse approximations not published to date.
92:
93: Section \ref{aia} presents basics
94: of what we call the adiabatic-impulse approximation. Section \ref{LZ} shows
95: adiabatic-impulse, diabatic and finally exact
96: solutions to different versions of the Landau-Zener problem.
97: In Section \ref{2n} we discuss solutions for a whole class of two level systems
98: on the basis of adiabatic-impulse and diabatic schemes.
99: Section \ref{ising} presents how results of Section \ref{LZ} can be used to study
100: quantum Ising model dynamics.
101: Details of
102: analytic calculations are in Appendix \ref{a_l} (diabatic
103: solutions), Appendix \ref{a_lz} (exact solutions of Landau-Zener problems)
104: and Appendix \ref{a_r} (adiabatic-impulse solution from Sec. \ref{2n}).
105:
106: \section{Adiabatic-Impulse approximation}
107: \label{aia}
108: We will study dynamics of quantum systems by
109: assuming that it includes
110: {\it adiabatic} (no population transfer between instantaneous energy eigenstates),
111: and {\it impulse} (no changes in the wave function except for an overall phase factor) stages
112: only.
113: The nature of this approach suggests the name {\it
114: adiabatic-impulse} (AI) approximation.
115: The AI approach
116: originates from the Kibble-Zurek (KZ) theory of nonequilibrium classical phase transitions
117: and we refer the reader to \cite{bodzio} for a detailed discussion of AI-KZ connections.
118: Below, we summarize basic ideas of the KZ theory, and the
119: relevant assumptions used later on.
120:
121: Second order phase transitions and avoided level crossings share one key
122: distinguishing characteristic: in both cases sufficiently near the critical point
123: (defined by the
124: relaxation time or by the inverse of the size of the gap between the two
125: levels) system ``reflexes'' become very bad.
126: In phase transitions this is known as ``critical slowing down''.
127:
128: In the treatment of the second order phase transitions this leads to the
129: behavior where the state of the system can initially -- in the adiabatic region
130: far away from the transition -- adjust to the change of the relevant parameter
131: that induces the transition, but sufficiently close to the critical point
132: its reflexes become too slow for it to react at all. As a consequence, a
133: sequence of three regimes (adiabatic, impulse near the critical point, and
134: adiabatic again on the other side of the transition) can be anticipated.
135: In the second order phase transitions this allows one to employ a
136: configuration dominated by the pre-transition fluctuations to calculate
137: salient features of the post-transition state of the order parameter
138: (e.g., the density of topological defects). Key predictions of this KZ
139: mechanism have been by now verified in numerical simulations \cite{antunes,laguna} and
140: more importantly in the laboratory experiments \cite{kz1,kz2,kz3,kz4}.
141:
142: The crux of this story is of course the moments when the transition
143: from adiabatic to impulse and back to adiabatic behavior takes place.
144: Assuming that the transition point is crossed at time $t=0$,
145: this must happen around the instants $\pm\hat t$, where
146: $\hat t$ equals system relaxation time $\tau$ \cite{zurek}.
147: As the relaxation time depends on system parameter $\varepsilon$, which
148: is changed as a function of time, one can write
149: \begin{equation}
150: \label{whz}
151: \tau(\varepsilon(\hat t\,))=\hat t.
152: \end{equation}
153: This basic equation proposed in \cite{zurek} can be solved when dependence of $\tau$ on some measure of
154: the distance from the critical point (e.g., relative temperature or coupling
155: $\varepsilon$) is known. Assuming, e.g.,
156: $\tau=\tau_0/\varepsilon$ and $\varepsilon= t/\tau_Q$,
157: where $\tau_Q$, the ``quench time", contains information about how fast the
158: system is driven through the transition, one arrives at
159: $\tau_0/(\hat{t}/\tau_Q)= \hat{t}$. Hence,
160: the time $\hat{t}$ at which the behavior switches between approximately
161: adiabatic and approximately impulse is $\sqrt{\tau_0\tau_Q}$.
162: Once this last ``adiabatic" instant in the evolution of the system is
163: known, interesting features of the post transition state (such as the size of
164: the regions $\hat{\xi}$ in which the order parameter is smooth) can be
165: computed. Generalization to other universality classes is straightforward
166: \cite{zurek} and leads to
167: \begin{equation}
168: \hat{t}\sim \tau_Q^{1/(1+\nu z)} \ \ \ , \ \ \
169: \hat{\xi}\sim \tau_Q^{\nu/(1+\nu z)},
170: \end{equation}
171: where $z$ and $\nu$ are universal critical exponents (see, e.g., \cite{antunes}).
172:
173: In what follows we adopt this approach to quantum systems where
174: dynamics can be approximated as an avoided level crossing (see Fig.
175: \ref{gapanty}).
176: For simplicity we
177: restrict ourselves to two-level systems that possess a single
178: anti-crossing (Fig. \ref{gapanty}) in the
179: excitation spectrum and a gap preferably going to $+\infty$
180: far away from the anti-crossing.
181: The latter condition guarantees that
182: asymptotically the system enters adiabatic regime.
183:
184:
185: The passage through avoided level crossing
186: can be divided into an adiabatic and an impulse regime according to the
187: size of the gap in comparison with the energy scale that characterizes the rate of the
188: imposed changes in system Hamiltonian. If the gap is large enough
189: the system is in adiabatic regime, while
190: when the gap is small it undergoes impulse time evolution as depicted in Fig.
191: \ref{gapanty}(b). The instants $\pm\hat{t}$ are supposed to be such that
192: the discrepancy between time dependent exact results and
193: those coming from splitting of the evolution into only adiabatic and impulse parts
194: is minimized. Therefore, the AI approximation looks like a
195: {\it time dependent variational method} where $\hat{t}$ is a variational parameter
196: while adiabatic-impulse assumptions provide a form of a variational wave
197: function.
198:
199:
200: To make sure that assumptions behind the AI approach are
201: correctly understood, let's consider time evolution of the system depicted
202: in Fig. \ref{gapanty}. Let the
203: evolution start at $t_i\to-\infty$ from a ground state (GS) and
204: last till $t_f\to+\infty$. The AI method assumes that the system wave
205: function, $|\Psi(t)\rangle$, satisfies the following three approximations
206: coming from passing through first adiabatic, then impulse and finally
207: adiabatic regions
208: \begin{eqnarray}
209: t\in[-\infty,-\hat{t}\,]&:&
210: |\Psi(t)\rangle\approx({\rm phase \ factor})|{\rm GS \ at \ t}\rangle \nonumber\\
211: t\in[-\hat{t},\hat{t}\,]&:& |\Psi(t)\rangle\approx
212: ({\rm phase \ factor})
213: |{\rm GS \ at \ -\hat{t}}\rangle\nonumber\\
214: t\in[\hat{t},+\infty]&:&
215: |\langle\Psi(t)|{\rm GS \ at \ t}\rangle|^2={\rm const}. \nonumber
216: \end{eqnarray}
217:
218: \begin{figure}
219: \includegraphics[width=\columnwidth, clip=true]{fig1.eps}
220: \caption{Plot (a): structure of energy spectrum (parametrized by time) of a generic
221: two-level system
222: under consideration. Note the anticrossing at
223: $t=0$ and asymptotic form of eigenstates: $|1\rangle$, $|2\rangle$.
224: Plot (b): adiabatic-impulse regimes in system dynamics. Compare it
225: to a plot of relaxation time scale in the Kibble-Zurek theory \cite{zurek,bodzio}. Note that
226: instants separating adiabatic and impulse regimes does not have to be
227: placed symmetrically with respect to $t=0$ -- see Sec. \ref{LZa} and Fig. \ref{asym}.
228: }
229: \label{gapanty}
230: \end{figure}
231:
232: Finally, one needs to know how to get the instant $\hat{t}$. As proposed in
233: \cite{bodzio}, the proper generalization of Eq. (\ref{whz})
234: to the quantum case (after rescaling everything to dimensionless
235: quantities) reads
236: \begin{equation}
237: \frac{1}{{\rm gap}(\hat{t}\,)}=\alpha\hat{t},
238: \label{bodzik}
239: \end{equation}
240: where $\alpha={\cal O}(1)$ is a constant. Similarity between Eq. (\ref{bodzik})
241: and Eq. (\ref{whz}) suggests (in accord with physical intuition) that the
242: quantum mechanical equivalent of the relaxation time scale is an inverse of the
243: gap.
244:
245: In this paper we
246: give a simple and systematic way for (i) obtaining
247: $\alpha$ analytically; (ii) verification that Eq. (\ref{bodzik}) leads to correct
248: results in the lowest nontrivial order.
249: This method, illustrated on specific examples in
250: Appendix \ref{a_l}, is based on the observation that a time dependent
251: Schr\"odinger equation can be solved exactly in the diabatic limit
252: if one looks at the lowest nontrivial terms in expressions for
253: excitation amplitudes. This should be true
254: even when getting an exact solution turns out to be very
255: complicated or even impossible. After obtaining $\alpha$ this way, the whole AI
256: approximation is complete in a sense that there are no free parameters,
257: so its predictions can be rigorously checked by comparison to exact (either
258: analytic or numeric) solution.
259:
260:
261:
262: \section{Dynamics in the Landau-Zener model}
263: \label{LZ}
264: In this section
265: we illustrate the AI approximation by considering the Landau-Zener model. In this way
266: we supplement and extend the results of \cite{bodzio}.
267:
268: The Landau-Zener model, after rescaling all the quantities to dimensionless
269: variables, is defined by the Hamiltonian:
270: \begin{equation}
271: \label{H}
272: H=\frac{1}{2}
273: \left(
274: \begin{array}{cc}
275: \frac{t}{\tau_Q} & 1 \\
276: 1 & -\frac{t}{\tau_Q}
277: \end{array}
278: \right),
279: \end{equation}
280: where $\tau_Q$ is time independent and provides a time scale on which the
281: system stays in the neighborhood of an anti-crossing.
282: As $\tau_Q\to0$ the system undergoes diabatic time evolution, while
283: $\tau_Q\gg1$ means adiabatic evolution.
284: There are two
285: eigenstates of this model for any fixed time $t$:
286: the ground state $|\downarrow(t)\rangle$ and the excited state
287: $|\uparrow(t)\rangle$ given by:
288: \begin{equation}
289: \label{eig}
290: \left[
291: \begin{array}{c}
292: |\uparrow(t)\rangle \\ |\downarrow(t)\rangle
293: \end{array}
294: \right] =
295: \left(
296: \begin{array}{rr}
297: \cos(\theta(t)/2) & \sin(\theta(t)/2) \\
298: -\sin(\theta(t)/2) & \cos(\theta(t)/2)
299: \end{array}
300: \right)
301: \left[
302: \begin{array}{c}
303: |1\rangle \\ |2\rangle
304: \end{array}
305: \right],
306: \end{equation}
307: where $|1\rangle$ and $|2\rangle$ are time-independent basis states
308: of the Hamiltonian (\ref{H}); $\cos(\theta)=\varepsilon/\sqrt{1+\varepsilon^2}$;
309: $\sin(\theta)=1/\sqrt{1+\varepsilon^2}$;
310: \begin{equation}
311: \label{varepsilon}
312: \varepsilon= \frac{t}{\tau_Q},
313: \end{equation}
314: and $\theta\in[0,\pi]$. The gap in this model is
315: $$
316: {\rm gap}= \sqrt{1+\varepsilon^2}.
317: $$
318: The instant $\hat{t}$ is obtained from Equation (\ref{bodzik}):
319: $$
320: \frac{1}{\sqrt{1+\left(\frac{\hat{t}}{\tau_Q}\right)^2}}=\alpha\hat{t},
321: $$
322: which leads to the following solution
323: \begin{equation}
324: \label{hatt}
325: \hat{\varepsilon}=\frac{\hat{t}}{\tau_Q}=
326: \frac{1}{\sqrt{2}}\sqrt{\sqrt{1+\frac{4}{(\alpha\tau_Q)^2}}-1}.
327: \end{equation}
328: Dynamics of the system is governed by the Schr\"odinger equation:
329: $$i \frac{d}{dt}|\Psi\rangle=H|\Psi\rangle,$$
330: and it will be assumed that evolution happens in the interval $[t_i,t_f]$.
331:
332: As was discussed in \cite{bodzio}, by using AI approximation one can easily arrive
333: at the following predictions for the probability of finding the system
334: in the excited state at $t_f\gg\hat{t}$
335: \begin{itemize}
336: \item time evolution starts at $t_i\ll-\hat{t}$ from the ground state
337: \begin{equation}
338: \label{infty}
339: P_{AI}=|\langle\uparrow(\hat{t}\,)|\downarrow(-\hat{t}\,)\rangle|^2=
340: \frac{\hat{\varepsilon}^2}{1+\hat{\varepsilon}^2}.
341: \end{equation}
342: In this case the system undergoes in the AI scheme the adiabatic time
343: evolution from $t_i$ to $-\hat{t}$, then impulse one from $-\hat{t}$
344: to $\hat{t}$ and finally adiabatic from $\hat{t}$ to $t_f\gg\hat{t}$.
345: \item time evolution starts at $t_i=0$ from the ground state
346: \begin{equation}
347: \label{srodek}
348: p_{AI}=|\langle\uparrow(\hat{t}\,)|\downarrow(0)\rangle|^2=
349: \frac{1}{2}\left(1-\frac{1}{\sqrt{1+\hat{\varepsilon}^2}}\right).
350: \end{equation}
351: Now evolution is first impulse from $t=0$ to $t=\hat{t}$
352: and then adiabatic from $\hat{t}$ to $t_f\gg\hat{t}$.
353: \end{itemize}
354: In the following we consistently denote excitation probability when the
355: system evolves from $t_i\ll-\hat{t}$, $t_i=0$ by $P$, $p$ respectively.
356: Additionally the subscript AI will be attached to predictions based on
357: the AI approximation.
358:
359: In the first case, the substitution
360: of (\ref{hatt}) into (\ref{infty}) leads to
361: \begin{equation}
362: P_{AI}=1- \alpha\tau_Q+\frac{(\alpha\tau_Q)^2}{2}-\frac{(\alpha\tau_Q)^3}{8}+
363: {\cal O}(\tau_Q^4).
364: \label{dupa1}
365: \end{equation}
366: Now we have to determine $\alpha$. It turns out that it can be
367: done by looking at the diabatic excitation probability. The proper
368: expression and calculation can be found in Appendix \ref{a_l}. Substituting
369: $\eta=1/2$ into (\ref{alfa_1}) and (\ref{infty_1})
370: one gets that to the lowest nontrivial order $P_{AI}=1-\pi\tau_Q/2$,
371: which implies that $\alpha=\pi/2$ and verifies Eq.
372: (\ref{bodzik})
373: for this case. Note that the calculation leading to Eqs.
374: (\ref{infty_1}) and (\ref{alfa_1}) is not only much easier then determination
375: of exact LZ solution \cite{zener}, but also really elementary. Therefore,
376: we expect that it can be done comparably easily for any model of
377: interest.
378:
379:
380: Now we are ready to compare our AI approximation with $\alpha$ determined as
381: above, to the exact result, i.e.,
382: \begin{equation}
383: P=\exp\left(-\frac{\pi\tau_Q}{2}\right).
384: \label{dupa987}
385: \end{equation}
386: First, the
387: agreement between the exact and AI result is up to ${\cal O}(\tau_Q^3)$,
388: i.e., one order above the first nontrivial term. This is the advantage that the
389: AI approximation provides over a simple diabatic approximation performed in
390: Appendix \ref{a_l}.
391: Second, we see that the AI expansion contains the same powers of $\tau_Q$
392: as the diabatic (small $\tau_Q$) expansion of the exact result. Third,
393: Fig. \ref{dupa123} quantifies the discrepancies between exact, AI and diabtic
394: results. For the AI prediction, we plot in Fig. \ref{dupa123}
395: instead of a Taylor series (\ref{dupa1}) the full expression evaluated in
396: \cite{bodzio}
397: \begin{equation}
398: P_{AI}= \frac{2}{(\alpha\tau_Q)^2+\alpha\tau_Q\sqrt{(\alpha\tau_Q)^2+4}+2},
399: \label{infty_ful}
400: \end{equation}
401: with $\alpha=\pi/2$. As easily seen the AI approximation significantly
402: outperforms a diabatic solution. In other words, the combination of
403: AI simplification of dynamics and diabatic prediction for the purpose of
404: getting the constant $\alpha$ leads to fully satisfactory results considering
405: simplicity of the whole approach.
406:
407: \begin{figure}
408: \includegraphics[width=\columnwidth, clip=true]{fig2.eps}
409: \caption{Transition probability when the system starts time evolution from a
410: ground state
411: at $t_i\to-\infty$ and evolves to $t_f\to+\infty$.
412: Dots: exact expression (\ref{dupa987}). Solid line: AI prediction
413: (\ref{infty_ful}) with $\alpha=\pi/2$ determined from diabatic solution
414: in Appendix \ref{a_l}. Dashed thick line: lowest order diabatic result,
415: $1-\pi\tau_Q/2$, coming from (\ref{infty_1}) and (\ref{alfa_1}) with $\eta=1/2$.
416: }
417: \label{dupa123}
418: \end{figure}
419:
420: It is instructive to consider now separately three situations: (i) dynamics
421: in a nonsymmetric avoided level crossing (Sec. \ref{LZa}); (ii) dynamics
422: beginning at the anti-crossing center (Sec. \ref{LZb});
423: and (iii) dynamics starting at $t_i\to-\infty$ but ending at the anti-crossing center
424: (Sec. \ref{LZc}).
425: The first case will give us a hint whether ${\cal O}(\tau_Q^3)$
426: agreement we have seen above is accidental and has something to do with the symmetry of the
427: Landau-Zener problem. The second problem
428: was preliminarily considered in \cite{bodzio}, but without
429: comparing the AI prediction to exact analytic one being interesting on its own.
430: Finally, the third problem
431: is an example where AI approximation correctly suggests at
432: a first sight unexpected symmetry between this problem and the one considered in Sec.
433: \ref{LZb}.
434:
435: \subsection{Nonsymmetric Landau-Zener problem}
436: \label{LZa}
437:
438: \begin{figure}
439: \includegraphics[width=\columnwidth, clip=true]{fig3.eps}
440: \caption{The same as in Fig. \ref{gapanty} but for nonsymmetric Landau-Zener
441: problem with $\delta>1$ -- see (\ref{Hdelta}).}
442: \label{asym}
443: \end{figure}
444:
445: We assume that system Hamiltonian is provided by
446: the following expression
447: \begin{equation}
448: H=\frac{1}{2}
449: \left(
450: \begin{array}{cc}
451: \frac{1}{\chi}\frac{t}{\tau_Q} & 1 \\
452: 1 & -\frac{1}{\chi}\frac{t}{\tau_Q}
453: \end{array}
454: \right) \ \ \ , \ \ \
455: \chi= \left\{
456: \begin{array}{c}
457: 1 \ {\rm for} \ t\le0 \\
458: \delta \ {\rm for} \ t>0 \\
459: \end{array}
460: \right.
461: \label{Hdelta}
462: \end{equation}
463: with $\delta>0$ being the asymmetry parameter -- see Fig. \ref{asym}(a) for
464: schematic plot of the spectrum.
465:
466: Once again, evolution starts at $t_i\to-\infty$ from a ground state.
467: The exact expression for finding the system
468: in the excited eigenstate at the end of time evolution ($t_f\to+\infty)$ is
469: \begin{eqnarray}
470: \label{exact_nonsym}
471: P&=&1-
472: \frac{e^{-\frac{1}{8}\pi(1+\delta)\tau_Q}}{2}
473: \sinh\left(\frac{1}{4}\pi\tau_Q\delta\right)\nonumber\\&&
474: \left|\frac{\Gamma(1/2+i\tau_Q\delta/8)}{\Gamma(1/2+i\tau_Q/8)}+
475: \sqrt{\frac{1}{\delta}}
476: \frac{\Gamma(1+i\tau_Q\delta/8)}{\Gamma(1+i\tau_Q/8)}\right|^2,
477: \end{eqnarray}
478: and its derivation is presented in Appendix \ref{a_lz}.
479: Naturally, for $\delta=1$, i.e., in a symmetric LZ problem,
480: the expression (\ref{exact_nonsym}) reduces to (\ref{dupa987}).
481:
482: Now we would like to compare (\ref{exact_nonsym}) to predictions coming from
483: AI approximation. Due to asymmetry of the Hamiltonian the systems enters
484: the impulse regime in the time interval $[-\hat{t}_L,\hat{t}_R]$
485: -- see Fig. \ref{asym}(b)
486: for illustration of these concepts.
487: The instants $\hat{t}_L$ and $\hat{t}_R$
488: are easily found in the same way as in the symmetric case.
489: It is a straightforward exercise to verify that according to
490: AI approximation, the probability of finding the system in the upper state is
491: \begin{equation}
492: \label{nonsym}
493: P_{AI}= |\langle \downarrow(-\hat{t}_L)|\uparrow(\hat{t}_R)\rangle|^2.
494: \end{equation}
495:
496: To simplify comparison between exact (\ref{exact_nonsym}) and approximate
497: (\ref{nonsym}) excitation probabilities we will present their diabatic Taylor expansions:
498: \begin{eqnarray}
499: P_{AI}&=& 1- \frac{1}{4}\left( 1+\sqrt{\delta}\right)^2 \alpha\tau_Q +
500: \nonumber\\&&
501: \frac{1}{16}(1+\delta)(1+\sqrt{\delta})^2 (\alpha\tau_Q)^2+
502: {\cal O}(\tau_Q^3).
503: \label{dupa}
504: \end{eqnarray}
505: Determination of the constant $\alpha$ is easy and is presented in
506: Appendix \ref{a_l}. By putting $\eta=1/2$ into (\ref{alfa_1}) and (\ref{2nasym}), one
507: gets that $\alpha=\pi/2$.
508:
509: The AI prediction can be easily compared to the exact result (\ref{exact_nonsym})
510: after series expansion
511: $$
512: P=
513: 1- \frac{\pi}{8}\left( 1+\sqrt{\delta}\right)^2 \tau_Q +
514: \frac{\pi^2}{64}(1+\delta)(1+\sqrt{\delta})^2 \tau_Q^2
515: + {\cal O}(\tau_Q^3).
516: $$
517: This comparison shows that once again there is perfect matching between exact and
518: AI description up to ${\cal O}(\tau_Q^3)$, i.e., one order better then the simple
519: diabatic approximation from Appendix \ref{a_l}, despite the fact that we deal
520: now with nonsymmetric LZ problem.
521: Moreover, the constant $\alpha$ is the same in symmetric and nonsymmetric cases.
522: Finally the same powers of $\tau_Q$ show up in both exact and AI results.
523:
524: The nonsymmetric Landau-Zener model is also interesting in the light of
525: a recent paper \cite{ray}, where it is argued that the final state of the
526: system that passed through a classical
527: phase transition point, is
528: determined by details of dynamics after phase transition.
529: In our system, there are two different quench rates: $\tau_Q$
530: before the transition and $\tau_Q'=\tau_Q\delta$ after the transition.
531: Substitution of $\delta=\tau_Q'/\tau_Q$ into
532: (\ref{exact_nonsym}) shows that the final state of nonsymmetric Landau-Zener
533: model expressed in terms of the excitation probability depends on both $\tau_Q$ and
534: $\tau_Q'$, so that it behaves differently than the system described in
535: \cite{ray}.
536: %%%%%%%%%%%
537:
538:
539: \subsection{Landau-Zener problem when time evolution starts at the
540: anti-crossing center}
541: \label{LZb}
542: Now we consider the LZ problem characterized by Hamiltonian (\ref{H}) in the
543: case when time evolution starts from the ground state at the anticrossing
544: center, i.e., $t_i=0$. It means that
545: $|\Psi(0)\rangle\propto\left(|1\rangle-|2\rangle\right)/\sqrt{2}$.
546: Excitation probability at $t_f\to+\infty$ equals exactly (see Appendix \ref{a_lz})
547: \begin{eqnarray}
548: p&=& 1- \frac{2}{\pi\tau_Q} \sinh\left(\frac{\pi\tau_Q}{4}\right)e^{-\pi\tau_Q/8}
549: \left|\Gamma\left(1+\frac{i\tau_Q}{8}\right)+\right. \nonumber\\&&\left.
550: e^{i\pi/4}\sqrt{\frac{\tau_Q}{8}}
551: \Gamma\left(\frac{1}{2}+\frac{i\tau_Q}{8}\right)\right|^2.
552: \label{srodek_ex}
553: \end{eqnarray}
554:
555: From the point of view of AI approximation the evolution is now simplified by
556: assuming that it is impulse from $t_i=0$ to $\hat{t}$ and then adiabatic from
557: $\hat{t}$ to $t_f\to+\infty$. This case was discussed in \cite{bodzio}, and it
558: was shown that substitution of (\ref{hatt}) into (\ref{srodek}) leads to a
559: simple expression
560: \begin{equation}
561: p_{AI}=
562: \frac{1}{2}-\frac{1}{2}\sqrt{1-\frac{2}{(\alpha\tau_Q)^2+
563: \alpha\tau_Q\sqrt{(\alpha\tau_Q)^2+4}+2}},
564: \label{ai_full}
565: \end{equation}
566: expanding it into diabatic series one gets
567: \begin{equation}
568: p_{AI}= \frac{1}{2}-\frac{1}{2}\sqrt{\alpha\tau_Q}+\frac{1}{8}(\alpha\tau_Q)^{3/2}+
569: {\cal O}(\tau_Q^{5/2}).
570: \label{srodek_series_ai}
571: \end{equation}
572: Determination of a constant $\alpha$ is straightforward
573: (see Appendix \ref{a_l}). Putting $\eta=1/2$ into (\ref{infty_2},\ref{alfa_2})
574: one easily gets $\alpha=\pi/4$ and confirms that the first nontrivial term
575: in (\ref{srodek_series_ai}) is indeed $\propto\sqrt{\tau_Q}$.
576: This value of $\alpha$ ($\approx0.785$)
577: is in agreement with the numerical fit done in \cite{bodzio}
578: where optimal $\alpha$ was found to be equal to $0.77$. Small disagreement
579: comes from the fact that now we use just the
580: $\frac{1}{2}-\frac{1}{2}\sqrt{\alpha\tau_Q}$ part for getting $\alpha$,
581: which is equivalent to making the fit to exact results in the limit of
582: $\tau_Q\ll1$. In \cite{bodzio} the fit of the whole expression
583: (\ref{ai_full}) in the range of $0<\tau_Q<6$ was done.
584:
585: Having the exact solution (\ref{srodek_ex}) at hand, we can verify
586: rigorously accuracy of (\ref{ai_full},\ref{srodek_series_ai}) with the choice
587: of $\alpha=\pi/4$. Expanding
588: (\ref{srodek_ex}) into diabatic series one gets
589: \begin{equation}
590: p= \frac{1}{2}-\frac{\sqrt{\pi}}{4}\sqrt{\tau_Q}+
591: \frac{\pi^{3/2}}{64}\left(2-\frac{4\ln2}{\pi}\right)\tau_Q^{3/2}+
592: {\cal O}(\tau_Q^{5/2}).
593: \label{srodek_series}
594: \end{equation}
595: As expected the first nontrivial term is in perfect agreement with (\ref{srodek_series_ai})
596: once $\alpha=\pi/4$. On the other hand, the term
597: proportional to $\tau_Q^{3/2}$ is about $10\%$ off. Indeed,
598: after substitution $\alpha=\pi/4$ into
599: (\ref{srodek_series_ai}) one gets the third term equal to
600: $\pi^{3/2}\tau_Q^{3/2}/64$, which has to be compared to
601: $\approx1.1\pi^{3/2}\tau_Q^{3/2}/64$ from (\ref{srodek_series}).
602: It shows that the AI approximation in this case does not predict exactly
603: higher nontrivial terms in the diabatic expansion.
604:
605: Nonetheless, expression (\ref{ai_full}) works much better then the above
606: comparison suggests. It not only significantly outperforms the
607: lowest order diabatic expansion,
608: $\frac{1}{2}-\frac{1}{4}\sqrt{\pi\tau_Q}$,
609: but also beautifully fits to the exact result pretty far from the $\tau_Q\ll1$
610: regime.
611: Fig. \ref{srodek_plot} leaves no doubt about these statements.
612: It suggests that
613: AI predictions should be compared to exact results
614: not only term by term as in the diabatic expansion, but also in a full form.
615:
616:
617: \begin{figure}
618: \includegraphics[width=\columnwidth, clip=true]{fig4.eps}
619: \caption{Transition probability when the system starts time evolution
620: at the anti-crossing center from a ground state and evolves to $t_f\to+\infty$.
621: Dots: exact expression (\ref{srodek_ex}). Solid line: AI prediction
622: (\ref{ai_full}) with $\alpha=\pi/4$ determined from the diabatic solution
623: in Appendix \ref{a_l}. Dashed thick line: lowest order diabatic prediction
624: $1/2-\sqrt{\pi\tau_Q}/4$ coming from (\ref{infty_2},\ref{alfa_2}) with $\eta=1/2$.
625: }
626: \label{srodek_plot}
627: \end{figure}
628:
629: %%%%%%%%%%%
630:
631: \subsection{One half of Landau-Zener evolution}
632: \label{LZc}
633: In this section we consider exactly one half of the LZ problem. Namely,
634: we take the Hamiltonian (\ref{H}) and evolve the system to the anti-crossing
635: center, $t_f=0$, while starting from the ground state at $t_i\to-\infty$.
636:
637: Let's see what the AI approximation predicts for excitation probability
638: of the system at $t_f=0$. According to AI, evolution is
639: adiabatic in the interval $[-\infty, -\hat{t}\,]$ and then impulse in time range
640: $[-\hat{t},0]$.
641: It implies that
642: $$|\langle\uparrow(0)|\Psi(0)\rangle|^2\approx
643: |\langle\uparrow(0)|\Psi(-\hat{t}\,)\rangle|^2\approx|\langle\uparrow(0)|\downarrow(-\hat{t}\,)
644: \rangle|^2.
645: $$
646: A simple calculation then shows that
647: \begin{equation}
648: P_{AI}= |\langle\uparrow(0)|\downarrow(-\hat{t}\,)\rangle|^2=
649: \frac{1}{2}\left(1-\frac{1}{\sqrt{1+\hat{\varepsilon}^2}}\right).
650: \label{half_ai}
651: \end{equation}
652: A quick look at (\ref{srodek}) shows that the excitation probability for
653: one-half of the Landau-Zener problem is supposed to be, according to the AI scheme,
654: equal to the excitation
655: probability when the system starts from the anti-crossing and evolves toward
656: $t_f\to+\infty$.
657:
658: To check that prediction we have solved the one-half LZ model exactly, see
659: Appendix \ref{a_lz}, and found that indeed both probabilities are exactly
660: the same, so the AI approximation provides us here with a prediction that is
661: correct not only qualitatively but also quantitatively.
662:
663: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
664: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
665: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
666:
667: \section{Dynamics in a general class of quantum two level systems}
668: \label{2n}
669: To test predictions coming from the AI approximation on two level systems different
670: than the classic
671: Landau-Zener model, we consider in this section dynamics induced by the Hamiltonian
672: \begin{equation}
673: H=\label{H2n}
674: \frac{1}{2}
675: \left(
676: \begin{array}{cc}
677: {\rm sgn}(t)\left|\frac{t}{\tau_Q}\right|^\frac{\eta}{1-\eta} & 1 \\
678: 1 & -{\rm sgn}(t)\left|\frac{t}{\tau_Q}\right|^\frac{\eta}{1-\eta}
679: \end{array}
680: \right),
681: \end{equation}
682: where $\eta\in(0,1)$ is a constant parameter
683: (when $\eta=1/2$ the system reduces to the Landau-Zener model).
684: The eigenstates of Hamiltonian (\ref{H2n}) are expressed by formula (\ref{eig})
685: with
686: $$
687: \cos(\theta)= \frac{{\rm sgn}(t)\left|\frac{t}{\tau_Q}\right|^\frac{\eta}{1-\eta}}{
688: \sqrt{1+\left|\frac{t}{\tau_Q}\right|^\frac{2\eta}{1-\eta}}},
689: $$
690: and
691: $$
692: \sin(\theta)= \frac{1}{\sqrt{1+\left|\frac{t}{\tau_Q}\right|^\frac{2\eta}{1-\eta}}}.
693: $$
694: Extending the notation of Sec. \ref{LZ} to the system (\ref{H2n}) one has to
695: replace definition (\ref{varepsilon}) by
696: $$
697: \varepsilon= {\rm sgn}(t)\left|\frac{t}{\tau_Q}\right|^\frac{\eta}{1-\eta}
698: $$
699: and then expressions (\ref{infty},\ref{srodek}) providing excitation
700: probabilities are valid also for the model (\ref{H2n}).
701:
702:
703: The instant $\hat{t}$ for the Hamiltonian (\ref{H2n}) is determined by
704: the following version of Eq. (\ref{bodzik})
705: \begin{equation}
706: \label{gap2n}
707: \frac{1}{\sqrt{1+\left|\frac{\hat{t}}{\tau_Q}\right|^\frac{2\eta}{1-\eta}}}
708: =\alpha \hat{t}.
709: \end{equation}
710: It seems to be impossible to obtain solution of (\ref{gap2n}) exactly. Fortunately,
711: a diabatic perturbative expansion for $\hat{t}$
712: can be found -- see Appendix \ref{a_r}. Once the
713: diabatic expansion of $\hat{t}$ is known, we can present different predictions
714: about dynamics of the model (\ref{H2n}).
715:
716: First, we consider transitions starting from a ground state at $t_i\to-\infty$ and
717: lasting till $t_f\to+\infty$.
718: Using the results of Appendix \ref{a_r} one gets the following expression for
719: the excitation probability:
720: \begin{eqnarray}
721: P_{AI}&=&\frac{\hat{\varepsilon}^2}{1+\hat{\varepsilon}^2}= 1-(\alpha\tau_Q)^{2\eta}+
722: (1-\eta)(\alpha\tau_Q)^{4\eta}-\nonumber\\&&
723: \frac{1}{2}(1-\eta)(2-3\eta)(\alpha\tau_Q)^{6\eta}+\frac{1}{3}(1-\eta)(1-2\eta)
724: \nonumber\\ & &(3-4\eta)(\alpha\tau_Q)^{8\eta} +
725: {\cal O}\left(\tau_Q^{10\eta}\right).
726: \label{infty2n}
727: \end{eqnarray}
728: This result is equivalent to (\ref{dupa1}) when one puts $\eta=1/2$. The constant
729: $\alpha$ was found in Appendix \ref{a_l} from the diabatic solution
730: and equals (\ref{alfa_1})
731: $$\alpha= (1-\eta)\Gamma(1-\eta)^\frac{1}{\eta}.$$
732: The exponent $2\eta$ in the lowest order term in (\ref{infty2n})
733: was positively verified (Appendix \ref{a_l}).
734: Additionally, we have performed numerical simulations for $\eta=1/3,2/3, 3/4, 5/6$.
735: A fit to numerics in the range of $\tau_Q\ll1$ has confirmed both the
736: exponent $2\eta$ and the value of $\alpha$. The comparison between the
737: AI prediction for $\eta=1/3,2/3$ and numerics is presented in Fig. \ref{p_excitations}(a).
738: The lack of exact solution for this problem does not allow us for a more
739: systematic analytic investigation of the AI approximation in this case.
740: Nonetheless, a good agreement between AI and numerics is easily noticed.
741: Once again, the AI prediction outperforms the lowest order diabatic
742: expansion.
743:
744: As before, we would like to consider transitions starting
745: from the anticrossing center,
746: i.e., $t_i=0$. In this case the excitation probability equals
747: \begin{eqnarray}
748: p_{AI}&=&\frac{1}{2}\left(1-\frac{1}{\sqrt{1+\hat{\varepsilon}^2}}\right)=
749: \frac{1}{2}-\frac{1}{2}(\alpha\tau_Q)^\eta+ \frac{1}{4}(1-\eta)\nonumber\\ &&
750: (\alpha\tau)^{3\eta}- \frac{1}{16}(1-\eta)(3-5\eta)
751: (\alpha\tau_Q)^{5\eta} + \nonumber\\ && {\cal
752: O}\left(\tau_Q^{7\eta}\right),
753: \label{srodek2n}
754: \end{eqnarray}
755: As above, the diabatic calculation of Appendix
756: \ref{a_l} verifies the exponent of the first nontrivial term in
757: (\ref{srodek2n}) and provides us with the following prediction for $\alpha$
758: (\ref{alfa_2})
759: $$
760: \alpha= (1-\eta)\Gamma(1-\eta)^\frac{1}{\eta}
761: \cos\left(\frac{\pi}{2}\eta\right)^\frac{1}{\eta}.
762: $$
763: After setting $\eta=1/2$ formula (\ref{srodek2n})
764: becomes the same as (\ref{srodek_series_ai}). Due to the lack of exact results
765: we have carried out numerical simulations for $\eta=1/3,2/3,3/4, 5/6$. As expected,
766: this calculation has confirmed that the exponent of $\tau_Q$ is indeed
767: $\eta$ for small $\tau_Q$ and that $\alpha$ is
768: provided by (\ref{alfa_2}). The AI approximation vs. numerics and diabatic
769: expansion for $\eta=1/3,2/3$ is presented in Fig. \ref{p_excitations}(b).
770: Once again an overall agreement is strikingly good.
771:
772: \begin{figure}
773: \includegraphics[width=\columnwidth, clip=true]{fig5a.eps}\\
774: \includegraphics[width=\columnwidth, clip=true]{fig5b.eps}
775: \caption{Numerics: bold dashed line, AI approximation: solid line, lowest
776: order exact diabatic expansion (Appendix \ref{a_l}): dotted line.
777: (a): evolution starts from $t_i\to-\infty$. (b): evolution starts
778: from $t_i=0$. In both (a) and (b) plots evolution ends in $t_f\to+\infty$.
779: }
780: \label{p_excitations}
781: \end{figure}
782:
783: \section{Quantum phase transition in Ising model vs. adiabatic-impulse
784: approach}
785: \label{ising}
786:
787: In this section we will illustrate how ideas coming from the adiabatic-impulse
788: approach can be used in studies of dynamics of quantum phase transition. As an
789: example we choose the quantum Ising model
790: recently considered
791: in this context \cite{dorner,jacek,polkovnikov}.
792: The quantum Ising model is defined by the following Hamiltonian:
793: \begin{equation}
794: \label{His}
795: H=-\sum_{n=1}^N\left(g\sigma^x_n+\sigma^z_n\sigma^z_{n+1}\right),
796: \end{equation}
797: where $N$ is the number of spins.
798: The quantum phase transition in this
799: model is driven by the change of a dimensionless coupling $g$. The transition point
800: between paramagnetic phase ($g>1$) and ferromagnetic one ($0\le g<1$)
801: is at $g=1$. To study dynamics of (\ref{His}) one assumes
802: that the system evolves from time $t'=-\infty$
803: to time $t'=0$, and takes
804: $$
805: g= -\frac{t'}{\tau_Q'},
806: $$
807: where $\tau_Q'$ provides a quench rate. The quantity of interest is
808: density of topological defects (kinks) after
809: completion of the transition, i.e., at $g=0$. It equals \cite{jacek}
810: \begin{equation}
811: \label{n}
812: n=\lim_{N\to+\infty}\left\langle\frac{1}{2N}\sum_{n=1}^{N-1}
813: (1-\sigma^z_n\sigma^z_{n+1})\right\rangle.
814: \end{equation}
815: The quantum Ising model obviously possesses $2^N$ different energy
816: eigenstates so it seems to be hopeless to expect that the two level
817: approximation would be sufficient. Therefore, it is a remarkable
818: result of Dziarmaga \cite{jacek}, that dynamics in this system can be
819: exactly described by a series of uncoupled Landau-Zener systems.
820: Due to lack of space, we refer the reader to \cite{jacek},
821: and present just the main results and their AI equivalents.
822:
823: The density of defects (\ref{n}) in Dziarmaga's notation reads as
824: $$
825: n\stackrel{N\gg1}{=}\frac{1}{2\pi}\int_{-\pi}^\pi{\rm d}k\,p_k,
826: $$
827: where $p_k$'s are defined as
828: \begin{equation}
829: \label{pk}
830: p_k=|\cos(k/2)u_k(t'=0)-\sin(k/2)v_k(t'=0)|^2,
831: \end{equation}
832: with $u_k$, $v_k$ satisfying the following Landau-Zener system
833: \begin{equation}
834: \label{lz_ising}
835: \frac{d}{dt}
836: \left(
837: \begin{array}{c}
838: v_k \\ u_k
839: \end{array}
840: \right)
841: =
842: \frac{1}{2}
843: \left(
844: \begin{array}{cc}
845: \frac{t}{\tau_Q} & 1 \\
846: 1 & -\frac{t}{\tau_Q}
847: \end{array}
848: \right)
849: \left(
850: \begin{array}{c}
851: v_k \\ u_k
852: \end{array}
853: \right),
854: \end{equation}
855: $$t=4\left(t'+ \tau_Q'\cos k\right)\sin k,$$
856: $$\tau_Q= 4\tau_Q'\sin^2k,$$
857: where $t'$ changes from $-\infty$
858: to $0$. The initial conditions for that LZ system are the following
859: \begin{equation}
860: |u_k(t'=-\infty)|=1 \ \ \ , \ \ \ v_k(t'=-\infty)=0.
861: \label{ukvk}
862: \end{equation}
863: Due to the symmetry of the whole problem, it is
864: possible to show that $p_k\equiv p_{-k}$, and therefore it is sufficient
865: to restrict to $k\ge0$ from now on.
866:
867: Evolution in (\ref{lz_ising})
868: lasts from $t_i=-\infty$ to $t_f=2\tau_Q'\sin(2k)$. Since obviously $t_f$
869: does not go to $+\infty$, it is important to determine where $t_f$ is placed
870: with respect to $\pm\hat{t}$.
871:
872: We solve the equation $\hat{t}=|t_f|$ (where we substitute $k=k_c$). After
873: using (\ref{hatt}) with $\alpha=\pi/2$, and simple algebra one arrives at
874: $$
875: \label{ising_end}
876: \frac{1}{\sqrt{2}}\sqrt{\sqrt{1+\frac{4}{(\pi\tau_Q/2)^2}}-1}=\left|\frac{\cos
877: k_c}{\sin k_c}\right|,
878: $$
879: which leads to the result
880: $$
881: \sin^2 k_c = 1- \frac{1}{4\pi^2\tau_Q'^{\,2}}.
882: $$
883: Defining $k_c$ to be between $0$ and $\pi/2$ and doing
884: some additional easy calculations one gets that
885: when $\tau_Q'>\frac{1}{2\pi}$ we have: (i) $t_f\ge\hat{t}$ for $k\in[0,k_c]$;
886: (ii) $-\hat{t}< t_f < \hat{t}$ for $k\in(k_c,\pi-k_c)$; (iii)
887: $t_f\le-\hat{t}$ for $k\in [\pi-k_c,\pi]$.
888: Moreover, when $\tau_Q'<\frac{1}{2\pi}$ one can show that
889: $|t_f|<\hat{t}$ for any $k$'s of interest. It means that the evolution
890: ends in the impulse regime for small enough $\tau_Q'$.
891:
892: Now, as in \cite{dorner,jacek,polkovnikov}, we consider adiabatic time evolutions of the Ising
893: model. In our scheme
894: it clearly corresponds to $\tau_Q'\gg\frac{1}{2\pi}$. As the system undergoes
895: slow evolution it is safe to assume that only long wavelength modes are
896: excited, which means that $k\ll\pi/4$ are of interest. This allows
897: to approximate $\sin k\sim k$ and $\cos k\sim1$, which
898: implies that (\ref{pk}) turns into $p_k\approx|u_k(t_f)|^2$ and that
899: $$
900: n\approx\frac{1}{\pi}\int_0^{\epsilon}{\rm d}k\,|u_k(t_f)|^2,
901: $$
902: with $\epsilon\ll\pi/4$. Since $k\in[0,\epsilon]$ corresponds to the above
903: mentioned (i) case, $t_f>\hat{t}$ , the AI approximation says that
904: $|u_f(t_f)|\approx|u_f(+\infty)|$. Initial conditions (\ref{ukvk}) mean that
905: the system starts time evolution at $t=-\infty$ from {\it excited} state and
906: we are interested in probability of finding it in the {\it ground} state
907: at $t_f$. Elementary algebra based on AI approximation
908: shows that this probability equals
909: $$
910: |u_k(t_f)|^2=|\langle \uparrow(-\hat{t}\,)|\downarrow(\hat{t}\,)\rangle|^2=
911: |\langle \downarrow(-\hat{t}\,)|\uparrow(\hat{t}\,)\rangle|^2.
912: $$
913: Therefore, it is provided by (\ref{infty}) and (\ref{infty_ful}).
914: For any fixed $\epsilon$ consistent with lowest order approximation
915: of $\sin k$ and $\cos k$ one gets
916: \begin{equation}
917: \label{n_final}
918: n\cong 0.172\frac{1}{\sqrt{2\tau_Q'}},
919: \end{equation}
920: where the prefactor, was found from expansion of the
921: integral into $1/\sqrt{\tau_Q'}$ adiabatic series with $\tau_Q'\to+\infty$ at fixed $\epsilon$.
922: In the derivation of this result $\sin^2 k$ in the expression for $\tau_Q$ was
923: approximated by $k^2$.
924:
925: First of all, the prediction (\ref{n_final})
926: provides correct scaling
927: of defect density with $\tau_Q'$. The prefactor, $0.172$, has to be compared to
928: $\frac{1}{2\pi}\approx0.16$ (exact result from \cite{jacek} and numerical
929: estimation from \cite{dorner}) or $0.18$ (approximate result from \cite{polkovnikov}).
930: Our prefactor matches these results remarkably closely
931: concerning simplicity of the whole AI approximation. Slight overestimation
932: of the prefactor in comparison to exact result
933: comes from the fact that AI transition probability, Eq. (\ref{infty_ful}),
934: overestimates the exact result, Eq. (\ref{dupa987}), for large $\tau_Q$'s (Fig. \ref{dupa123}).
935: It is interesting, to note that in the Kibble-Zurek scheme used for
936: description of classical phase transitions an overestimation
937: of defect density is usually of the order of ${\cal O}(1)$
938: \cite{dorner,antunes,laguna}, while here discrepancy is smaller than $10\%$.
939:
940:
941:
942: Therefore, the AI approximation provides us with two predictions concerning
943: dynamics of the quantum Ising model: (i) it estimates when evolution ends
944: in the asymptotic limit; (ii) it correctly
945: predicts scaling exponent and number of defects produced during
946: adiabatic transitions.
947: The part (ii) can be calculated exactly in the quantum Ising model (\ref{His}).
948: Nonetheless, in other systems undergoing quantum phase transition
949: the two level simplification might be too difficult for
950: exact analytic treatment, e.g., as in the system governed by (\ref{H2n}). Then
951: the AI analysis might be the only analytic approach working beyond
952: the lowest order diabatic expansion.
953: Besides that the AI approach provides us with a quite intuitive
954: description of system dynamics, which is of interest in its own. Especially, when
955: one looks at connections between classical and quantum phase transitions.
956:
957: \section{Summary}
958:
959: We have shown that the adiabatic-impulse approximation,
960: based on the ideas underlying Kibble-Zurek mechanism \cite{kibble,zurek},
961: provides good quantitative predictions concerning
962: diabatic dynamics of two level Landau-Zener like systems.
963: After supplementing the splitting of the evolution
964: into adiabatic and impulse regimes by the exact lowest order
965: diabatic calculation, the whole approach is complete and
966: can be in principle applied to different systems
967: possessing anti-crossings, e.g., those lacking exact analytic
968: solution as the model (\ref{H2n}). We expect that the AI approach
969: will provide the link between dynamics of classical and quantum
970: phase transition thanks to usage of the same terminology and
971: similar assumptions.
972:
973: We also expect that different variations of the classic Landau-Zener system
974: discussed in this paper can be experimentally realized
975: in the setting where the sweep rate can be manipulated.
976: Indeed, the model (\ref{H2n})
977: arises once a proper nonlinear change of external system parameter is
978: performed. The experimental access to studies of nonsymmetric Landau-Zener model can be
979: obtained by change of the sweep rate after passing an anti-crossing center.
980: The case of evolution starting from or ending at the anticrossing center
981: can also be subjected to experimental investigations, e.g., in a beautiful
982: system consisting of the smallest available quantum magnets (cold Fe$_8$
983: clusters) \cite{magnet}.
984:
985: We are grateful to Uwe Dorner for collaboration on nonsymmetric
986: Landau-Zener problem, and Jacek Dziarmaga for comments and useful
987: suggestions on the manuscript.
988: Work supported by the U.S. Department of Energy,
989: National Security Agency, and the ESF COSLAB program.
990:
991: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
992: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
993: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
994:
995: \appendix
996: \section{Exact diabatic expressions for transition probabilities}
997: \label{a_l}
998: We would like to provide lowest order exact expressions for transition
999: probabilities in a class of two level systems described by the
1000: Hamiltonian (\ref{H2n}).
1001:
1002: First, we express the wave function as
1003: \begin{eqnarray}
1004: |\Psi(t)\rangle&=&
1005: C_1(t)
1006: \exp\left(\frac{-i(1-\eta)|t|^\frac{1}{1-\eta}}{2\tau_Q^\frac{\eta}{1-\eta}}\right)
1007: |1\rangle+\nonumber\\&&
1008: C_2(t)
1009: \exp\left(\frac{i(1-\eta)|t|^\frac{1}{1-\eta}}{2\tau_Q^\frac{\eta}{1-\eta}}\right)
1010: |2\rangle,
1011: \label{psi2n}
1012: \end{eqnarray}
1013: were exponentials are $\mp i\int {\rm dt}\, {\rm sgn}(t)|t/\tau_Q|^\frac{\eta}{1-\eta}/2$.
1014: Within this representation dynamics governed by the Hamiltonian (\ref{H2n})
1015: reduces to
1016: \begin{eqnarray}
1017: \label{c1nc2n0}
1018: i \dot{C_1}(t)&=& \frac{C_2(t)}{2}
1019: \exp\left(\frac{i(1-\eta)|t|^\frac{1}{1-\eta}}{\tau_Q^\frac{\eta}{1-\eta}}\right)\\
1020: \label{c1nc2n}
1021: i \dot{C_2}(t)&=& \frac{C_1(t)}{2}
1022: \exp\left(\frac{-i(1-\eta)|t|^\frac{1}{1-\eta}}{\tau_Q^\frac{\eta}{1-\eta}}\right)
1023: \end{eqnarray}
1024:
1025: {\bf Time evolution starting from the ground state at $t_i\to-\infty$:}
1026: we would like to integrate (\ref{c1nc2n}) from $-\infty$ to $+\infty$.
1027: The initial condition is such that $C_1(-\infty)=1$ and $C_2(-\infty)=0$.
1028: The simplification comes when one assumes a very fast transition, i.e.,
1029: $\tau_Q\to0$. Then it is clear that $C_1(t)=1+{\cal O}(\tau_Q^\beta)$ with
1030: some $\beta>0$. Putting such $C_1(t)$ into (\ref{c1nc2n}) one gets
1031: \begin{eqnarray}
1032: C_2(+\infty)&=& \frac{1}{2i}\tau_Q^\eta\int_{-\infty}^{+\infty} {\rm d}x\,
1033: \exp\left(-i(1-\eta)|x|^\frac{1}{1-\eta}\right)+\nonumber\\&&
1034: ({\rm higher \ order \ terms \ in \ \tau_Q}),
1035: \end{eqnarray}
1036: which after some algebra results in
1037: \begin{eqnarray}
1038: P=|C_1(+\infty)|^2&=& 1-|C_2(+\infty)|^2= 1-
1039: (\alpha\tau_Q)^{2\eta}\nonumber+\\&&
1040: ({\rm higher \ order \ terms \ in \ \tau_Q}),
1041: \label{infty_1}
1042: \end{eqnarray}
1043: where
1044: \begin{eqnarray}
1045: \alpha= (1-\eta)\Gamma\left(1-\eta\right)^\frac{1}{\eta}.
1046: \label{alfa_1}
1047: \end{eqnarray}
1048:
1049: %%
1050:
1051: {\bf Time evolution starting from the ground state at $t_i=0$ (anti-crossing
1052: center):} since initial wave function is $(|1\rangle-|2\rangle)/\sqrt{2}$
1053: we have $C_1(0)=1/\sqrt{2}$ and $C_2(0)=-1/\sqrt{2}$ and we evolve the system
1054: till $t_f\to+\infty$ with the Hamiltonian (\ref{H2n}). For fast transitions one has
1055: that $C_1(t)=1/\sqrt{2}+{\cal O}(\tau_Q^\beta)$, where $\beta>0$ is some
1056: constant. Integrating (\ref{c1nc2n}) from $0$ to $+\infty$ one gets:
1057: \begin{eqnarray}
1058: C_2(+\infty)+\frac{1}{\sqrt{2}}&=& \frac{\tau_Q^\eta}{2\sqrt{2}i}
1059: \int_{0}^{+\infty} {\rm d}x
1060: \exp\left(-i(1-\eta)x^\frac{1}{1-\eta}\right)\nonumber\\&&
1061: +({\rm higher \ order \ terms \ in \ \tau_Q}),
1062: \end{eqnarray}
1063: which can be easily shown to lead to
1064: \begin{eqnarray}
1065: p&=&|C_1(+\infty)|^2= 1-|C_2(+\infty)|^2= \frac{1}{2}-\frac{1}{2}
1066: (\alpha\tau_Q)^\eta\nonumber+\\&&
1067: ({\rm higher \ order \ terms \ in \ \tau_Q}),
1068: \label{infty_2}
1069: \end{eqnarray}
1070: where
1071: \begin{eqnarray}
1072: \alpha= (1-\eta)\Gamma(1-\eta)^\frac{1}{\eta}
1073: \cos\left(\frac{\pi}{2}\eta\right)^\frac{1}{\eta}.
1074: \label{alfa_2}
1075: \end{eqnarray}
1076:
1077: {\bf Time evolution from $t_i\to-\infty$ to $t_f\to+\infty$ in the nonsymmetric LZ model:}
1078: we assume that the Hamiltonian is given by (\ref{H2n}) for $t\le0$, while for
1079: $t>0$ it is given by (\ref{H2n}) with $\tau_Q$ exchanged by $\tau_Q\delta$
1080: where $\delta>0$ is the asymmetry constant -- see (\ref{Hdelta}) for
1081: $\eta=1/2$ case. Integration of (\ref{c1nc2n})
1082: separately in the intervals $[-\infty,0]$ and $[0,+\infty]$
1083: leads to the following prediction
1084: \begin{eqnarray}
1085: \label{2nasym}
1086: P&=&|C_1(+\infty)|^2= 1-|C_2(+\infty)|^2= 1-\frac{1}{4}
1087: \nonumber\\&&\left(1+\delta^\eta\right)^2
1088: (\alpha\tau_Q)^{2\eta}+\nonumber\\&&
1089: ({\rm higher \ order \ terms \ in \ \tau_Q}),
1090: \end{eqnarray}
1091: with $\alpha$ given by (\ref{alfa_1}).
1092:
1093:
1094: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1095:
1096:
1097: \section{Exact solutions of the Landau-Zener system}
1098: \label{a_lz}
1099: In this Appendix we present derivation of exact solutions of the Landau-Zener problem
1100: that are used in the main part of the paper. The general exact solution for Landau-Zener
1101: model evolving
1102: according to Hamiltonian (\ref{H}) was first discussed in
1103: \cite{zener} and then elaborated in a number of papers, e.g.,
1104: \cite{vitanov}.
1105: Here we will use that general solution
1106: for getting predictions about different time evolutions.
1107: The problem is simplified by writing the wave function as
1108: (\ref{psi2n}) with $\eta=1/2$. Then one arrives at the equations
1109: (\ref{c1nc2n0}) and (\ref{c1nc2n}) once again with $\eta=1/2$. Combining the latter ones
1110: with the substitution
1111: $$
1112: U_2(t)=C_2(t)e^{\frac{it^2}{4\tau_Q}}
1113: $$
1114: one gets
1115: $$
1116: \ddot{U}_2(z)+\left(k-\frac{z^2}{4}+\frac{1}{2}\right)U_2(z)=0,
1117: $$
1118: where
1119: \begin{equation}
1120: \label{k}
1121: k=\frac{i\tau_Q}{4} \ \ , \ \ z=\frac{t}{\sqrt{\tau_Q}}e^{-i\pi/4}.
1122: \end{equation}
1123: The general solution of this equation is expressed in
1124: terms of linearly independent Weber functions $D_{-k-1}(\pm iz)$ \cite{zener,whittaker}.
1125: By combining this observation with $\eta=1/2$ version of (\ref{c1nc2n}) one gets
1126: \begin{eqnarray}
1127: |\Psi(t)\rangle&=& 2i\left[\partial_t-\frac{it}{2\tau_Q}\right]
1128: [aD_{-k-1}(iz)+bD_{-k-1}(-iz)]|1\rangle
1129: \nonumber\\&&+
1130: [aD_{-k-1}(iz)+bD_{-k-1}(-iz)]|2\rangle,
1131: \label{general}
1132: \end{eqnarray}
1133: where $z$ is defined in (\ref{k}). Though there is a simple one-to-one correspondence
1134: between $z$ and $t$, we will use both $z$ and $t$ in different expressions to shorten
1135: notation.
1136:
1137: To determine constants $a$ and $b$ in different cases one has to know
1138: the following properties of Weber functions \cite{whittaker}.
1139: First, $\forall \ |arg(s)|<3\pi/4$ one has
1140: \begin{eqnarray}
1141: D_m(s)&=& e^{-s^2/4}s^m \left[1+{\cal O}(s^{-2})\right].
1142: \label{as1}
1143: \end{eqnarray}
1144: Second, $\forall -5\pi/4<arg(s)<-\pi/4$
1145: \begin{equation}
1146: D_m(s)= e^{-im\pi}D_m(-s)+\frac{\sqrt{2\pi}}{\Gamma(-m)}e^{-i(m+1)\pi/2}D_{-m-1}(is).
1147: \label{as2}
1148: \end{equation}
1149: Third, as $s\to0$ one has
1150: \begin{equation}
1151: D_m(s)=2^{m/2}\frac{\sqrt{\pi}}{\Gamma(1/2-m/2)}+{\cal O}(s).
1152: \label{as0}
1153: \end{equation}
1154: Finally,
1155: \begin{equation}
1156: \frac{d}{ds}D_m(s)=mD_{m-1}(s)-\frac{1}{2}sD_m(s).
1157: \label{diff}
1158: \end{equation}
1159:
1160: {\bf Exact solution of nonsymmetric LZ problem:}
1161: the evolution starts at $t_i\to-\infty$, i.e., up to a phase factor
1162: $|\Psi(-\infty)\rangle\sim|1\rangle$. The Hamiltonian is given by
1163: (\ref{Hdelta}). Using (\ref{as1},\ref{as2},\ref{diff}) one finds that $a=0$ and
1164: $b=\sqrt{\tau_Q}\exp(-\pi\tau_Q/16)/2$. Substituting them into (\ref{general})
1165: results in
1166: \begin{eqnarray}
1167: |\Psi(t\le0)\rangle &=& e^{-\pi\tau_Q/16}e^{i3\pi/4}\nonumber\\&&
1168: [(k+1)D_{-k-2}(-iz)-izD_{-k-1}(-iz)]|1\rangle\nonumber\\&&
1169: +\frac{\sqrt{\tau_Q}}{2}e^{-\pi\tau_Q/16} D_{-k-1}(-iz)]|2\rangle.
1170: \end{eqnarray}
1171: Using (\ref{as0}) one finds that this solution at $t=0$ becomes
1172: \begin{eqnarray}
1173: |\Psi(0)\rangle&=& e^{-\pi\tau_Q/16} e^{i3\pi/4}
1174: \frac{\sqrt{\pi}2^{-k/2}}{\Gamma(1/2+k/2)}|1\rangle+
1175: \nonumber\\&&
1176: \frac{\sqrt{\tau_Q}}{2}
1177: e^{-\pi\tau_Q/16}
1178: \sqrt{\frac{\pi}{2}}\frac{2^{-k/2}}{\Gamma(1+k/2)} |2\rangle.
1179: \label{t0}
1180: \end{eqnarray}
1181: For $t>0$ the Hamiltonian changes its form, see (\ref{Hdelta}), and one has to match
1182: (\ref{t0}) with (\ref{general}) having $\tau_Q$ replaced by
1183: $\tau_Q\delta$. Substitution of
1184: \begin{eqnarray}
1185: a&=&\frac{e^{-\pi\tau_Q/16}}{4}2^{-k(1-\delta)/2} \sqrt{\tau_Q\delta}
1186: \left[\frac{\Gamma(1+k\delta/2)}{\Gamma(1+k/2)}\sqrt{\frac{1}{\delta}} \right.
1187: \nonumber\\&&\left.
1188: -\frac{\Gamma(1/2+k\delta/2)}{\Gamma(1/2+k/2)}\right],\nonumber
1189: \end{eqnarray}
1190: \begin{eqnarray}
1191: b&=&\frac{e^{-\pi\tau_Q/16}}{4}2^{-k(1-\delta)/2} \sqrt{\tau_Q\delta}
1192: \left[\frac{\Gamma(1+k\delta/2)}{\Gamma(1+k/2)}\sqrt{\frac{1}{\delta}} \right.
1193: \nonumber\\&&\left.
1194: +\frac{\Gamma(1/2+k\delta/2)}{\Gamma(1/2+k/2)}\right],\nonumber
1195: \end{eqnarray}
1196: into (\ref{general}) gives the wave function $|\Psi(t\ge0)\rangle$.
1197: Taking one minus squared modulus of the amplitude of finding the system in the state
1198: $|2\rangle$ one obtains the excitation probability of the system at
1199: $t_f\to+\infty$ in the form (\ref{exact_nonsym}).
1200:
1201: {\bf Exact solution of LZ problem when evolution starts from a ground state at
1202: anticrossing center:}
1203: the Hamiltonian of the system is given by (\ref{H}) and we look for a solution
1204: that starts at $t=0$ from the (ground) state:
1205: $|\Psi(0)\rangle=(|1\rangle-|2\rangle)/\sqrt{2}$. The constants $a$
1206: and $b$ from (\ref{general}) turn out to be equal to
1207: $$
1208: a= \frac{2^{k/2}\Gamma(1/2+k/2)e^{i\pi/4}\sqrt{\tau_Q}}{4\sqrt{2\pi}}
1209: -2^{k/2}\frac{\Gamma(1+k/2)}{2\sqrt{\pi}},
1210: $$
1211: $$
1212: b= -\frac{2^{k/2}\Gamma(1/2+k/2)e^{i\pi/4}\sqrt{\tau_Q}}{4\sqrt{2\pi}}
1213: -2^{k/2}\frac{\Gamma(1+k/2)}{2\sqrt{\pi}}.
1214: $$
1215: Having this at hand,
1216: it is straightforward to show that one minus squared modulus of the
1217: amplitude of finding the system in the state $|2\rangle$ at $t_f\to+\infty$,
1218: i.e., excitation probability, equals (\ref{srodek_ex}).
1219:
1220: {\bf One-half of the Landau-Zener problem:}
1221: the excitation probability of the system
1222: that started time evolution at $t_i=-\infty$ from ground state,
1223: and stopped evolution at $t_f=0$,
1224: turns out to be equal to (\ref{srodek_ex}).
1225: The simplest way to prove it comes from an observation that this
1226: probability equals one minus
1227: the squared overlap between the state $(|1\rangle-|2\rangle)/\sqrt{2}$
1228: and (\ref{t0}). Then, straightforward calculation leads immediately to
1229: expression (\ref{srodek_ex}).
1230:
1231: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1232: \section{Solution of Eq. (\ref{gap2n})}
1233: \label{a_r}
1234:
1235: \begin{figure}[t]
1236: \includegraphics[width=\columnwidth, clip=true]{fig6.eps}
1237: \caption{Schematic plot of the recurrence method for getting solution of
1238: (\ref{gap2n_new}). Solid lines $x^\frac{1}{1-\eta}$ and $\beta-x$, dashed line is a
1239: construction of a recurrence solution (\ref{rec1}) and (\ref{rec2}).
1240: The plots are in a $\beta\times\beta$ box.}
1241: \label{rec}
1242: \end{figure}
1243:
1244: Equation (\ref{gap2n}) can be solved in the following recursive way.
1245: First, one introduces new variables:
1246: \begin{equation}
1247: \label{variables}
1248: x=\left(\frac{\hat{t}}{\tau_Q}\right)^2 \ \ , \ \ \beta=\frac{1}{\alpha^2\tau_Q^2}.
1249: \end{equation}
1250: In these variables Equation (\ref{gap2n}) becomes:
1251: \begin{equation}
1252: \label{gap2n_new}
1253: x^\frac{1}{1-\eta}= \beta-x,
1254: \end{equation}
1255: and we assume that $\beta>1$ (diabatic evolutions).
1256:
1257: A quick look at Fig. \ref{rec}, makes clear that the solution can
1258: be obtained by considering a series of inequalities, numbered by index $i$,
1259: in the form
1260: \begin{equation}
1261: x_{L(i)} < x < x_{U(i)},
1262: \label{rec1}
1263: \end{equation}
1264: where both lower ($x_{L(i)}$) and upper ($x_{U(i)}$) bounds
1265: satisfy the the same recurrence equation
1266: \begin{equation}
1267: x_{L,U(i)}= \left[\beta -
1268: \left[\beta-x_{L,U(i-1)}\right]^{1-\eta}\right]^{1-\eta},
1269: \label{rec2}
1270: \end{equation}
1271: with the initial conditions $x_{L(0)}=0$, $x_{U(0)}=\beta$. Considering a few
1272: iterations one can show that the solution of (\ref{gap2n_new}) can be
1273: conveniently written as
1274: \begin{eqnarray}
1275: \label{xdelta}
1276: x^\frac{1}{1-\eta} &=& \frac{1}{(\alpha\tau_Q)^2}\left[1-
1277: (\alpha\tau)^{2\eta}+
1278: (1-\eta)(\alpha\tau_Q)^{4\eta}+\right.\nonumber\\
1279: & & \frac{1}{2}(3\eta-2)(1-\eta)(\alpha\tau_Q)^{6\eta}+
1280: \frac{1}{3}(1-\eta)(1-2\eta)\nonumber\\ & &(3-4\eta) \left.
1281: (\alpha\tau_Q)^{8\eta}+{\cal O}\left(\tau_Q^{10\eta}\right)\right].
1282: \end{eqnarray}
1283: Using (\ref{xdelta}), different quantities of interest can be determined,
1284: e.g.,
1285: \begin{eqnarray}
1286: \label{epsilon2}
1287: \hat{\varepsilon}^2&=&\frac{1}{(\alpha\tau_Q)^{2\eta}}\left[1-\eta
1288: (\alpha\tau_Q)^{2\eta}+
1289: \frac{1}{2}\eta(1-\eta)(\alpha\tau_Q)^{4\eta}- \right.\nonumber\\
1290: &&\frac{1}{3}\eta(1-\eta)(1-2\eta)(\alpha\tau_Q)^{6\eta}+
1291: \frac{1}{8}\eta(1-\eta)\nonumber\\ & & \left. (1-3\eta)(2-3\eta)(\alpha\tau_Q)^{8\eta}+
1292: {\cal O}\left(\tau_Q^{10\eta}\right)\right],\\
1293: \label{that2n}
1294: \hat{t}&=& \tau_Q^\eta\alpha^{\eta-1}+ {\cal
1295: O}\left(\tau_Q^{3\eta}\right).
1296: \end{eqnarray}
1297: Substitution of $\eta=1/2$ into (\ref{epsilon2}) and (\ref{that2n}) reproduces
1298: diabatic series expansions of $\hat{\varepsilon}$, $\hat{t}$ from the standard
1299: Landau-Zener theory (\ref{hatt}).
1300:
1301:
1302: \begin{thebibliography}{99}
1303:
1304:
1305: \bibitem{subir} S. Sachdev, {\it Quantum Phase Transitions} (Cambridge University
1306: Press, Cambridge UK, 2001).
1307:
1308: \bibitem{dorner} W.H. Zurek, U. Dorner, and P. Zoller, Phys. Rev. Lett. {\bf 95}, 105701 (2005).
1309:
1310: \bibitem{jacek} J. Dziarmaga, Phys. Rev. Lett. {\bf 95}, 245701 (2005).
1311:
1312: \bibitem{polkovnikov} A. Polkovnikov, Phys. Rev. B {\bf 72}, 161201(R) (2005).
1313:
1314: \bibitem{ralf} R. Sch\"utzhold, Phys. Rev. Lett. {\bf 95}, 135703 (2005).
1315:
1316: \bibitem{ad_q_com} E. Farhi, J. Goldstone, S. Gutmann, J. Lapan, A. Lundgren, and D. Preda,
1317: Science {\bf 292}, 472 (2001); A.M. Childs, E. Farhi, and J. Preskill, Phys. Rev. A {\bf 65},
1318: 012322 (2002).
1319:
1320: \bibitem{magnet} W. Wernsdorfer and R. Sessoli, Science {\bf 284},
1321: 133 (1999); W. Wernsdorfer {\it et al.}, J. Appl. Phys. {\bf 87},
1322: 5481 (2000).
1323:
1324: \bibitem{bloch} I. Bloch, Nature Physics {\bf 1}, 23 (2005).
1325:
1326: \bibitem{bodzio} B. Damski, Phys. Rev. Lett. {\bf 95}, 035701 (2005).
1327:
1328: \bibitem{kibble} T.W.B. Kibble, J. Phys. A {\bf 9}, 1387 (1976);
1329: Phys. Rep. {\bf 67}, 183 (1980).
1330:
1331: \bibitem{zurek} W.H. Zurek, Nature (London) {\bf 317}, 505 (1985);
1332: Acta Phys. Pol. B {\bf 24}, 1301 (1993); Phys. Rep. {\bf 276}, 177 (1996).
1333:
1334: \bibitem{antunes} N.D. Antunes, L.M.A. Bettencourt, and W.H. Zurek,
1335: Phys. Rev. Lett. {\bf 82}, 2824 (1999)
1336:
1337: \bibitem{laguna} P. Laguna and W.H. Zurek,
1338: Phys. Rev. Lett. {\bf 78}, 2519 (1997); A. Yates and W. H. Zurek, Phys. Rev.
1339: Lett. {\bf 80}, 5477 (1998).
1340:
1341: \bibitem{kz1} C. Bauerle, Y.M. Bunkov, S.N. Fisher, H. Godfrin,
1342: and G.R. Pickett, Nature (London) {\bf 382}, 332 (1996);
1343: V.M.H. Ruutu {\it et al.}, Nature (London) {\bf 382}, 334 (1996).
1344:
1345: \bibitem{kz2} M.J. Bowick, L. Chandar, E.A. Schiff, and
1346: A.M. Srivastava, Science {\bf 263}, 943 (1994);
1347: I. Chuang, R. Durrer, N. Turok, and B. Yurke,
1348: Science {\bf 251}, 1336 (1991).
1349:
1350: \bibitem{kz3} R. Carmi, E. Polturak, and G. Koren,
1351: Phys. Rev. Lett. {\bf 84} 4966 (2000);
1352: A. Maniv, E. Polturak, and G. Koren, Phys. Rev. Lett.
1353: {\bf 91}, 197001 (2003).
1354:
1355: \bibitem{kz4} R. Monaco, J. Mygind, and R.J. Rivers, Phys. Rev. Lett.
1356: {\bf 89}, 080603 (2002); Phys. Rev. B {\bf 67}, 104506 (2003). R. Monaco,
1357: U.L. Olsen, J. Mygind, R.J. Riviers, and V.P. Koshelets, cond-mat/0503707.
1358:
1359: \bibitem{zener} C. Zener, Proc. R. Soc. A {\bf 137}, 696 (1932).
1360:
1361: \bibitem{ray} N. D. Antunes, P. Gandra, R.J. Rivers, hep-ph/0504004.
1362:
1363:
1364: \bibitem{vitanov} N. V. Vitanov and B. M. Garraway, Phys. Rev. A {\bf 53},
1365: 4288 (1996); N. V. Vitanov, Phys. Rev. A {\bf 59}, 988 (1999).
1366:
1367: \bibitem{whittaker} E.T. Whittaker and G.N. Watson, {\it A Course of Modern
1368: Analysis} (Cambridge University Press, Cambridge UK, 1958).
1369:
1370:
1371: \end{thebibliography}
1372:
1373: \end{document}
1374: