1: % Submitted to PRB 17 July 2002, bu8239
2: % Include figure files
3: % Align table columns on decimal point
4: % bold math
5: % standard setting
6: %\topmargin= -32 truemm % for bloody cond-mat
7:
8:
9: \documentclass[aps,prb,showpacs,twocolumn,floats]{revtex4}
10: \usepackage{amssymb}
11:
12: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
13: \usepackage{graphicx,psfig}
14: \usepackage{dcolumn}
15: \usepackage{bm}
16:
17: %TCIDATA{OutputFilter=LATEX.DLL}
18: %TCIDATA{LastRevised=Wed Feb 05 18:29:42 2003}
19: %TCIDATA{<META NAME="GraphicsSave" CONTENT="32">}
20:
21: \newcommand{\onefourth}{\mbox{\scriptsize
22: \raisebox{1.5mm}{1}\hspace{-2.7mm}
23: \raisebox{0.5mm}{$-$}\hspace{-2.8mm}
24: \raisebox{-0.9mm}{4}\hspace{-0.7mm} \normalsize }}
25: \textheight= 243 truemm
26: \topmargin= -15 truemm
27: \newcommand{\erfc}{ {\rm erfc} }
28: \newcommand{\const}{ {\rm const} }
29: \newcommand{\arctanh}{ {\rm arctanh} }
30: \input{tcilatex}
31:
32: \begin{document}
33:
34: \title{Landau-Zener-Stueckelberg effect in a model of interacting tunneling
35: systems}
36: \author{D. A. Garanin}
37: \affiliation{ Institut f\"ur Physik,
38: Johannes-Gutenberg-Universit\"at,
39: D-55099 Mainz, Germany}
40: \date{\today}
41:
42:
43: \begin{abstract}
44: The Landau-Zener-Stueckelberg (LZS) effect in a model system of
45: interacting tunneling particles is studied numerically and
46: analytically. Each of $N$ tunneling particles interacts with each
47: of the others with the same coupling $J.$ This problem maps onto
48: that of the LZS effect for a large spin $S=N/2.$ The mean-field
49: limit $N\rightarrow \infty $ corresponds to the classical limit
50: $S\rightarrow \infty $ for the effective spin. It is shown that
51: the ferromagnetic coupling $J>0$ tends to suppress the LZS
52: transitions. For $N\rightarrow \infty $ there is a critical value
53: of $J$ above which the staying probability $P$ does not go to zero
54: in the slow sweep limit, unlike the standard LZS effect. In the
55: same limit for $J>0$ LZS transitions are boosted and $P=0$ for a
56: set of finite values of the sweep rate. Various limiting cases
57: such as strong and weak interaction, slow and fast sweep are
58: considered analytically. It is shown that the mean-field approach
59: works well for arbitrary $N$ if the interaction $J$ is weak.
60: \end{abstract}
61:
62: \pacs{03.65.-w, 75.10.Jm}
63:
64: \maketitle
65:
66:
67:
68: \section{Introduction}
69:
70: The Landau-Zener-Stueckelberg (LZS)
71: problem\cite{lan32,zen32,stu32} of quantum-mechanical transitions
72: at an avoided level crossing induced by a linear-in-time energy
73: sweep is well known in many areas of physics, in particular, in
74: the physics of atomic and molecular collisions (see, e.g., Ref.\
75: \onlinecite{crohug77}). Recently the LZS\ effect was observed in
76: crystals of single-molecule magnets and it was used to extract
77: their tunneling level splittitng $\Delta
78: $.\cite{werses99,weretal00epl}
79:
80: The specifics of the LZS effect in solid-state systems is the
81: macroscopically large number $N$ ot tunneling species such as spins of
82: magnetic molecules in the above experiments.\cite{werses99,weretal00epl}
83: These tunneling species are usually coupled with each other (mainly via the
84: dipole-dipole interaction in the case of magnetic molecules) and the
85: coupling energy $J$ can easily exceed the splitting $\Delta .$ In this
86: situation tunneling species (that can be described by pseudospins $S=1/2)$
87: do not tunnel independently, and one has to consider the Schr\"{o}dinger
88: equation for the whole system that is described by $2^{N}$ variables. The
89: latter is a new and tremendous problem since the whole Schr\"{o}dinger
90: equation becomes intractable even for a moderate number $N$ while analytical
91: treatment is difficult because the system is far from the equilibrium.
92:
93: As a plausible first step towards the solution of the LZS problem for a
94: system of interacting particles one can consider a simplified model in which
95: each particle interacts with all $N-1$ other particles with the same
96: strength $J,$ as was done in Ref.\ \onlinecite{hamraemiysai00}. For this
97: model the Schr\"{o}dinger equation simplifies so that one has to solve only $%
98: N+1$ equations instead of the $2^{N}$ equations in the general case. In the
99: limit $N\rightarrow \infty $ the problem simplifies since the mean-filed
100: approximation (MFA) becomes exact. The latter allows one to gain insights
101: into the problem by analyzing its mean-field solution numerically as well as
102: analytically in different limiting cases. Because the \emph{molecular field}
103: exerted on a tunneling particle from the others depends nonlinearly on time
104: in the region of the level crossing, the problem cannot be linearized (c.f.
105: Ref. \onlinecite{hamraemiysai00}) and\ the results differ essentially from
106: the standard LZS solution. For instance, the ferromagnetic coupling, $J>0,$
107: makes the energy sweep faster, and thus the probability $P$ to stay on the
108: initial bare level increases. The nonlinear time dependence of the sweep
109: within the MFA has stimulated Ref.\ \onlinecite{garsch02prb}, where the
110: different kinds of sweep nonlinearities were investigated for a single
111: tunneling system. Certainly the MFA for systems of interacting tunneling
112: particles is a more complicated problem since in this case the form of the
113: sweep is unknown from the beginning and it should be found self-consistently.
114:
115: It is very interesting to study the difference between the exact
116: quantum-mechanical and the mean-field solutions of the LZS problem for
117: finite $N.$ The original model maps of the model of a large spin $S=N/2$
118: with the uniaxial anisotropy $D\sim J$, transverse field $H_{x}\sim \Delta ,$
119: and the sweeped longitudinal field $H_{z}(t).$ That is, the mean-field limit
120: $N\rightarrow \infty $ of our model corresponds to the classical limit $%
121: S\rightarrow \infty $ for the effective large spin. In Ref.\ %
122: \onlinecite{hamraemiysai00} it was shown numerically that for the model with
123: $N=4$ and $J>0$ the exact and MFA results are qualitatively similar. It is
124: very interesting, however, to investigate the problem for larger $N$ and for
125: $J<0.$
126:
127: Numerical and analytical study of these problems is the objective of this
128: paper, the rest of which is organized as follows. In Sec.\ \ref
129: {Sec-Hamiltonian} the Hamiltonian is written down, the Schr\"{o}dinger
130: equation is simplified making use of the symmetry of the interaction, the
131: problem is mapped onto that of a large spin and its general properties are
132: studied. In Sec.\ \ref{Sec-separated} the case of well-separated resonances,
133: $\left| J\right| \gg \Delta $ where the problem can be reduced to that of
134: successive standard LZS transitions is investigated. In Sec.\ \ref
135: {Sec-fast-sweep} the opposite case $J\ll \Delta ,$ where the problem can be
136: solved perturbatively in $J,$ is considered. Final results are worked out in
137: the cases of fast sweep and slow sweep. In Sec.\ \ref{Sec-mfa} general
138: properties of the mean-field solution of the LZS problem with interaction
139: are investigated, in particular, with the help of the mapping onto the
140: classical-spin problem. Numerical solution shows, in particular, that for $%
141: J<0$ complete LZS transition, $P=0,$ is achieved at some values of the sweep
142: rate. Analytical treatment of the slow-sweep limit within the MFA is
143: provided in Sec.\ \ref{Sec-slow-MFA}.
144:
145: \section{The Hamiltonian}
146:
147: \label{Sec-Hamiltonian}
148:
149: We consider the model of $N$ double-level tunneling systems described by
150: pseudospins 1/2 and interacting each with each with an equal strength
151: \begin{equation}
152: \widehat{H}=-\frac{1}{2}\sum_{i=1}^{N}\left[ W(t)\sigma _{iz}+\Delta \sigma
153: _{ix}\right] -\frac{J}{2}\sum_{i\neq j=1}^{N}\sigma _{iz}\sigma _{jz}.
154: \label{Ham}
155: \end{equation}
156: Here $\mathbf{\sigma }_{i}$ are Pauli matrices, $W(t)$ is the energy sweep
157: that is taken to be linear in time
158: \begin{equation}
159: W(t)\equiv E_{-1}(t)-E_{1}(t)=vt, \label{WDef}
160: \end{equation}
161: $\Delta $ is the level splitting at resonance in the absence of interaction (%
162: $t=0$ and $J=0)$, whereas $J$ is the interaction constant. The case $J>0$
163: corresponds to the ferromagnetic (FM) coupling, whereas that of $J<0$
164: corresponds to the antiferromagnetic frustrating (AFMF) coupling. In the
165: latter case, the ground state of the system with $W=\Delta =0$ is a highly
166: degenerate state with the minimal total spin. If $J=0,$ the problem
167: simplifies to the well known LZS problem for individual tunneling systems.
168: If these system at $t\rightarrow -\infty $ are in the bare ground state $%
169: \psi _{-1}\equiv |\downarrow \rangle $ before crossing the resonance, the
170: probability to stay in this state after crossing the resonance at $%
171: t\rightarrow \infty $ \ is given by \cite{zen32,stu32,akusch92,dobzve97}
172: \begin{equation}
173: P(\infty )\equiv P=e^{-\varepsilon },\qquad \varepsilon \equiv \frac{\pi
174: \Delta ^{2}}{2\hbar v} \label{PLZ}
175: \end{equation}
176: (Probabilities without the argument are shortcuts for the final-state
177: probabilities at $t=\infty $ throughout the paper). For the problem with
178: interaction, the wave function of the system can be written as the expansion
179: over the direct-product states
180: \begin{eqnarray}
181: \Psi (t) &=&\sum_{m_{1},\ldots ,m_{N}=-1,1}C_{m_{1},\ldots ,m_{N}}(t)\psi
182: _{m_{1}}\otimes \ldots \otimes \psi _{m_{N}} \nonumber \\
183: \psi _{-1} &=&\left(
184: \begin{array}{c}
185: 0 \\
186: 1
187: \end{array}
188: \right) =|\downarrow \rangle ,\qquad \psi _{1}=\left(
189: \begin{array}{c}
190: 1 \\
191: 0
192: \end{array}
193: \right) =|\uparrow \rangle . \label{PsiDef}
194: \end{eqnarray}
195: The initial condition for $\Psi (t)$ is $C_{-1,\ldots ,-1}(-\infty )=1$
196: whereas all other coefficients are zero, i.e., the system starts in the
197: state with all spins down. With time the state of the system becomes a
198: superposition of all possible basis states in Eq.\ (\ref{PsiDef}). The
199: one-particle probability to remain in the initial state $-1$ for our $N$%
200: -particle system is given by
201: \begin{equation}
202: P_{N}(t)=\sum_{m_{2},\ldots ,m_{N}=-1,1}\left| C_{-1,m_{2},\ldots
203: ,m_{N}}(t)\right| ^{2} \label{PtDef}
204: \end{equation}
205: and it starts from $P_{N}(-\infty )=1.$
206:
207: The solution of the LZS problem for this model is simplified by the fact
208: that the coefficients $C_{m_{1},\ldots ,m_{N}}(t)$ depend only on the number
209: $k$ of spins up while they are independent on the choice of these spins.
210: Thus one can label the states by the index $k$ only that results in the
211: Schr\"{o}dinger equation
212: \begin{equation}
213: i\hbar \dot{C}_{k}=E_{k}(t)C_{k}-\frac{k\Delta }{2}C_{k-1}-\frac{(N-k)\Delta
214: }{2}C_{k+1} \label{SchrEq}
215: \end{equation}
216: for $k=0,\ldots ,N$ and with the bare energies
217: \begin{equation}
218: E_{k}(t)=\left( \frac{N}{2}-k\right) W(t)+2Jk(N-k)+\mathrm{const}.
219: \label{Enk}
220: \end{equation}
221: Each $k$-state is realized by $N!/\left[ (N-k)!k!\right] $ different choices
222: of the indices $m_{1},\ldots ,m_{N}$ in Eq.\ (\ref{PsiDef}). For $J<0,$ the
223: ground state with $W=\Delta =0$\ corresponds to $k=N/2$ and is highly
224: degenerate. In Eq.\ (\ref{PtDef}) $m_{1}$ is fixed to $-1,$ and the
225: remaining indices $m_{2},\ldots ,m_{N}$ can also be parametrized by $k.$ In
226: the corresponding number of realizations, one should use $N-1$ instead of $N$
227: that leads to
228: \begin{equation}
229: P_{N}(t)=\sum_{k=0}^{N-1}\frac{(N-1)!}{(N-1-k)!k!}\left| C_{k}(t)\right|
230: ^{2}. \label{PtDefk}
231: \end{equation}
232:
233: It is convenient to introduce the coefficients
234: \begin{equation}
235: c_{k}\equiv \sqrt{\frac{N!}{(N-k)!k!}}C_{k} \label{ckDef}
236: \end{equation}
237: that satisfy the normalization condition $1=\sum_{k=0}^{N}\left|
238: c_{k}\right| ^{2}$ and the Schr\"{o}dinger equation
239: \begin{equation}
240: i\hbar \dot{c}_{k}=E_{k}(t)c_{k}-\frac{\Delta }{2}l_{k-1,k}c_{k-1}-\frac{%
241: \Delta }{2}l_{k,k+1}c_{k+1} \label{SchrEqc}
242: \end{equation}
243: with $l_{k,k+1}\equiv \sqrt{(N-k)(k+1)}.$ Then Eq.\ (\ref{PtDefk})
244: simplifies to
245: \begin{equation}
246: P_{N}(t)=\sum_{k=0}^{N}\left( 1-\frac{k}{N}\right) p_{k}(t),\qquad
247: p_{k}(t)\equiv \left| c_{k}(t)\right| ^{2}. \label{PtFinal}
248: \end{equation}
249: Eq.\ (\ref{SchrEqc}) maps on the corresponding Schr\"{o}dinger equation for
250: a large pseudospin $S=N/2$ with the Hamiltonian
251: \begin{equation}
252: \widehat{H}_{S}=-H_{z}(t)S_{z}-H_{x}S_{x}-DS_{z}^{2}+\mathrm{const,}
253: \label{HamS}
254: \end{equation}
255: where
256: \begin{equation}
257: H_{z}(t)=W(t),\qquad H_{x}=\Delta ,\qquad D=2J \label{ParIdent}
258: \end{equation}
259: (we set $g\mu _{B}=1$ so that the ``magnetic field'' is energy dimensional).
260: Note that the excitation number $k$ is related to the eigenvalue $m$ of $%
261: S_{z}$ by $k=N/2+m.$ Thus $P(t)$ maps onto
262: \begin{equation}
263: P_{N}(t)=\frac{1}{2}\left( 1-\frac{\langle S_{z}\rangle _{t}}{S}\right) .
264: \label{PtMapping}
265: \end{equation}
266:
267: The large-spin model defined by Eq.\ (\ref{HamS}) is well known from the
268: physics of the single-molecule magnet Mn$_{12}.$ For small $\Delta $ the
269: energy levels of the system as function of $W$ are nearly straight lines
270: with avoided crossings (see Fig.\ \ref{Fig-lzn-En}). Splittings at avoided
271: crossings are or order $\Delta (\Delta /J)^{n},$ where $n$ are appropriate
272: powers.\cite{gar91jpa}
273:
274: %TCIMACRO{
275: %\TeXButton{Fig-lzn-En}{\begin{figure}[t]
276: %\unitlength1cm
277: %\begin{picture}(11,6)
278: %\centerline{\psfig{file=Fig-lzn-En-F.eps,angle=-90,width=9cm}}
279: %\end{picture}
280: %\begin{picture}(11,6)
281: %\centerline{\psfig{file=Fig-lzn-En-AF.eps,angle=-90,width=9cm}}
282: %\end{picture}
283: %\caption{ \label{Fig-lzn-En}
284: %$a$ -- Exchange-split resonances for a ferromagnetically coupled spin cluster for $|J|\gtrsim \Delta$.
285: %Transitions shown by the dashed and dotted arrows arise in higher orders in $\Delta/J$ and are much weaker.
286: %$b$ -- Same for an AFMF coupled spin cluster.
287: %}
288: %\end{figure}%
289: %}}%
290: %BeginExpansion
291: \begin{figure}[t]
292: \unitlength1cm
293: \begin{picture}(11,6)
294: \centerline{\psfig{file=Fig-lzn-En-F.eps,angle=-90,width=9cm}}
295: \end{picture}
296: \begin{picture}(11,6)
297: \centerline{\psfig{file=Fig-lzn-En-AF.eps,angle=-90,width=9cm}}
298: \end{picture}
299: \caption{ \label{Fig-lzn-En}
300: $a$ -- Exchange-split resonances for a ferromagnetically coupled spin cluster for $|J|\gtrsim \Delta$.
301: Transitions shown by the dashed and dotted arrows arise in higher orders in $\Delta/J$ and are much weaker.
302: $b$ -- Same for an AFMF coupled spin cluster.
303: }
304: \end{figure}%
305: %
306: %EndExpansion
307:
308: %TCIMACRO{
309: %\TeXButton{Fig-lzn-P3Ft}{\begin{figure}[t]
310: %\unitlength1cm
311: %\begin{picture}(11,6)
312: %\centerline{\psfig{file=Fig-lzn-P3Ft.eps,angle=-90,width=9cm}}
313: %\end{picture}
314: %\caption{ \label{Fig-lzn-P3Ft}
315: %Time dependence of the one-particle staying probability $P_N$ as function of the energy sweep $W(t)$
316: %for a system of $N=3$ ferromagnetically coupled particles.
317: %At intermediate sweep rate, $\varepsilon=1$ only the first-order LZS transition at $W=2J_0$ (the $\Delta$ transition) is seen.
318: %For rather slow sweep rates (here $\varepsilon=10$), higher-order transitions (here the $\Delta^2/J$ transition at $W=0$) are switching on.
319: %}
320: %\end{figure}%
321: %}}%
322: %BeginExpansion
323: \begin{figure}[t]
324: \unitlength1cm
325: \begin{picture}(11,6)
326: \centerline{\psfig{file=Fig-lzn-P3Ft.eps,angle=-90,width=9cm}}
327: \end{picture}
328: \caption{ \label{Fig-lzn-P3Ft}
329: Time dependence of the one-particle staying probability $P_N$ as function of the energy sweep $W(t)$
330: for a system of $N=3$ ferromagnetically coupled particles.
331: At intermediate sweep rate, $\varepsilon=1$ only the first-order LZS transition at $W=2J_0$ (the $\Delta$ transition) is seen.
332: For rather slow sweep rates (here $\varepsilon=10$), higher-order transitions (here the $\Delta^2/J$ transition at $W=0$) are switching on.
333: }
334: \end{figure}%
335: %
336: %EndExpansion
337:
338: Our model described by Eqs.\ (\ref{Ham}) or (\ref{HamS}), although not
339: completely realistic, allows one to trace down the influence of the
340: interaction on the transition probabilities and to obtain a number of
341: interesting analytical results. We will see that, in particular, suppression
342: of transitions for the ferromagnetic interactions ($J>0)$ can be easily
343: understood. An advantage of this simplified model is the possibility to
344: solve the problem numerically up to much larger system sizes, compared to
345: models with realistic interactions. For the latter, the number of different
346: coefficients in Eq.\ (\ref{PsiDef}) and thus of differential equations to
347: solve is $2^{N},$ whereas in our case there are only $N+1$ equations.
348:
349: For numerical calculations we use Wolfram Mathematica 4.0 that employs, in
350: particular, a very accurate differential equation solver needed for handling
351: strongly oscillating solutions over large time intervals. The results of
352: solving Eq.\ (\ref{SchrEqc}) are shown in Fig.\ \ref{Fig-lzn-P3Ft} for a
353: cluster of three ferromagnetically coupled particles.
354:
355: \section{Well-separated resonances}
356:
357: \label{Sec-separated}
358:
359: The interaction can profoundly influence the LZS effect if it is strong
360: enough, \TEXTsymbol{\vert}$J|\gtrsim \Delta $. The general tendency can be
361: already seen from the well-known mean-field argument. If one of the
362: tunneling particles flips to another bare state during the resonance
363: crossing, then the total field (the external sweep field plus the molecular
364: field) on all other particles jumps. For the ferromagnetic coupling, $J>0,$
365: the jump of the total field is positive, other particles are brought past
366: the resonance and lose their chance to flip, and thus transition probability
367: is suppressed. For the antiferromagnetic frustrating coupling, the jump of
368: the total field is negative and other particles are being set back before
369: the resonance and are getting one more chance to cross the resonance at some
370: larger field value and flip; Thus the transition probability for the system
371: should increase.
372:
373: The influence of the interaction can be readily illustrated within a
374: rigorous many-body quantum language for our model if one considers the
375: energy levels of the system shown in Fig.\ \ref{Fig-lzn-En}. One can see
376: that instead of a single resonance at $W=0$ for individual or noninteracting
377: particles there is an interaction-split resonance that consists of many
378: avoided line crossings. These avoided line crossings are well separated from
379: each other on the plot if \TEXTsymbol{\vert}$J|\gg \Delta .$ The question
380: whether these well-separated crossings can be considered as a succession of
381: independent LZS transitions (i.e., whether they are \emph{dynamically} well
382: separated) depends on the sweep rate. Whereas for the slow sweep, $%
383: \varepsilon \gtrsim 1$ [see Eq.\ (\ref{PLZ})] the condition \TEXTsymbol{\vert%
384: }$J|\gg \Delta $ is sufficient, for the fast sweep, $\varepsilon \lesssim 1,$
385: a more stringent condition is required that follows from Eq.\ (16) of Ref.\ %
386: \onlinecite{garsch02prb}. The resulting combined condition for the
387: dynamically well-separated resonances thus would be
388: \begin{equation}
389: |J|\gg \left\{
390: \begin{array}{cc}
391: \Delta , & \varepsilon \gtrsim 1 \\
392: \sqrt{\hbar v}\sim \Delta /\varepsilon , & \varepsilon \lesssim 1.
393: \end{array}
394: \right. \label{SepResCond}
395: \end{equation}
396: This is, however, only an \emph{apriori} estimation considering two
397: resonances. For $N\gg 1$ there are many resonances, and the deviations from
398: the single-resonance picture can accumulate with the increase of $N.$ One
399: can do an \emph{aposteriori} estimation in the slow-sweep limit where the
400: result for the well-separated resonances $P\rightarrow 1-1/N$ for $%
401: \varepsilon \rightarrow \infty $ following from Eq.\ (\ref{PNF}) as well as
402: the mean-field result for $N\gg 1$ and non-separated resonances [the first
403: line of Eq.\ (\ref{PhxLims})] are available, in both cases $1-P\ll 1$. The
404: crossover between these results should take place at \TEXTsymbol{\vert}$%
405: J|\sim $ $\Delta N^{1/2},$ thus the first line of Eq.\ (\ref{SepResCond})
406: should be replaced by
407: \begin{equation}
408: |J|\gg \Delta N^{1/2},\qquad \varepsilon \gtrsim 1. \label{SepResCond2}
409: \end{equation}
410: The second line of Eq.\ (\ref{SepResCond}) could be also modified by $N$ but
411: it is difficult to derive an exact condition.
412:
413: A perturbative analysis of the stationary Schr\"{o}dinger equation for our
414: model shows that tunnel level splitting of the bare levels with different
415: values of $k$ has the form\cite{gar91jpa}
416: \begin{equation}
417: \delta E_{k_{1},k_{2}}\sim \Delta \left( \frac{\Delta }{J}\right)
418: ^{|k_{1}-k_{2}|-1}. \label{SplittDiffOrd}
419: \end{equation}
420: Thus for \TEXTsymbol{\vert}$J|\gg \Delta $ only the first-order or direct
421: transitions with $|k_{1}-k_{2}|=1$ should be taken into account whereas the
422: higher-order transitions are weak (see Fig.\ \ref{Fig-lzn-En}).
423:
424: Now it becomes clear that for the ferromagnetic coupling and well-separated
425: resonances only the strong transition between the initial level $k=0$ and
426: the level $k=1$ happens. That is, in another language, only one particle of $%
427: N$ has a chance to tunnel, and the tunneling probability for the whole
428: system is strongly reduced. For a quantitative analysis one can neglect all $%
429: c_{k}$ with $k>1$ in Eq.\ (\ref{SchrEqc}) that yields the system of
430: equations
431: \begin{eqnarray}
432: i\hbar \dot{c}_{0} &=&E_{0}c_{0}-\frac{\Delta }{2}\sqrt{N}c_{1} \nonumber \\
433: i\hbar \dot{c}_{1} &=&E_{1}c_{1}-\frac{\Delta }{2}\sqrt{N}c_{0}
434: \label{SchrEqcF}
435: \end{eqnarray}
436: that maps on the standard LZS problem with $\Delta \Rightarrow \Delta
437: _{N}\equiv \Delta \sqrt{N}$ and with the resonance at $W=2(N-1)J$ instead of
438: $W=0.$ The final-state probabilities for Eq.\ (\ref{SchrEqcF}) are according
439: to Eq.\ (\ref{PLZ})
440: \begin{equation}
441: p_{0}=P^{N}=e^{-N\varepsilon },\qquad p_{1}=1-P^{N}. \label{p0p1F}
442: \end{equation}
443: Then with the help of Eq.\ (\ref{PtFinal}) the one-particle staying
444: probability for $J>0$ can be cast into the form
445: \begin{equation}
446: P_{N}=1-\frac{1}{N}\left( 1-P^{N}\right) \label{PNF}
447: \end{equation}
448: and it varies between $1$ in the fast-sweep limit and $1-1/N$ in the
449: slow-sweep limit. Thus suppression of transitions by the ferromagnetic
450: coupling is extremely strong for a large number of particles $N.$ Expansion
451: of the transition probability of Eq.\ (\ref{PNF}) in the fast-sweep limit
452: reads
453: \begin{equation}
454: 1-P_{N}\cong \varepsilon -\frac{N}{2}\varepsilon ^{2}+O(\varepsilon
455: ^{3}),\qquad (\varepsilon \ll 1/N). \label{PNFfast}
456: \end{equation}
457: Note that the first term of this expansion is insensitive to the interaction
458: (in this context to the number of interacting particles) and is the same as
459: for the standard LZS problem, $1-P=1-e^{-\varepsilon }\cong \varepsilon
460: -\varepsilon ^{2}/2!+\varepsilon ^{3}/3!-\ldots $
461:
462: %TCIMACRO{
463: %\TeXButton{Fig-lzn-separatedresonances}{\begin{figure}[t]
464: %\unitlength1cm
465: %\begin{picture}(11,6)
466: %\centerline{\psfig{file=Fig-lzn-separatedresonances.eps,angle=-90,width=9cm}}
467: %\end{picture}
468: %\caption{ \label{Fig-lzn-separatedresonances}
469: %One-particle staying probability $P_N$ vs the sweep-rate parameter $\varepsilon$ for systems with FM and AFMF interaction in the case of well separated resonances.
470: %}
471: %\end{figure}%
472: %}}%
473: %BeginExpansion
474: \begin{figure}[t]
475: \unitlength1cm
476: \begin{picture}(11,6)
477: \centerline{\psfig{file=Fig-lzn-separatedresonances.eps,angle=-90,width=9cm}}
478: \end{picture}
479: \caption{ \label{Fig-lzn-separatedresonances}
480: One-particle staying probability $P_N$ vs the sweep-rate parameter $\varepsilon$ for systems with FM and AFMF interaction in the case of well separated resonances.
481: }
482: \end{figure}%
483: %
484: %EndExpansion
485:
486: The AFMF coupling $J<0$ favors transitions, as can be seen from Fig.\ \ref
487: {Fig-lzn-En}b. The original resonance is splitted by the interaction into $N$
488: resonances filling equidistantly the range $-2(N-1)J\leq W\leq 2(N-1)J.$ One
489: can see that in the limit of slow sweep the system remains on the lowest
490: adiabatic energy level thus $P_{N}\rightarrow 0,$ in contrast to the
491: ferromagnetic case. For well-separated resonances the problem described by
492: Eq.\ (\ref{SchrEqc}) splits into independent standard LZS problems for the
493: resonances between the levels $k$ and $k+1$ that are described by an
494: effective splitting $\Delta _{N,k}=\Delta l_{k,k+1}$ [cf. Eq.\ (\ref
495: {SchrEqcF})]. It is convenient to designate the probability to stay in state
496: $k$ after crossing with state $k-1$ but before crossing with state $k+1$
497: (see Fig.\ \ref{Fig-lzn-En}b) as $p_{k}($\textrm{mid}$)$. Then solutions of
498: the LZS problems for individual crossings along with conditions of the
499: probability conservation can be written as
500: \begin{eqnarray}
501: p_{k}(\infty ) &=&P^{(N-k)(k+1)}p_{k}(\text{\textrm{mid}}) \nonumber \\
502: p_{k}(\text{\textrm{mid}}) &=&p_{k-1}(\text{\textrm{mid}})-p_{k-1}(\infty )
503: \nonumber \\
504: p_{k-1}(\infty ) &=&P^{(N-k+1)k}p_{k-1}(\text{\textrm{mid}}), \label{EqsAF}
505: \end{eqnarray}
506: etc. These can be combined into the recurrence relation
507: \begin{eqnarray}
508: p_{k}(\infty ) &=&P^{(N-k)(k+1)}\left( P^{-(N-k+1)k}-1\right) p_{k-1}(\infty
509: ) \nonumber \\
510: &=&P^{N-2k}\left( 1-P^{(N-k+1)k}\right) p_{k-1}(\infty ) \label{RecurrAF}
511: \end{eqnarray}
512: that has to be iterated with the initial condition $p_{0}(\infty )=P^{N}.$
513: In the slow-sweep limit $\varepsilon \gg 1/N,$ one has $P^{N}\ll 1,$ thus
514: one can drop the factor $\left( 1-P^{(N-k+1)k}\right) \cong 1$ in the second
515: line of Eq.\ (\ref{RecurrAF}) that after iteration results in $p_{k}\cong
516: P^{(N-k)(k+1)}.$ In this case, in Eq.\ (\ref{PtFinal}) $p_{0}\cong
517: p_{N-1}\cong P^{N}$ are dominant for large $N$ whereas all other summands
518: are much smaller. This yields
519: \begin{equation}
520: P_{N}\cong \left( 1+\frac{1}{N}\right) P^{N}\ll 1,\qquad \left( \varepsilon
521: \gg 1/N\right) \label{PNAFslow}
522: \end{equation}
523: for the antiferromagnetic coupling. In the fast-sweep limit the solution of
524: Eq.\ (\ref{RecurrAF}) can be expanded in powers of $\varepsilon ,$ after
525: which Eq.\ (\ref{PtFinal}) yields for the transition probability
526: \begin{equation}
527: 1-P_{N}\cong \varepsilon +\left( \frac{3N}{2}-2\right) \varepsilon
528: ^{2}+O(\varepsilon ^{3}),\quad (\varepsilon \ll 1/N). \label{PNAFfast}
529: \end{equation}
530: Again, the first term of this expansion is the same as in the standard LZS
531: problem. For intermediate sweep rates one can iterate Eq.\ (\ref{RecurrAF})
532: numerically and substitute the solution for $p_{k}$ into Eq.\ (\ref{PtFinal}%
533: ). The resulting curves $P_{N}$ vs $\varepsilon $ are shown in Fig.\ \ref
534: {Fig-lzn-separatedresonances} along with those for the ferromagnetic
535: coupling, Eq.\ (\ref{PNF}).
536:
537: To conclude this section, we show the numerically computed dependences of $%
538: P_{N}$ on $J/\Delta $ for $N=3$ and different $\varepsilon $ in Fig.\ \ref
539: {Fig-lzn-PvsJ}. The case of well-separated resonances that was considered in
540: this section corresponds to the plateaus on the left and on the right sides
541: of the plot.
542:
543: %TCIMACRO{
544: %\TeXButton{Fig-lzn-PvsJ}{\begin{figure}[t]
545: %\unitlength1cm
546: %\begin{picture}(11,6)
547: %\centerline{\psfig{file=Fig-lzn-PvsJ.eps,angle=-90,width=9cm}}
548: %\end{picture}
549: %\caption{ \label{Fig-lzn-PvsJ}
550: %Numerically computed staying probability $P_{N}$ vs $J/\Delta $ for $N=3$ and different sweep rate parameters $\varepsilon $.
551: %The horizontal dashed lines on the left side of the plot are asymptotes corresponding to well-separated resonances.
552: %}
553: %\end{figure}%
554: %}}%
555: %BeginExpansion
556: \begin{figure}[t]
557: \unitlength1cm
558: \begin{picture}(11,6)
559: \centerline{\psfig{file=Fig-lzn-PvsJ.eps,angle=-90,width=9cm}}
560: \end{picture}
561: \caption{ \label{Fig-lzn-PvsJ}
562: Numerically computed staying probability $P_{N}$ vs $J/\Delta $ for $N=3$ and different sweep rate parameters $\varepsilon $.
563: The horizontal dashed lines on the left side of the plot are asymptotes corresponding to well-separated resonances.
564: }
565: \end{figure}%
566: %
567: %EndExpansion
568:
569: \section{Fast sweep and weak coupling}
570:
571: \label{Sec-fast-sweep}
572:
573: In the preceding section we have considered the strong-coupling limit
574: \TEXTsymbol{\vert}$J|\gg \max \left( \Delta N^{1/2},\Delta /\sqrt{%
575: \varepsilon }\right) $ in which individual resonances are well separated
576: from each other and one thus deals with successive standard LZS transitions.
577: The opposite limiting case is that of the weak coupling $J.$ This case can
578: be solved by a perturbative expansion in $J$ with the standard LZS solution
579: as the zeroth-order appriximation. The situation further simplifies in the
580: fast-sweep limit, where the transition probability is small and the
581: coefficients $c_{k}$ in Eq.\ (\ref{SchrEqc}) decrease with $k$ as powers of $%
582: \varepsilon \ll 1.$ In particular, $c_{0}$ contains terms of orders $%
583: \varepsilon ^{0},\varepsilon ^{1},$ and $\varepsilon ^{2},$ the coefficient $%
584: c_{1}$ contains $\varepsilon ^{1/2}$ and $\varepsilon ^{3/2},$ the
585: coefficient $c_{2}$ starts with $\varepsilon ,$ etc. Knowing these
586: contributions into $c_{0},$ $c_{1},$ and $c_{2}$ is sufficient to set up the
587: expansion of $P_{N}$ up to the first nontrivial order $\varepsilon ^{2}.$ A
588: straighforward but cumbersome calculation yields
589: \begin{eqnarray}
590: 1-P_{N} &\cong &\varepsilon -\left( \frac{1}{2}+\frac{4J_{0}}{\sqrt{2\pi
591: \hbar v}}\right) \varepsilon ^{2}+O(\varepsilon ^{3}) \nonumber \\
592: &=&\varepsilon -\frac{1}{2}\varepsilon ^{2}-\frac{4J_{0}}{\Delta }%
593: \varepsilon ^{5/2}+\ldots \label{PNSmallJfast}
594: \end{eqnarray}
595: where we have defined
596: \begin{equation}
597: J_{0}\equiv (N-1)J. \label{J0Def}
598: \end{equation}
599: One can see, again that ferromagnetic interaction suppresses transitions
600: whereas the AFMF interaction facilitates transitions, and that the effect of
601: interaction is increased by the number particles in the system. In fact,
602: however, in our model the interaction should scale, on physical grounds,
603: with the system's size, i.e., $J_{0}\equiv $ $(N-1)J$ should be size
604: independent. Note that in the fast-sweep limit one could do the expansion in
605: $\varepsilon $ for arbitrary $J.$ This leads, however, to cumbersome
606: multiple integrals. The strong-coupling limit for the fast sweep $%
607: \varepsilon \ll 1$ has been considered above, and the corresponding
608: counterparts of Eq.\ (\ref{PNSmallJfast}) for ferro-and antiferromagnetic
609: coupling are Eqs.\ (\ref{PNFfast}) and (\ref{PNAFfast}).
610:
611: It should be noted that although Eq.\ (\ref{PNSmallJfast}) is valid for $%
612: \varepsilon \ll 1,$ the method of its derivation above is only justified for
613: $\varepsilon \ll 1/N,$ i.e. for much faster sweep rates, if $N\gg 1.$
614: Indeed, in the course of the derivation the terms of orders $\left(
615: \varepsilon N\right) ^{1/2},$ $\varepsilon N,$ etc., appear that are
616: required to be small. Only at the very end the leading $N$ contributions
617: cancel each other that leads to Eq.\ (\ref{PNSmallJfast}) that fortunately
618: has a larger applicability range. In fact, for large systems with weak
619: interaction the most populated final states are
620: \begin{equation}
621: k_{\max }=N(1-P)\cong N\varepsilon , \label{kmaxDef}
622: \end{equation}
623: so that keeping only the states with $k=0,1,2$ above was not justified for $%
624: N\varepsilon \gtrsim 1.$ Eq.\ (\ref{kmaxDef}) can be easily derived for an
625: assembly of noninteracting tunneling species. To this end, we use the
626: coefficients $C_{k}$ of Eq.\ (\ref{SchrEq}) and represent them in the
627: factorized form
628: \begin{equation}
629: C_{k}=a_{-1}^{N-k}a_{1}^{k}, \label{CFactorized}
630: \end{equation}
631: where $a_{-1}$ and $a_{1}$ are the coefficients of the wave function of the
632: one-particle problem, $\psi =a_{-1}\psi _{-1}+a_{1}\psi _{1}.$ This yields
633: \begin{equation}
634: |C_{k}|^{2}=|a_{-1}|^{2\left( N-k\right) }|a_{1}|^{2k}=P^{N-k}(1-P)^{k}
635: \label{C2Factorized}
636: \end{equation}
637: that can be plugged into Eq.\ (\ref{PtDefk}) to give
638: \begin{equation}
639: P_{N}=\sum_{k=0}^{N-1}\frac{(N-1)!}{(N-1-k)!k!}P^{N-k}(1-P)^{k}.
640: \label{PNNoninteracting}
641: \end{equation}
642: The summand in this formula has its maximum at $k=k_{\max }$ given by Eq.\ (%
643: \ref{kmaxDef}) and it becomes a narrow Gaussian packet around $k_{\max }$\
644: for $N\gg 1.$
645:
646: A more rigorous method of handling the fast-sweep and weak coupling limits
647: uses the one-particle density matrix that is defined by
648: \begin{equation}
649: \rho _{-1,-1}=P_{N},\qquad \rho _{1,1}=1-\rho _{-1,-1}=1-P_{N}
650: \label{rhodiagDef}
651: \end{equation}
652: with $P_{N}$ of Eq.\ (\ref{PtDefk}) and
653: \begin{eqnarray}
654: \rho _{1,-1} &=&\sum_{m_{2},\ldots ,m_{N}=-1,1}C_{1,m_{2},\ldots
655: ,m_{N}}^{\ast }C_{-1,m_{2},\ldots ,m_{N}} \nonumber \\
656: &=&\sum_{k=0}^{N-1}\frac{(N-1)!}{(N-1-k)!k!}C_{k+1}^{\ast }C_{k} \nonumber
657: \\
658: \rho _{-1,1} &=&\rho _{1,-1}^{\ast }. \label{rhonondiagDef}
659: \end{eqnarray}
660: The density-matrix equation (DME) can be obtained by time differentiating of
661: $\rho _{mn}$ and employing Eq.\ (\ref{SchrEq}) that yields
662: \begin{eqnarray}
663: &&\hbar \dot{\rho}_{-1,-1}=\frac{i\Delta }{2}\left( \rho _{-1,1}-\rho
664: _{1,-1}\right) =\Delta \func{Im}\rho _{1,-1} \nonumber \\
665: &&\hbar \dot{\rho}_{1,-1}=-iW(t)\rho _{1,-1}+\frac{i\Delta }{2}\left( \rho
666: _{1,1}-\rho _{-1,-1}\right) \nonumber \\
667: &&{}+2iJ\sum_{k=0}^{N-1}\frac{(N-1)!}{(N-1-k)!k!}\left( N-2k-1\right)
668: C_{k}C_{k+1}^{\ast }. \label{DME}
669: \end{eqnarray}
670: In the absence of interaction, $J=0,$ the last term in $\dot{\rho}_{-1,1}$
671: disappears and one obtains a DME for one isolated particle that is
672: equivalent to the one-particle Schr\"{o}dinger equation. Note that solving
673: this one-particle equation in the fast-sweep limit requires $\varepsilon \ll
674: 1,$ in contrast to Eq.\ (\ref{SchrEq}) with $J=0$ that requires $\varepsilon
675: \ll 1/N.$
676:
677: For $J\neq 0$ Eqs.\ (\ref{DME}) do not form a closed system of equations. In
678: this case Eqs.\ (\ref{DME}) can only be useful if an approximation for the
679: interaction term be found. In particular, one can consider the interaction
680: term perturbatively in $J$ and use
681: \begin{equation}
682: \rho _{mn}=\rho _{mn}^{(0)}+\delta \rho _{mn} \label{rhoExpansion}
683: \end{equation}
684: where $\rho _{mn}^{(0)}$ is the density matrix without interaction that
685: satisfies
686: \begin{eqnarray}
687: &&\hbar \dot{\rho}_{-1,-1}^{(0)}=\Delta \func{Im}\rho _{1,-1}^{(0)}
688: \nonumber \\
689: &&\hbar \dot{\rho}_{1,-1}^{(\text{0)}}=-iW(t)\rho _{1,-1}^{(0)}+i\Delta
690: \left( \frac{1}{2}-\rho _{-1,-1}^{(0)}\right) \label{DME0}
691: \end{eqnarray}
692: and $\delta \rho _{mn}$ is the correction term. The latter satisfies
693: equations similar to Eqs.\ (\ref{DME}) in which, however, $%
694: C_{k}C_{k+1}^{\ast }$ is replaced by its noninteracting-particle expression
695: following from Eq.\ (\ref{CFactorized}),
696: \begin{eqnarray}
697: &&C_{k}C_{k+1}^{\ast }=a_{-1}^{N-k}a_{1}^{k}\left( a_{-1}^{\ast }\right)
698: ^{N-k-1}\left( a_{1}^{\ast }\right) ^{k+1} \nonumber \\
699: &&{}\qquad =|a_{-1}|^{2(N-k-1)}|a_{1}|^{2k}a_{-1}a_{1}^{\ast } \nonumber \\
700: &&\qquad =\left( \rho _{-1,-1}^{(0)}\right) ^{N-k-1}\left( 1-\rho
701: _{-1,-1}^{(0)}\right) ^{k}\rho _{-1,1}^{(0)}. \label{CkCkp1}
702: \end{eqnarray}
703: With the use of the latter, the sum over $k$ in Eqs.\ (\ref{DME}) can be
704: performed to yield
705: \begin{eqnarray}
706: &&\hbar \delta \dot{\rho}_{-1,-1}=\Delta \func{Im}\delta \rho _{1,-1}
707: \nonumber \\
708: &&\hbar \delta \dot{\rho}_{1,-1}=-iW(t)\delta \rho _{-1,1}-i\Delta \delta
709: \rho _{-1,-1} \nonumber \\
710: &&{}\qquad -2iJ_{0}\left( 1-2\rho _{-1,-1}^{(0)}\right) \rho _{1,-1}^{(0)}.
711: \label{DMEdelta}
712: \end{eqnarray}
713: Solving Eqs.\ (\ref{DME0}) and (\ref{DMEdelta}) perturbatively for $%
714: \varepsilon \ll 1$ results in Eq.\ (\ref{PNSmallJfast}).
715:
716: In accord with the remark at the beginning of this section, $J$-perturbative
717: Eqs.\ (\ref{DME0}) and (\ref{DMEdelta}) can be solved in terms of the
718: hypergeometric functions for any $\varepsilon $ (For the standard LZS
719: problem, $J=0,$ this was done by Zener. \cite{zen32}) It can be shown that
720: in the slow-sweep limit this solution simplifies to
721: \begin{equation}
722: \left. \frac{dP}{d\rho }\right| _{\rho =0}=c\varepsilon e^{-\varepsilon
723: },\qquad \varepsilon \gg 1,\qquad \rho \equiv \frac{2J_{0}}{\Delta },
724: \label{dPdrhoslowsweep}
725: \end{equation}
726: where $c$ is a number and the parameter $\rho $ should not be confused with
727: the density matrix.
728:
729: %TCIMACRO{
730: %\TeXButton{Fig-lzn-Pvsrho}{\begin{figure}[t]
731: %\unitlength1cm
732: %\begin{picture}(11,6)
733: %\centerline{\psfig{file=Fig-lzn-Pvsrho.eps,angle=-90,width=9cm}}
734: %\end{picture}
735: %\caption{ \label{Fig-lzn-Pvsrho}
736: %Numerically computed staying probability $P_{N}$ vs $\rho\equiv 2J_0/\Delta $ for $\varepsilon=1$ and different system sizes $N$, including the mean-field result for $N\to\infty$.}
737: %\end{figure}%
738: %}}%
739: %BeginExpansion
740: \begin{figure}[t]
741: \unitlength1cm
742: \begin{picture}(11,6)
743: \centerline{\psfig{file=Fig-lzn-Pvsrho.eps,angle=-90,width=9cm}}
744: \end{picture}
745: \caption{ \label{Fig-lzn-Pvsrho}
746: Numerically computed staying probability $P_{N}$ vs $\rho\equiv 2J_0/\Delta $ for $\varepsilon=1$ and different system sizes $N$, including the mean-field result for $N\to\infty$.}
747: \end{figure}%
748: %
749: %EndExpansion
750:
751: \section{The mean field approximation}
752:
753: \label{Sec-mfa}
754:
755: If each of tunneling particles interacts with all the other particles with a
756: coupling of the same sign, as is the case in our model for $N\gg 1,$ then
757: the mean-field approximation (MFA) should work well. The MFA employs a
758: one-particle density-matrix equation or a one-particle Schr\"{o}dinger
759: equation in which the interaction enters via the molecular field. For the
760: model of Eq.\ (\ref{Ham}) one has $W(t)\Rightarrow W_{\mathrm{eff}}(t)$
761: where
762: \begin{equation}
763: W_{\mathrm{eff}}(t)=W(t)+2J_{0}(1-2\rho _{-1,-1}) \label{WeffDef}
764: \end{equation}
765: in the DME of Eq.\ (\ref{DME0}), i.e.,
766: \begin{eqnarray}
767: &&\hbar \dot{\rho}_{-1,-1}=\Delta \func{Im}\rho _{1,-1} \nonumber \\
768: &&\hbar \dot{\rho}_{1,-1}=-iW_{\mathrm{eff}}(t)\rho _{1,-1}+i\Delta \left(
769: \frac{1}{2}-\rho _{-1,-1}\right) . \label{DMEMFA}
770: \end{eqnarray}
771: This is equivalent to the nonlinear Schr\"{o}dinger equation
772: \begin{eqnarray}
773: i\hbar \dot{C}_{-1} &=&\frac{1}{2}W_{\mathrm{eff}}(t)C_{-1}-\frac{\Delta }{2}%
774: C_{1} \nonumber \\
775: i\hbar \dot{C}_{1} &=&-\frac{1}{2}W_{\mathrm{eff}}(t)C_{1}-\frac{\Delta }{2}%
776: C_{-1} \label{NonlSchEq}
777: \end{eqnarray}
778: with $W_{\mathrm{eff}}(t)=W(t)+2J_{0}(1-2|C_{-1}|^{2}).$ It is interesting
779: to note that if one solves Eq.\ (\ref{DMEMFA}) perturbatively in $J$ \ using
780: Eq.\ (\ref{rhoExpansion}) one obtains Eqs.\ (\ref{DME0}) and (\ref{DMEdelta}%
781: ). This implies that the MFA works well in the weak-coupling limit in our
782: model (see Fig.\ \ref{Fig-lzn-Pvsrho}).
783:
784: In the case $N\gg 1$ one can assume that the state of the system is
785: described by a narrow packet in the $k$-space, as is indeed the case for
786: noninteracting particles, see Eq.\ (\ref{PNNoninteracting}). Then in Eq.\ (%
787: \ref{DME}) one can replace $\left( N-2k-1\right) \Rightarrow \left(
788: N-2k_{\max }-1\right) $ \ with $k_{\max }$ of Eq.\ (\ref{kmaxDef}), after
789: which the sum over $k$ converts to $\rho _{-1,1}$ according to the
790: definition in Eq.\ (\ref{rhonondiagDef}). This leads to Eq.\ (\ref{DMEMFA}).
791: One should, however, realize that this is no more than a heuristic
792: derivation since it was not proved that the state of a system \emph{with
793: interaction} is described by a narrow packet for $N\gg 1.$ On the other
794: hand, the narrow-packet assumption is quite plausible since the limit $N\gg
795: 1 $ corresponds to the quasiclassical limit $S\gg 1$ of the equivalent model
796: of Eq.\ (\ref{HamS}) and the states of quasiclassical systems should be
797: narrow packets. It is remarkable that the density-matrix equation within the
798: MFA, Eq.\ (\ref{DMEMFA}), can be interpreted as a purely classical equation
799: of motion for a spin vector, the Landau-Lifschitz equation (LLE). Indeed,
800: rewriting Eq.\ (\ref{rhonondiagDef}) in terms of $c_{k}$ with the help of
801: Eq.\ (\ref{ckDef}) and adding Eq.\ (\ref{PtMapping}) one obtains the mapping
802: relations
803: \begin{eqnarray}
804: &&\rho _{1,-1}=\frac{1}{N}\sum_{k=0}^{N}l_{k,k+1}c_{k+1}^{\ast }c_{k}=\frac{%
805: \langle S_{+}\rangle }{2S} \nonumber \\
806: &&\rho _{-1,-1}=\frac{1}{2}\left( 1-\frac{\langle S_{z}\rangle }{S}\right) .
807: \label{rhoSMapping}
808: \end{eqnarray}
809: Now one can see that the DME in the MFA, Eq.\ (\ref{DMEMFA}) is equivalent
810: to the LLE for the classical spin vector components
811: \begin{equation}
812: s_{z}=\frac{\langle S_{z}\rangle }{S},\qquad s_{+}=\frac{\langle
813: S_{+}\rangle }{S} \label{sSavrIdent}
814: \end{equation}
815: that reads
816: \begin{eqnarray}
817: &&\hbar \dot{s}_{z}=-\Delta s_{y} \nonumber \\
818: &&\hbar \dot{s}_{+}=-i\left[ W(t)+4J(S-1/2)s_{z}\right] s_{+}+i\Delta s_{z}.
819: \label{LLEzp}
820: \end{eqnarray}
821: The vector form of this LLE is
822: \begin{equation}
823: \mathbf{\dot{s}}=\gamma \left[ \mathbf{s\times H}_{\mathrm{eff}}\right]
824: ,\qquad \mathbf{H}_{\mathrm{eff}}=-\frac{\partial \mathcal{H}_{\mathrm{eff}}%
825: }{\partial \mathbf{s}}, \label{LLE}
826: \end{equation}
827: where $\gamma =1/\hbar $ and the effective classical energy is given by
828: \begin{equation}
829: \mathcal{H}_{\mathrm{eff}}=-H_{z}(t)s_{z}-H_{x}s_{x}-D_{\mathrm{cl}}s_{z}^{2}
830: \label{HamClass}
831: \end{equation}
832: with
833: \begin{eqnarray}
834: &&H_{z}(t)=W(t),\qquad H_{x}=\Delta ,\qquad \nonumber \\
835: &&D_{\mathrm{cl}}=D(S-1/2)=J(N-1)\equiv J_{0} \label{ParIdentClass}
836: \end{eqnarray}
837: [cf. Eqs.\ (\ref{HamS}) and (\ref{ParIdent})]. The one-particle staying
838: probability $P_{N}$ is given by the formula
839: \begin{equation}
840: P_{N}(t)=\frac{1}{2}\left[ 1-s_{z}(t)\right] \label{PtMappingClass}
841: \end{equation}
842: that is analogous to Eq.\ (\ref{PtMapping}).
843:
844: %TCIMACRO{
845: %\TeXButton{Fig-lzn-PMFA-AF}{\begin{figure}[t]
846: %\unitlength1cm
847: %\begin{picture}(11,6)
848: %\centerline{\psfig{file=Fig-lzn-PMFA-AF.eps,angle=-90,width=9cm}}
849: %\end{picture}
850: %\caption{ \label{Fig-lzn-PMFA-AF}
851: %The mean-field solution for the one-particle staying probability $P_N$ for the AFMF coupling, $J_0\equiv (N-1)J<0$.
852: %}
853: %\end{figure}%
854: %}}%
855: %BeginExpansion
856: \begin{figure}[t]
857: \unitlength1cm
858: \begin{picture}(11,6)
859: \centerline{\psfig{file=Fig-lzn-PMFA-AF.eps,angle=-90,width=9cm}}
860: \end{picture}
861: \caption{ \label{Fig-lzn-PMFA-AF}
862: The mean-field solution for the one-particle staying probability $P_N$ for the AFMF coupling, $J_0\equiv (N-1)J<0$.
863: }
864: \end{figure}%
865: %
866: %EndExpansion
867:
868: %TCIMACRO{
869: %\TeXButton{Fig-lzn-PWeffvst-AF-MFA-j10m}{\begin{figure}[t]
870: %\unitlength1cm
871: %\begin{picture}(11,6)
872: %\centerline{\psfig{file=Fig-lzn-Pvst-AF-MFA-j10m.eps,angle=-90,width=9cm}}
873: %\end{picture}
874: %\begin{picture}(11,6)
875: %\centerline{\psfig{file=Fig-lzn-Weffvst-AF-MFA-j10m.eps,angle=-90,width=9cm}}
876: %\end{picture}
877: %\caption{ \label{Fig-lzn-PWeffvst-AF-MFA-j10m}
878: %$a$ -- Time dependence of the staying probability $P_N$ for the AFMF coupling
879: %$J_0\equiv (N-1)J=-10$ for the first three complete conversion cases,
880: %$P_N(\infty)=0$ (see Fig.\ \protect\ref{Fig-lzn-PMFA-AF}).
881: %$b$ -- Same for the effective sweep $W_{\rm eff}$ of Eq.\ (\protect\ref{WeffDef}).
882: %}
883: %\end{figure}%
884: %}}%
885: %BeginExpansion
886: \begin{figure}[t]
887: \unitlength1cm
888: \begin{picture}(11,6)
889: \centerline{\psfig{file=Fig-lzn-Pvst-AF-MFA-j10m.eps,angle=-90,width=9cm}}
890: \end{picture}
891: \begin{picture}(11,6)
892: \centerline{\psfig{file=Fig-lzn-Weffvst-AF-MFA-j10m.eps,angle=-90,width=9cm}}
893: \end{picture}
894: \caption{ \label{Fig-lzn-PWeffvst-AF-MFA-j10m}
895: $a$ -- Time dependence of the staying probability $P_N$ for the AFMF coupling
896: $J_0\equiv (N-1)J=-10$ for the first three complete conversion cases,
897: $P_N(\infty)=0$ (see Fig.\ \protect\ref{Fig-lzn-PMFA-AF}).
898: $b$ -- Same for the effective sweep $W_{\rm eff}$ of Eq.\ (\protect\ref{WeffDef}).
899: }
900: \end{figure}%
901: %
902: %EndExpansion
903:
904: Let us now analyze the mean-field solution of the LZS problem based on Eqs.\
905: (\ref{DMEMFA}) or (\ref{LLE}) and compare it with the exact solution of Eq.\
906: (\ref{SchrEqc}). We start with the AFMF coupling, for convenience. Fig.\ \ref
907: {Fig-lzn-PMFA-AF} shows the dependence of the one-particle staying
908: probability $P_{N}$ on the sweep-rate parameter $\varepsilon $ for different
909: coupling strengths $J_{0}.$ The curves on the plot are qualitatively similar
910: to those in the case of well-separated resonances, Fig.\ \ref
911: {Fig-lzn-separatedresonances} and they show that AFMF interaction boosts
912: transitions, especially for slow sweep, $\varepsilon \gg 1.$ The difference
913: is that in the MFA only the combined parameter $J_{0}=(N-1)J$ enters,
914: whereas for well-separated resonances the results depend on $N$ only and not
915: on $J.$ Another difference is that for strong enough interaction the
916: mean-field solution for $P_{N}$ oscillates and vanishes at some values of $%
917: \varepsilon .$ Oscillations of this kind have been found in the solution of
918: the LZS problem with nonlinear sweep.\cite{ternak97,garsch02prb} As was
919: explained in Ref.\ \onlinecite{garsch02prb}, oscillations arise when the
920: sweep slows down in the vicinity of the resonance so that (in the extreme
921: case) the particle oscillates between the two quantum states and the
922: resulting staying probability depends on the phase of this oscillation at
923: the end of the stay near the resonance. This is also the case for the model
924: with AFMF interaction in the MFA.Tunneling of particles creates a molecular
925: field that changes in the direction opposite to the sweep so that the
926: resulting effective sweep$\ W_{\mathrm{eff}}(t)$ of Eq.\ (\ref{WeffDef}) can
927: even become nonmonotonic (in this case the analytical method of Ref.\ %
928: \onlinecite{hamraemiysai00} based on the linearization of the problem near
929: the resonance and introducing an effective sweep rate breaks down). The MFA
930: solution becomes time symmetric in the case of complete conversion, $%
931: P_{N}(\infty )=0$ that for $J_{0}/\Delta =-10$ is realized for $\varepsilon
932: =0.1137983$, 0.13207, 0.1741, 0.2233, etc. The corresponding time
933: dependences of $P_{N}(t)$ and $W_{\mathrm{eff}}(t)$ are shown in Fig.\ (\ref
934: {Fig-lzn-PWeffvst-AF-MFA-j10m}). Although they resemble the complete
935: conversion solutions found in Ref.\ \onlinecite{garsch02}, there are no
936: apparent analytical solutions for this model. The difficulty of the present
937: mean-field model is that $W_{\mathrm{eff}}(t)$ is not known from the
938: beginning but is a solution of a self-consistent nonlinear problem.
939:
940: Comparison of the exact and the MFA solutions for the one-particle staying
941: probability $P_{N}$ for the antiferromagnetic coupling $J_{0}=-10$ is shown
942: in Fig.\ \ref{Fig-lzn-P-AF-j10}. One can see that the exact quantum solution
943: converges to the classical mean-field solution for $N\rightarrow \infty $,
944: although in the vicinity of the first complete-conversion point convergence
945: is extremely slow.
946:
947: %TCIMACRO{
948: %\TeXButton{Fig-lzn-P-AF-j10}{\begin{figure}[t]
949: %\unitlength1cm
950: %\begin{picture}(11,6)
951: %\centerline{\psfig{file=Fig-lzn-P-AF-j10.eps,angle=-90,width=9cm}}
952: %\end{picture}
953: %\caption{ \label{Fig-lzn-P-AF-j10}
954: %Comparison of the exact quantum and the MFA solutions for the one-particle staying probability $P_N$ for the AFMF coupling
955: %showing partially slow convergence to the classical mean-field limit for $N\to\infty$.
956: %}
957: %\end{figure}%
958: %}}%
959: %BeginExpansion
960: \begin{figure}[t]
961: \unitlength1cm
962: \begin{picture}(11,6)
963: \centerline{\psfig{file=Fig-lzn-P-AF-j10.eps,angle=-90,width=9cm}}
964: \end{picture}
965: \caption{ \label{Fig-lzn-P-AF-j10}
966: Comparison of the exact quantum and the MFA solutions for the one-particle staying probability $P_N$ for the AFMF coupling
967: showing partially slow convergence to the classical mean-field limit for $N\to\infty$.
968: }
969: \end{figure}%
970: %
971: %EndExpansion
972:
973: %TCIMACRO{
974: %\TeXButton{Fig-lzn-PMFA-F}{\begin{figure}[t]
975: %\unitlength1cm
976: %\begin{picture}(11,6)
977: %\centerline{\psfig{file=Fig-lzn-PMFA-F.eps,angle=-90,width=9cm}}
978: %\end{picture}
979: %\caption{ \label{Fig-lzn-PMFA-F}
980: %The mean-field solution for the one-particle staying probability $P_N$ for the ferromagnetic coupling, $J_0\equiv (N-1)J>0$.
981: %The dashed line is the large-$\varepsilon$ asymptote of Eq.\ (\protect\ref{PN0alSmall}).
982: %}
983: %\end{figure}%
984: %}}%
985: %BeginExpansion
986: \begin{figure}[t]
987: \unitlength1cm
988: \begin{picture}(11,6)
989: \centerline{\psfig{file=Fig-lzn-PMFA-F.eps,angle=-90,width=9cm}}
990: \end{picture}
991: \caption{ \label{Fig-lzn-PMFA-F}
992: The mean-field solution for the one-particle staying probability $P_N$ for the ferromagnetic coupling, $J_0\equiv (N-1)J>0$.
993: The dashed line is the large-$\varepsilon$ asymptote of Eq.\ (\protect\ref{PN0alSmall}).
994: }
995: \end{figure}%
996: %
997: %EndExpansion
998:
999: %TCIMACRO{
1000: %\TeXButton{Fig-lzn-P-F-j3}{\begin{figure}[t]
1001: %\unitlength1cm
1002: %\begin{picture}(11,6)
1003: %\centerline{\psfig{file=Fig-lzn-P-F-j3.eps,angle=-90,width=9cm}}
1004: %\end{picture}
1005: %\caption{ \label{Fig-lzn-P-F-j3}
1006: %Comparison of the exact quantum and the MFA solutions for the one-particle staying probability $P_N$ for the ferromagnetic coupling.
1007: %}
1008: %\end{figure}%
1009: %}}%
1010: %BeginExpansion
1011: \begin{figure}[t]
1012: \unitlength1cm
1013: \begin{picture}(11,6)
1014: \centerline{\psfig{file=Fig-lzn-P-F-j3.eps,angle=-90,width=9cm}}
1015: \end{picture}
1016: \caption{ \label{Fig-lzn-P-F-j3}
1017: Comparison of the exact quantum and the MFA solutions for the one-particle staying probability $P_N$ for the ferromagnetic coupling.
1018: }
1019: \end{figure}%
1020: %
1021: %EndExpansion
1022:
1023: Let us now turn to the ferromagnetic interaction. The mean-field solution
1024: for different interaction strengths is shown in Fig.\ \ref{Fig-lzn-PMFA-F}.
1025: Again, qualitatively it is similar to the solution in the limit of
1026: well-separated resonances, Eq.\ (\ref{PNF}) that is shown in Fig.\ \ref
1027: {Fig-lzn-separatedresonances}. The difference is the same as for the AFMF
1028: coupling: The MFA solution depends on $J_{0}\equiv (N-1)J$ whereas Eq.\ (\ref
1029: {PNF}) depends on $N$ only. Another difference is that in the slow-sweep
1030: limit the mean-field curve tends to a constant (for $J_{0}/\Delta >1/2)$
1031: whereas all quantum curves tend to zero. The latter are in fact combinations
1032: of several LZS exponentials corresponding to transitions of different orders
1033: shown in Fig.\ \ref{Fig-lzn-En}a by dotted arrows (see also Fig.\ \ref
1034: {Fig-lzn-P3Ft}). One can see, in particular, that the curve $N=2$ consist of
1035: two different exponentials, the curve $N=3$ consist of three different
1036: exponentials, etc. With increasing of $N,$ however, the coupling $%
1037: J=J_{0}/(N-1)$ decreases and transitions become not well separated. In this
1038: case dependence $P_{N}$ of $\varepsilon $ becomes a long-tale curve.
1039:
1040: \section{Slow sweep in the MFA}
1041:
1042: \label{Sec-slow-MFA}
1043:
1044: Here we analytically consider the slow-sweep limit within the mean-field
1045: approximation, to get more insights into the mechanism that leads to
1046: suppressing transitions in the case of FM interactions (see Fig.\ \ref
1047: {Fig-lzn-PMFA-F}) as well as into that of facilitating transitions and
1048: probabillity oscillations in the case of the antiferromagnetic frustrating
1049: interactions (see Fig.\ \ref{Fig-lzn-PMFA-AF}).
1050:
1051: \subsection{Adiabatic case: Basic equations}
1052:
1053: For the one-particle LZS problem, a convenient method of treating the
1054: slow-sweep limit is that using the adiabatic basis.\cite{garsch02prb} Its
1055: advantage is that the probability to stay on the lowest adiabatic level is
1056: at any time close to 1, so that one can make an appropriate approximation
1057: after which the solution is obtained as a quadrature, including the case of
1058: nonlinear sweep. Direct extention of this method for the present model
1059: within the MFA is cumbersome, however, owing to the complicated
1060: self-consistent nature of the adiabatic energy levels for a \emph{nonlinear}
1061: Schr\"{o}dinger equation or \emph{nonlinear} density-matrix equation, Eq.\ (%
1062: \ref{DME}). Fortunately, an alternative description based on the LLE, Eq.\ (%
1063: \ref{LLE}), supports a physically transparent extension of the method. The
1064: adiabatic basis for a Schr\"{o}dinger equation corresponds to the adiabatic
1065: coordinate system for the classical problem of Eq.\ (\ref{LLE}). The $%
1066: z^{\prime }$ axis of this coordinate system is oriented in the direction
1067: minimizing the classical energy $\mathcal{H}_{\mathrm{eff}}$ of Eq.\ (\ref
1068: {HamClass}) at any fixed time $t.$ It belongs to the $x$-$z$ plane and it
1069: makes the angle $\theta (t)$ with the $z$ axis \ that satisfies the equation
1070: \begin{equation}
1071: h_{z}(t)\sin \theta +\nu \sin \theta \cos \theta -h_{x}\cos \theta =0
1072: \label{EMinEq}
1073: \end{equation}
1074: following from $\partial \mathcal{H}_{\mathrm{eff}}/\partial \theta =0.$
1075: Here $\nu =D_{\mathrm{cl}}/|D_{\mathrm{cl}}|=J/|J|$ and
1076: \begin{equation}
1077: h_{x}\equiv \frac{H_{x}}{2|D_{\mathrm{cl}}|}=\frac{\Delta }{2|J_{0}|},\qquad
1078: h_{z}(t)\equiv \frac{H_{z}(t)}{2|D_{\mathrm{cl}}|}=\frac{W(t)}{2|J_{0}|}.
1079: \label{hxhzDef}
1080: \end{equation}
1081: Eq.\ (\ref{EMinEq}) is well known in the theory of magnetism. In the AFMF
1082: case $\nu <0$, its solution is unique. For the ferromagnetic interaction $%
1083: \nu >0,$ Eq.\ (\ref{EMinEq}) has two solutions corresponding to the minimum
1084: and to the maximum of $\mathcal{H}_{\mathrm{eff}}$ in the range of reduced
1085: fields $h_{x}$ and $h_{z}$ outside the Stoner-Wohlfarth astroid \cite{stowoh}
1086: (see Fig.\ \ref{Fig-lzn-P-F-Astroid})
1087: \begin{equation}
1088: h_{x}^{2/3}+h_{z}^{2/3}=1. \label{Astroid}
1089: \end{equation}
1090: For $h_{x}$ and $h_{z}$ inside the astroid, Eq.\ (\ref{EMinEq}) has four
1091: solutions corresponding to the stable and metastable minima as well as to
1092: the saddle point and to the maximum of $\mathcal{H}_{\mathrm{eff}}.$ The
1093: adiabatic solution corresponds to the minimum of the energy that changes as $%
1094: h_{z}(t)$ is sweeped from $-\infty $ to $\infty $. For $h_{x}\geq 1$ the
1095: system follows this adiabatic solution in the whole range of $h_{z}.$ In
1096: this case dependence of the effective sweep $W_{\mathrm{eff}}$ of Eq.\ (\ref
1097: {WeffDef}) or, here, $W_{\mathrm{eff}}=H_{z}+2D_{\mathrm{cl}}\cos \theta ,$
1098: on the energy sweep $W$ is shown in Fig.\ \ref{Fig-lzn-P-F-Astroid-Weff}.
1099: One can see that for $J_{0}/\Delta =0.5$ this dependence becomes
1100: nonanalytic. This corresponds to the horizontal line in Fig.\ \ref
1101: {Fig-lzn-P-F-Astroid} that touches the upper corner of the astroid. In
1102: contrast, for $h_{x}<1$ the adiabatic solution exists only until the
1103: crossing of the right branch of the astroid. After that the spin rotates
1104: away from the disappeared metastable state and its behavior becomes
1105: nonadiabatic. The latter case will be considered later while at first we
1106: concentrate on the adiabatic situation.
1107:
1108: %TCIMACRO{
1109: %\TeXButton{Fig-lzn-P-F-Astroid}{\begin{figure}[t]
1110: %\unitlength1cm
1111: %\begin{picture}(11,4.5)
1112: %\centerline{\psfig{file=Fig-lzn-P-F-Astroid.eps,angle=-90,width=8cm}}
1113: %\end{picture}
1114: %\caption{ \label{Fig-lzn-P-F-Astroid}
1115: %Comparison of the exact quantum and the MFA solutions for the one-particle staying probability $P_N$ for the ferromagnetic coupling.
1116: %}
1117: %\end{figure}%
1118: %}}%
1119: %BeginExpansion
1120: \begin{figure}[t]
1121: \unitlength1cm
1122: \begin{picture}(11,4.5)
1123: \centerline{\psfig{file=Fig-lzn-P-F-Astroid.eps,angle=-90,width=8cm}}
1124: \end{picture}
1125: \caption{ \label{Fig-lzn-P-F-Astroid}
1126: Comparison of the exact quantum and the MFA solutions for the one-particle staying probability $P_N$ for the ferromagnetic coupling.
1127: }
1128: \end{figure}%
1129: %
1130: %EndExpansion
1131:
1132: %TCIMACRO{
1133: %\TeXButton{Fig-lzn-P-F-Astroid-Weff}{\begin{figure}[t]
1134: %\unitlength1cm
1135: %\begin{picture}(11,6)
1136: %\centerline{\psfig{file=Fig-lzn-P-F-Astroid-Weff.eps,angle=-90,width=9cm}}
1137: %\end{picture}
1138: %\caption{ \label{Fig-lzn-P-F-Astroid-Weff}
1139: %Effective sweep in the adiabatic case for the ferromagnetic coupling of different strengths.
1140: %}
1141: %\end{figure}%
1142: %}}%
1143: %BeginExpansion
1144: \begin{figure}[t]
1145: \unitlength1cm
1146: \begin{picture}(11,6)
1147: \centerline{\psfig{file=Fig-lzn-P-F-Astroid-Weff.eps,angle=-90,width=9cm}}
1148: \end{picture}
1149: \caption{ \label{Fig-lzn-P-F-Astroid-Weff}
1150: Effective sweep in the adiabatic case for the ferromagnetic coupling of different strengths.
1151: }
1152: \end{figure}%
1153: %
1154: %EndExpansion
1155:
1156: The Landau-Lifshitz equation, Eq.\ (\ref{LLE}) can be rewritten in the
1157: rotating adiabatic coordinate system as follows
1158: \begin{equation}
1159: \mathbf{\dot{s}}^{\prime }=\left[ \mathbf{s}^{\prime }\mathbf{\times }\left(
1160: \gamma \mathbf{H}_{\mathrm{eff}}^{\prime }+\mathbf{\Omega }\right) \right]
1161: ,\qquad \mathbf{\Omega =}\dot{\theta}\mathbf{e}_{y}, \label{LLErotating}
1162: \end{equation}
1163: where $\dot{\theta}$ is time derivative of the appropriate solution of Eq.\ (%
1164: \ref{EMinEq}). The reduced effective field
1165: \begin{equation}
1166: \mathbf{h}_{\mathrm{eff}}\equiv \frac{\mathbf{H}_{\mathrm{eff}}}{2|D_{%
1167: \mathrm{cl}}|}=\left( h_{z}+\nu s_{z}\right) \mathbf{e}_{z}+h_{x}\mathbf{e}%
1168: _{x}, \label{heff}
1169: \end{equation}
1170: in the adiabatic frame is calculated as follows
1171: \begin{eqnarray}
1172: \mathbf{h}_{\mathrm{eff}}^{\prime } &=&\left( h_{x}\cos \theta -h_{\mathrm{%
1173: eff,}z}\sin \theta \right) \mathbf{e}_{x^{\prime }} \nonumber \\
1174: &&+\left( h_{x}\sin \theta +h_{\mathrm{eff,}z}\cos \theta \right) \mathbf{e}%
1175: _{z^{\prime }} \nonumber \\
1176: h_{\mathrm{eff,}z} &=&h_{z}+\nu \left( -s_{x^{\prime }}\sin \theta
1177: +s_{z^{\prime }}\cos \theta \right) . \label{heffTrans}
1178: \end{eqnarray}
1179: With the help of Eq.\ (\ref{EMinEq}) components of $\mathbf{h}_{\mathrm{eff}%
1180: }^{\prime }$ simplify to
1181: \begin{eqnarray}
1182: h_{\mathrm{eff,}x^{\prime }} &=&\left( 1-s_{z^{\prime }}+s_{x^{\prime }}\tan
1183: \theta \right) \nu \sin \theta \cos \theta \nonumber \\
1184: h_{\mathrm{eff,}z^{\prime }} &=&h_{x}/\sin \theta -s_{x^{\prime }}\nu \sin
1185: \theta \cos \theta \nonumber \\
1186: &&+\left( 1-s_{z^{\prime }}\right) \left( h_{z}-h_{x}\cot \theta \right)
1187: \cos \theta . \label{heffTransFinal}
1188: \end{eqnarray}
1189: Note that if the spin is directed along the $z^{\prime }$ axis, i.e., $%
1190: s_{x^{\prime }}=s_{y}=0$ and s$_{z^{\prime }}=1,$ one has $h_{\mathrm{eff,}%
1191: x^{\prime }}=0,$ that is consistent with the choice of the adiabatic
1192: coordinate system. Introducing the dimensionless sweep variables
1193: \begin{equation}
1194: u\equiv \frac{W(t)}{\Delta }=\frac{vt}{\Delta }=\frac{h_{z}}{h_{x}},\qquad
1195: \tilde{\varepsilon}\equiv \frac{\Delta ^{2}}{\hbar v} \label{uDef}
1196: \end{equation}
1197: [$\tilde{\varepsilon}$ should not be confused with $\varepsilon =(\pi /2)%
1198: \tilde{\varepsilon}$ of Eq.\ (\ref{PLZ})] one can rewrite Eq.\ (\ref
1199: {LLErotating}) in the form
1200: \begin{equation}
1201: \frac{d\mathbf{s}^{\prime }}{du}=\left[ \mathbf{s}^{\prime }\mathbf{\times }%
1202: \left( \tilde{\varepsilon}\frac{\mathbf{h}_{\mathrm{eff}}^{\prime }}{h_{x}}+%
1203: \frac{d\theta }{du}\mathbf{e}_{y}\right) \right] . \label{epstilDef}
1204: \end{equation}
1205:
1206: In the slow-sweep limit $\varepsilon \gg 1$ the solution of Eq.\ (\ref
1207: {epstilDef}) is close to the adiabatic solution $s_{x^{\prime }}=s_{y}=0$
1208: and s$_{z^{\prime }}=1.$ Hence one can linearize this equation near the
1209: adiabatic solution:
1210: \begin{eqnarray}
1211: s_{z^{\prime }} &\Rightarrow &1,\qquad h_{\mathrm{eff,}z^{\prime
1212: }}\Rightarrow h_{x}/\sin \theta \nonumber \\
1213: h_{\mathrm{eff,}x^{\prime }} &\Rightarrow &s_{x^{\prime }}\nu \sin ^{2}\theta
1214: \label{Linearization}
1215: \end{eqnarray}
1216: that results in
1217: \begin{eqnarray}
1218: \frac{ds_{x^{\prime }}}{du} &=&\frac{\tilde{\varepsilon}}{\sin \theta }s_{y}-%
1219: \frac{d\theta }{du} \nonumber \\
1220: \frac{ds_{y}}{du} &=&-\frac{\zeta \tilde{\varepsilon}}{\sin \theta }%
1221: s_{x^{\prime }}, \label{LLElinearized}
1222: \end{eqnarray}
1223: where
1224: \begin{equation}
1225: \zeta \equiv 1-\rho \sin ^{3}\theta ,\qquad \rho \equiv 2J_{0}/\Delta .
1226: \label{arhoDef}
1227: \end{equation}
1228: The factor $\zeta $ in the second of Eqs.\ (\ref{LLElinearized}) makes spin
1229: precession elliptic for any nonzero coupling $J_{0},$ especially in the
1230: extreme case $\rho =1$. This is an important difference from the model of
1231: one tunneling particle with a nonlinear sweep, \cite{garsch02prb} where
1232: precession remains always circular. It is convenient to introduce
1233: \begin{equation}
1234: \tilde{c}_{+}\equiv -\frac{1}{2}\left( \zeta ^{1/4}s_{x^{\prime }}+i\zeta
1235: ^{-1/4}s_{y}\right) ,\qquad \overline{\Omega }(u)\equiv \frac{\sqrt{\zeta }}{%
1236: \sin \theta } \label{cplusOmegaDef}
1237: \end{equation}
1238: and rewrite Eqs.\ (\ref{LLElinearized}) in the form
1239: \begin{equation}
1240: \frac{d\tilde{c}_{+}}{du}=-i\tilde{\varepsilon}\overline{\Omega }(u)\tilde{c}%
1241: _{+}+\frac{1}{2}\frac{d\theta }{du}\zeta ^{1/4}. \label{cplusEq}
1242: \end{equation}
1243: This differential equation can be easily solved with the initial condition $%
1244: \tilde{c}_{+}(-\infty )=0,$ and the staying probability $P_{N}$ can be found
1245: from Eq.\ (\ref{PtMappingClass}). Keeping in mind that at $t=\infty $ both
1246: coordinate systems coincide whereby $\theta =0$ and $a=1$, one can rewrite $%
1247: P_{N}\equiv P_{N}(\infty )$ within our linearized theory as
1248: \begin{equation}
1249: P_{N}\cong \frac{1}{4}\left[ s_{x^{\prime }}^{2}(\infty )+s_{y}^{2}(\infty )%
1250: \right] =\left| \tilde{c}_{+}(\infty )\right| ^{2}. \label{PNLinearized}
1251: \end{equation}
1252: The final general expression for $P_{N}$ reads
1253: \begin{equation}
1254: P_{N}\cong \left| \frac{1}{2}\int_{-\infty }^{\infty }du\frac{d\theta }{du}%
1255: \zeta ^{1/4}\exp \left[ i\tilde{\varepsilon}\Phi (u)\right] \right| ^{2},
1256: \label{PNfinalgeneral}
1257: \end{equation}
1258: where
1259: \begin{equation}
1260: \Phi (u)\equiv \int_{0}^{u}du^{\prime }\,\overline{\Omega }(u^{\prime }).
1261: \label{PhiDef}
1262: \end{equation}
1263:
1264: In the absence of interaction $J=0,$ our solution recovers that for the
1265: standard LZS problem in the slow-sweep limit up to the prefactor in front of
1266: a small exponential. Indeed, in this case the solution of Eq.\ (\ref{EMinEq}%
1267: ) is
1268: \begin{equation}
1269: \cos \theta =\frac{h_{z}}{\sqrt{h_{x}^{2}+h_{z}^{2}}}=\frac{u}{\sqrt{1+u^{2}}%
1270: }, \label{CoSiJ0}
1271: \end{equation}
1272: in Eq.\ (\ref{arhoDef}) one has $\rho =0$ and $\zeta =1,$ hence
1273: \begin{equation}
1274: \overline{\Omega }(u)=\sqrt{1+u^{2}},\qquad \frac{d\theta }{du}=-\frac{1}{%
1275: 1+u^{2}}=-\frac{1}{\overline{\Omega }^{2}(u)}. \label{OmegaDerthetaJ0}
1276: \end{equation}
1277: One can see that in this case Eq.\ (\ref{cplusEq}) coincides with second of
1278: Eqs.\ (24) of Ref.\ \onlinecite{garsch02prb} and Eq.\ (\ref{PNfinalgeneral})
1279: coincides with Eq.\ (26) of Ref.\ \onlinecite{garsch02prb}, with $w^{\prime
1280: }(u)=1$ and $\tilde{c}_{-}\Rightarrow 1.$ The latter is exactly the
1281: approximation proposed in Ref.\ \onlinecite{garsch02prb} that for the
1282: standard LZS problem reproduces the well-known result of Eq.\ (\ref{PLZ})
1283: for $\varepsilon \gg 1$, however with a wrong prefactor: $P\approx (\pi
1284: /3)^{2}e^{-\varepsilon }.$ In Ref.\ \onlinecite{garsch02prb} we have shown
1285: how to correct the prefactor by accurately calculating $\tilde{c}_{-}.$
1286:
1287: \subsection{Adiabatic case: The results }
1288:
1289: In the general case $J\neq 0$ Eq.\ (\ref{EMinEq}) does not yield an
1290: analytical solution for $\theta (u).$ Fortunately, instead of $u$ one can
1291: use $x\equiv \cos \theta $ as the integration variable on the interval $%
1292: -1\leq x\leq 1$ and with the help of Eq.\ (\ref{EMinEq}) express $%
1293: u=h_{z}/h_{x}$ as a function of $x.$ Even better is then to parametrize $x$
1294: as $x=w/\sqrt{1+w^{2}},$ analogously to Eq.\ (\ref{CoSiJ0}). This brings
1295: Eq.\ (\ref{PNfinalgeneral}) into the explicit form
1296: \begin{equation}
1297: P_{N}\cong \left| \frac{1}{2}\int_{-\infty }^{\infty }\frac{dw}{1+w^{2}}%
1298: \left[ 1-\frac{\rho }{(1+w^{2})^{3/2}}\right] ^{1/4}e^{i\tilde{\varepsilon}%
1299: \Phi (w)}\right| ^{2}, \label{PNx}
1300: \end{equation}
1301: where
1302: \begin{equation}
1303: \Phi (w)=\int_{0}^{w}dw\sqrt{1+w^{2}}\left[ 1-\frac{\rho }{(1+w^{2})^{3/2}}%
1304: \right] ^{3/2}. \label{Phix}
1305: \end{equation}
1306: In the absence of interaction, $\rho =0,$ one has $w=u$ and the formulas
1307: above describe the standard LZS effect. For $\varepsilon \gg 1,$ the value
1308: of $P_{N}$ is exponentially small and defined by the singularities of the
1309: integrand in the complex plane of $w$
1310: \begin{equation}
1311: P_{N}=P_{N0}e^{-\varepsilon \func{Im}F(\rho )}, \label{PNExp}
1312: \end{equation}
1313: where
1314: \begin{equation}
1315: F(\rho )=\frac{4}{\pi }\int_{0}^{w_{c}}dw\sqrt{1+w^{2}}\left[ 1-\frac{\rho }{%
1316: (1+w^{2})^{3/2}}\right] ^{3/2}. \label{FrhoDef}
1317: \end{equation}
1318: For $\rho =0$ one obtains the LZS value $F(0)=i$. For the FM coupling the
1319: relevant singularity is
1320: \begin{equation}
1321: w_{c}=i\sqrt{1-\rho ^{2/3}},\qquad 0<\rho <1, \label{wcFM}
1322: \end{equation}
1323: that corresponds to vanishing of $\overline{\Omega }(u)$ of Eq.\ (\ref
1324: {cplusOmegaDef}) due to that of the ellipticity factor $\zeta .$ For $\rho
1325: >1,$ the motion of the spin becomes nonadiabatic (see Fig.\ \ref
1326: {Fig-lzn-P-F-Astroid}), and this method does not apply any longer. For $%
1327: 0<\rho <1$ one has $\func{Re}F(\rho )=0$ and the limiting forms of Im$F(\rho
1328: )$ are given by
1329: \begin{equation}
1330: \func{Im}F(\rho )\cong \left\{
1331: \begin{array}{ll}
1332: \displaystyle1-\frac{2}{\pi }\left( \ln \frac{32}{\rho }-1\right) \rho , &
1333: \rho \ll 1 \\
1334: \displaystyle\frac{\sqrt{3}}{2\sqrt{2}}(1-\rho )^{2}, & 1-\rho \ll 1.
1335: \end{array}
1336: \right. \label{FrhoFMLims}
1337: \end{equation}
1338: For the AFMF coupling $\rho <0,$ there is a pair of relevant singularities
1339: that also correspond to $\zeta =0$ and are given by
1340: \begin{eqnarray}
1341: w_{c\pm } &=&i\left( 1+|\rho |^{2/3}+|\rho |^{4/3}\right) ^{1/4}\exp \left(
1342: \mp i\varphi _{c}\right) \nonumber \\
1343: \varphi _{c} &=&\frac{1}{2}\arctan \frac{\sqrt{3}|\rho |^{2/3}}{2+|\rho
1344: |^{2/3}}. \label{wcAFMF}
1345: \end{eqnarray}
1346: Note that $\func{Im}(w_{c\pm })>1$ for $\rho <0$ and thus these
1347: singularities are \emph{further} from the real axis than the LZS singularity
1348: at $w=i.$ The latter, however, makes no contribution since, as can be easily
1349: checked, $\func{Im}F(\rho )=\infty $ for $w_{c}=i$ and $\rho <0.$ Analytical
1350: calculation of the limiting forms of $F(\rho )$ is more cumbersome for the
1351: antiferromagnetic coupling. For $|\rho |\ll 1$ one obtains
1352: \begin{equation}
1353: \func{Re}F(\rho )\cong 2|\rho |,\qquad \func{Im}F(\rho )\cong 1-\frac{2}{\pi
1354: }\left( \ln \frac{32}{|\rho |}-1\right) \rho . \label{FrhoAFMFsmall}
1355: \end{equation}
1356: Note that $\func{Im}F(\rho )$ is given by the same formula for $|\rho |\ll 1$
1357: and both signs of $\rho .$ In the limit $|\rho |\gg 1$ the result has the
1358: form
1359: \begin{eqnarray}
1360: F(\rho ) &\cong &\frac{2}{3\sqrt{\pi }}\frac{\Gamma (1/4)}{\Gamma (3/4)}%
1361: \left( |\rho |^{3/2}-\frac{1}{4}|\rho |^{5/6}\right) \nonumber \\
1362: &&+\frac{\sqrt{\pi }\left( \sqrt{3}+i\right) }{\Gamma (5/3)\Gamma (11/6)}%
1363: |\rho |^{2/3} \nonumber \\
1364: &\simeq &1.1|\rho |^{3/2}-0.3\,|\rho |^{5/6}+\left( 3.6+2.1i\right) |\rho
1365: |^{2/3}. \label{FrhoAFMFlarge}
1366: \end{eqnarray}
1367: The real and imaginary parts of $F(\rho )$ numerically calculated from Eq.\ (%
1368: \ref{FrhoDef}), as well as the analytical limiting forms, are shown in Fig.\
1369: \ref{Fig-lzn-Frho}.
1370:
1371: %TCIMACRO{
1372: %\TeXButton{Fig-lzn-Frho}{\begin{figure}[t]
1373: %\unitlength1cm
1374: %\begin{picture}(11,6)
1375: %\centerline{\psfig{file=Fig-lzn-Frho.eps,angle=-90,width=9cm}}
1376: %\end{picture}
1377: %\caption{ \label{Fig-lzn-Frho}
1378: %Function $F(\rho) $ of Eqs.\
1379: %(\protect\ref{PNExp}) and (\protect\ref{FrhoDef}).
1380: %}
1381: %\end{figure}%
1382: %}}%
1383: %BeginExpansion
1384: \begin{figure}[t]
1385: \unitlength1cm
1386: \begin{picture}(11,6)
1387: \centerline{\psfig{file=Fig-lzn-Frho.eps,angle=-90,width=9cm}}
1388: \end{picture}
1389: \caption{ \label{Fig-lzn-Frho}
1390: Function $F(\rho) $ of Eqs.\
1391: (\protect\ref{PNExp}) and (\protect\ref{FrhoDef}).
1392: }
1393: \end{figure}%
1394: %
1395: %EndExpansion
1396:
1397: The prefactor $P_{N0}$ in Eq.\ (\ref{PNExp}) is determined for $\varepsilon
1398: \gg 1$ by a close vicinity of the singularities $w_{c}.$ For the
1399: ferromagnetic coupling the result is
1400: \begin{equation}
1401: P_{N0}\cong \frac{\pi ^{2}}{15\varepsilon \rho \sqrt{1-\rho ^{2/3}}},\qquad
1402: 0<\rho <1. \label{P0F}
1403: \end{equation}
1404: For the AFMF coupling $\rho <0$ interference of the two contributions from $%
1405: w_{c+}$ and $w_{c-}$ of Eq.\ (\ref{wcAFMF}) results in the oscillating
1406: prefactor
1407: \begin{equation}
1408: P_{N0}\cong \frac{\pi ^{2}}{15\varepsilon |\rho ||w_{c}|}\cos ^{2}\left[
1409: \frac{\varepsilon }{2}\func{Re}F(\rho )\right] ,\qquad \rho <0. \label{P0AF}
1410: \end{equation}
1411: Oscillating prefactors of such a kind take place for\emph{\ one} tunneling
1412: particle if the sweep is nonlinear and decelerating in the vicinity of the
1413: resonance.\cite{garsch02prb} Here \emph{effective} nonlinearity of this type
1414: occurs because of the negative coupling between tunneling particles, even
1415: for the linear sweep.
1416:
1417: Both Eqs.\ (\ref{P0F}) and (\ref{P0AF}) break down in the linit $\rho
1418: \rightarrow 0$ since different singularities come close to each other. One
1419: could work out the crossover from Eqs.\ (\ref{P0F}) and (\ref{P0AF}) to the
1420: value $P_{N0}=(\pi /3)^{2}$ (see Ref.\ \onlinecite{garsch02prb}) at $\rho =0$
1421: that takes place in a narrow region \TEXTsymbol{\vert}$\rho |\varepsilon
1422: \sim 1.$ This makes no sence, however, since the result $P_{N0}=\pi /3$
1423: differs from the exact LZS prefactor $P_{0}=1$ and it has to be improved by
1424: taking into account nonlinear terms dropped during the derivation of Eq.\ (%
1425: \ref{cplusEq}). The corresponding procedure for $\rho =0$ \ is described in
1426: Ref.\ \onlinecite{garsch02prb}. In the present case $\rho \neq 0$ this would
1427: be too involved and we don't try to do it. It is clear that the logarithmic
1428: singularity of the exponent of Eq.\ (\ref{FrhoAFMFsmall}) should ve
1429: compensated for by the singularity of the prefactor so that the staying
1430: probability $P_{N}$ behaves linearly in $\rho $ near $\rho =0$ [see Eq.\ (%
1431: \ref{dPdrhoslowsweep})].
1432:
1433: Let us now consider the case $1-\rho \ll 1$ in more detail since the
1434: prefactor $P_{N0}$ of Eq.\ (\ref{P0F}) diverges at $\rho \rightarrow 1$.
1435: Here one has $\func{Im}F(\rho )\ll 1,$ and the exponential decrease of $%
1436: P_{N} $ is very slow. In this region according to Eq.\ (\ref{wcFM}) the
1437: integral in Eq.\ (\ref{PNx}) is dominated by small $w$ and it can be
1438: simplified to
1439: \begin{eqnarray}
1440: P_{N} &\cong &\frac{2}{3}\delta ^{3/2}\left| \frac{1}{2}\int_{-\infty
1441: }^{\infty }dt\left( 1+t^{2}\right) ^{1/4}e^{ia\widetilde{\Phi }(t)}\right|
1442: ^{2} \nonumber \\
1443: \widetilde{\Phi }(t) &=&\int_{0}^{t}dt^{\prime }\left( 1+t^{\prime 2}\right)
1444: ^{3/2}, \label{PNt}
1445: \end{eqnarray}
1446: where
1447: \begin{equation}
1448: \delta \equiv 1-\rho \ll 1,\qquad t\equiv \sqrt{\frac{3}{2\delta }}w,\qquad
1449: a\equiv \sqrt{\frac{2}{3}}\delta ^{2}\tilde{\varepsilon}. \label{atDef}
1450: \end{equation}
1451: It is convenient to compute the integral in Eq.\ (\ref{PNt}) by shifting the
1452: integration contour by $i$ to suppress oscillations of the integrand, i.e.,
1453: to parametrize $t=i+z,$ $-\infty <z<\infty .$ With $\widetilde{\Phi }%
1454: (i)=i3\pi /8$ this yields
1455: \begin{equation}
1456: P_{N}\cong P_{N0}e^{-a\func{Im}\widetilde{\Phi }(i)}=P_{N0}\exp \left( -%
1457: \frac{3\pi }{8}a\right) , \label{PNdelta}
1458: \end{equation}
1459: where the exponent coincides with the previously obtained $[\sqrt{3}/(2\sqrt{%
1460: 2})]\delta ^{2}\varepsilon $ [see second line of Eq.\ (\ref{FrhoFMLims})].
1461: The prefactor $P_{N0}$ can be calculated analytically for $a\gg 1$ and $a\ll
1462: 1.$ For $a\gg 1$ we need the small-$z$ expansions
1463: \begin{eqnarray}
1464: &&\delta \widetilde{\Phi }(z)\equiv \widetilde{\Phi }(i+z)-\widetilde{\Phi }%
1465: (i)\cong \frac{4}{5}\left( -1+i\right) z^{5/2} \nonumber \\
1466: &&[1+(i+z)^{2}]^{1/4}\cong (-1)^{1/8}(2z)^{1/4}. \label{SmallzExp}
1467: \end{eqnarray}
1468: Values of these functions for $z<0$ are obtained from $(-1)^{5/2}=-i$ and $%
1469: (-1)^{1/4}=(1-i)/\sqrt{2}.$ After that calculation in Eq.\ (\ref{PNt})
1470: yields
1471: \begin{equation}
1472: P_{N0}\cong \frac{2\pi }{15}\frac{\delta ^{3/2}}{a},\qquad a\gg 1
1473: \label{PN0alarge}
1474: \end{equation}
1475: that is a limiting form of Eq.\ (\ref{P0F}). In the opposite limit $a\ll 1,$
1476: the integral in Eq.\ (\ref{PNt}) is dominated by large $t,$ so that one can
1477: use $\widetilde{\Phi }(t)\cong t^{4}/4$ and $\left( 1+t^{2}\right)
1478: ^{1/4}\cong t^{1/2}.$ This yields
1479: \begin{equation}
1480: P_{N0}\cong \cos ^{2}\left( \frac{3\pi }{16}\right) \frac{\Gamma ^{2}(3/8)}{6%
1481: \sqrt{2}}\frac{\delta ^{3/2}}{a^{3/4}}\simeq \frac{0.747837}{\varepsilon
1482: ^{3/4}},\qquad \rho \rightarrow 1. \label{PN0alSmall}
1483: \end{equation}
1484: It is convenient to represent $P_{N0}(a,\delta )$ in the whole range of $a$
1485: with the help of the crossover function $f(a)$ according to
1486: \begin{eqnarray}
1487: &&P_{N0}(a,\delta )=\frac{2\pi }{15}\frac{\delta ^{3/2}}{a}f(a) \nonumber \\
1488: &&f(a)\cong \left\{
1489: \begin{array}{ll}
1490: 1, & a\gg 1 \\
1491: 1.09294a^{1/4}, & a\ll 1.
1492: \end{array}
1493: \right. \label{PN0Scaling}
1494: \end{eqnarray}
1495: Numerically computed function $f(a)$ is shown in Fig.\ (\ref
1496: {Fig-lzn-FScaling-a}).
1497:
1498: %TCIMACRO{
1499: %\TeXButton{Fig-lzn-FScaling-a.eps}{\begin{figure}[t]
1500: %\unitlength1cm
1501: %\begin{picture}(11,6)
1502: %\centerline{\psfig{file=Fig-lzn-FScaling-a.eps,angle=-90,width=9cm}}
1503: %\end{picture}
1504: %\caption{ \label{Fig-lzn-FScaling-a}
1505: %Scaling function $f(a)$ of Eq.\ (\protect\ref{PN0Scaling})
1506: %}
1507: %\end{figure}%
1508: %}}%
1509: %BeginExpansion
1510: \begin{figure}[t]
1511: \unitlength1cm
1512: \begin{picture}(11,6)
1513: \centerline{\psfig{file=Fig-lzn-FScaling-a.eps,angle=-90,width=9cm}}
1514: \end{picture}
1515: \caption{ \label{Fig-lzn-FScaling-a}
1516: Scaling function $f(a)$ of Eq.\ (\protect\ref{PN0Scaling})
1517: }
1518: \end{figure}%
1519: %
1520: %EndExpansion
1521:
1522: %TCIMACRO{
1523: %\TeXButton{Fig-lzn-Phx}{\begin{figure}[t]
1524: %\unitlength1cm
1525: %\begin{picture}(11,6)
1526: %\centerline{\psfig{file=Fig-lzn-Phx.eps,angle=-90,width=9cm}}
1527: %\end{picture}
1528: %\caption{ \label{Fig-lzn-Phx}
1529: %$P_N$ in the slow-sweep limit $\varepsilon\to\infty$ vs $h_x=1/\rho=\Delta/(2J_0)$.
1530: %}
1531: %\end{figure}%
1532: %}}%
1533: %BeginExpansion
1534: \begin{figure}[t]
1535: \unitlength1cm
1536: \begin{picture}(11,6)
1537: \centerline{\psfig{file=Fig-lzn-Phx.eps,angle=-90,width=9cm}}
1538: \end{picture}
1539: \caption{ \label{Fig-lzn-Phx}
1540: $P_N$ in the slow-sweep limit $\varepsilon\to\infty$ vs $h_x=1/\rho=\Delta/(2J_0)$.
1541: }
1542: \end{figure}%
1543: %
1544: %EndExpansion
1545:
1546: \subsection{Strong ferromagnetic interactions}
1547:
1548: For stronger ferromagnetic interactions, $\rho >1,$ the adiabatic
1549: approximation breaks down since the spin approximately follows the initial
1550: energy minimum that becomes unstable at some $h_{z}(t)>0$ (see Fig.\ \ref
1551: {Fig-lzn-P-F-Astroid}) and than it performes a large motion that does not
1552: approach the new stable energy minimum and thus cannot be linearized around
1553: it. While the problem becomes much more complicated in this case, one can
1554: still find analytically $\stackunder{\varepsilon \rightarrow \infty }{\lim }%
1555: P_{N}$ that is nonzero$.$ To this end, one can represent the LLE for the
1556: spin, Eq.\ (\ref{LLE}), in the Hamiltonian form in terms of the canonical
1557: angle variables $\left\{ \cos \theta ,\varphi \right\} \Leftrightarrow
1558: \left\{ p,q\right\} $ (in the arbitrary frame)
1559: \begin{equation}
1560: \frac{d}{dt}\cos \theta =-\frac{\partial \mathcal{H}}{\partial \varphi }%
1561: ,\qquad \frac{d}{dt}\varphi =\frac{\partial \mathcal{H}}{\partial \cos
1562: \theta } \label{HamEq}
1563: \end{equation}
1564: and use conservation of the action
1565: \begin{equation}
1566: \mathcal{S}=\oint pdq=\oint \cos \theta d\varphi \label{ActionDef}
1567: \end{equation}
1568: over the period of motion for very slowly changing parameters of the system
1569: [here $h_{z}(t)]$. Since in the final state $h_{z}\rightarrow \infty $ the
1570: spin precesses around the $z$ axis with a constant value of $s_{z}(\infty
1571: )=\cos \left[ \theta (\infty )\right] $ that is related to the staying
1572: probability $P_{N}$ by Eq.\ (\ref{PtMappingClass}) and the corresponding
1573: action is simply $\mathcal{S}(\infty )=2\pi \cos \theta (\infty ),$ one
1574: obtains
1575: \begin{equation}
1576: P_{N}=\frac{1}{2}\left( 1-\frac{1}{2\pi }\mathcal{S}\right) ,\qquad
1577: \varepsilon \rightarrow \infty , \label{PNActionDef}
1578: \end{equation}
1579: where $\mathcal{S}$ is the action over the trajectory that starts from the
1580: metastable energy minimum that is on the verge of being unstable
1581: \begin{equation}
1582: s_{zs}=-(1-h_{x}^{2/3})^{1/2},\qquad s_{xs}=h_{x}^{1/3},\qquad s_{ys}=0.
1583: \label{sStarting}
1584: \end{equation}
1585: This trajectory can be found from the conservation of energy,
1586: \begin{equation}
1587: -s_{z}^{2}-2(1-h_{x}^{2/3})^{3/2}s_{z}-2h_{x}s_{x}=1-3h_{x}^{2/3}.
1588: \label{EnergyConservation}
1589: \end{equation}
1590: Analysis shows that in the range $0<h_{x}<3\sqrt{3}/8$ this trajectory
1591: encircles the $z$ axis, thus the $z$ axis can be used as the polar axis ($%
1592: s_{z}=\cos \theta ,$ $s_{x}=\sin \theta \cos \varphi ,$ $s_{y}=\sin \theta
1593: \sin \varphi $) to compute $\mathcal{S}.$ In the overlapping range $%
1594: 1/8<h_{x}<1$ the trajectory encircles the $x$ axis that can be used as the
1595: polar axis ($s_{x}=\cos \theta ,$ $s_{z}=-\sin \theta \cos \varphi ,$ $%
1596: s_{y}=\sin \theta \sin \varphi $). With these choices, $\cos \theta $ can be
1597: found numerically from Eq.\ (\ref{EnergyConservation}) as a function of $%
1598: \varphi $ and the action $\mathcal{S}$ can be computed from Eq.\ (\ref
1599: {ActionDef}). The result for $P_{N}$ vs $h_{x}$ is shown in Fig.\ \ref
1600: {Fig-lzn-Phx} with asymptotes
1601: \begin{equation}
1602: P_{N}\cong \left\{
1603: \begin{array}{ll}
1604: 1-(3/2)h_{x}^{2/3}, & h_{x}\ll 1 \\
1605: 0.544861\left( 1-h_{x}\right) ^{3/2}, & 1-h_{x}\ll 1.
1606: \end{array}
1607: \right. \label{PhxLims}
1608: \end{equation}
1609:
1610: \section{Discussion}
1611:
1612: The simplified model of interacting tunneling particles that was considered
1613: above allows to quantify the influence of interaction on the
1614: Landau-Zener-Stueckelberg staying probability $P$. It was done here by
1615: numerically solving the problem for up to $N=200$ interacting particles as
1616: well as using a number of analytical and half-analytical approaches,
1617: including the mean-field limit $N\rightarrow \infty $.
1618:
1619: In accord with physical expectations and considerations of the exact levels
1620: of the system in Fig.\ \ref{Fig-lzn-En}a, the ferromagnetic coupling tends
1621: to suppress LZS transitions to another bare energy level (i.e., the state on
1622: the other side of the energy barrier), in agreement with Ref.\ %
1623: \onlinecite{hamraemiysai00}. For $N\rightarrow \infty $ there is a critical
1624: value of the coupling above which the staying probability $P$ does not go to
1625: zero in the slow-sweep limit $\varepsilon \rightarrow \infty $, in contrast
1626: with the standard LZS case. For finite $N$ the dependence of $P$ on $%
1627: \varepsilon $ is a sum of many exponentials with greatly differing
1628: relaxation rates that is very slow approaching zero (see Fig.\ \ref
1629: {Fig-lzn-P-F-j3}). The same should be the case for more realistic
1630: interactions of the ferromagnetic type.
1631:
1632: The negative coupling in our model (that corresponds to the
1633: antiferromagnetic frustrating coupling) tends to boost the LZS transitions.
1634: In the limit $N\rightarrow \infty $ the staying probability $P$ even turns
1635: to zero at \emph{finite} values of the sweep rate while the \emph{effective}
1636: sweep rate becomes an odd function of time with retardation in the
1637: resonance-crossing region (see Fig.\ \ref{Fig-lzn-PWeffvst-AF-MFA-j10m}). On
1638: the other hand, models with more realistic antiferromagnetic interactions
1639: such as the nearest-neighbor interaction have the energy-level scheme
1640: strongly differing from that shown in Fig.\ \ref{Fig-lzn-En}b. Preliminary
1641: results show that such interactions tend to hamper LZS transitions instead
1642: of boosting it, although not to such an extent as ferromagnetic interactions.
1643:
1644: Theoretically the model considered here is interesting since it maps on the
1645: problem of a single large spin $S=N/2$ and thus the mean-field limit $%
1646: N\rightarrow \infty $ corresponds to the classical limit $S\rightarrow
1647: \infty $ for the large spin. It would be very interesting to study
1648: deviations from the mean-field solutions for large but finite $N.$ These
1649: deviations can be very large, as can be seen in Fig.\ \ref{Fig-lzn-P-AF-j10}
1650: in the region $\varepsilon \simeq 0.11.$ How quantitatively well does the
1651: MFA work for model systems with more complicated interactions and for
1652: realistic systems remains unclear, and it is an interesting topic for
1653: further work.
1654:
1655: \section{\protect\bigskip Acknowledgments}
1656:
1657: Many useful discussions with Rolf Schilling are greatfully acknowledged.
1658:
1659: %\bigskip
1660:
1661: %\bibliographystyle{prsty}
1662: %\bibliography{gar-general,gar-tunneling,gar-own,gar-oldworks}
1663:
1664: \begin{thebibliography}{99}
1665: \bibitem{lan32} {L. D. Landau}, Phys. Z. Sowjetunion \textbf{2}, 46 (1932).
1666:
1667: \bibitem{zen32} C. Zener, Proc. R. Soc. London A \textbf{137}, 696 (1932).
1668:
1669: \bibitem{stu32} E.~C.~G. Stueckelberg, Helv. Phys. Acta \textbf{5}, 369
1670: (1932).
1671:
1672: \bibitem{crohug77} {D. S. F. Crothers and J. G Huges}, J. Phys. B \textbf{10%
1673: }, L557 (1977).
1674:
1675: \bibitem{werses99} {W. Wernsdorfer and R. Sessoli}, Science \textbf{284},
1676: 133 (1999).
1677:
1678: \bibitem{weretal00epl} {W. Wernsdorfer, R. Sessoli, A. Caneshi, D.
1679: Gatteschi, and A. Cornia}, Europhys. Lett. \textbf{50}, 552 (2000).
1680:
1681: \bibitem{hamraemiysai00} {A. Hams, H. De Raedt, S. Miyashita, and K. Saito}%
1682: , Phys. Rev. B \textbf{62}, 13880 (2000).
1683:
1684: \bibitem{garsch02prb} {D. A. Garanin and R. Schilling}, Phys. Rev. B
1685: \textbf{66}, 174438 (2002).
1686:
1687: \bibitem{akusch92} {V. M. Akulin and W. P. Schleich}, Phys. Rev. A \textbf{%
1688: 46}, 4110 (1992).
1689:
1690: \bibitem{dobzve97} {V. V. Dobrovitski and A. K. Zvezdin}, Europhys. Lett.
1691: \textbf{38}, 377 (1997).
1692:
1693: \bibitem{gar91jpa} {D. A. Garanin}, J. Phys. A \textbf{24}, L61 (1991).
1694:
1695: \bibitem{ternak97} {Y. Teranishi and H. Nakamura}, J. Chem. Phys. \textbf{%
1696: 107}, 1904 (1997).
1697:
1698: \bibitem{garsch02} {D. A. Garanin and R. Schilling}, Europhys. Lett.
1699: \textbf{59}, 7 (2002).
1700:
1701: \bibitem{stowoh} {E. C. Stoner and E. P. Wohlfarth}, Philos. Trans. R. Soc.
1702: London, Ser. A \textbf{\ 240}, 599 (1948); IEEE Trans. Magn. \textbf{MAG-27}%
1703: , 3475 (1991).
1704: \end{thebibliography}
1705:
1706: \end{document}
1707: