1: % -------------------------------------------------------------
2: % Surzhykov, Lubasch, Zinn-Justin and Jentschura
3: % -------------------------------------------------------------
4: % Quantum Dot Potentials:
5: % Symanzik Scaling, Resurgent Expansions and Quantum Dynamics
6: % -------------------------------------------------------------
7: % Submission to Physical Review B
8: % ------------------------------------------------------------
9: %\documentclass[preprint,superscriptaddress,aps,prb,showpacs,floatfix]{revtex4}
10: \documentclass[twocolumn,superscriptaddress,aps,prb,showpacs,floatfix]{revtex4}
11:
12:
13: %\usepackage{showkeys}
14:
15: \usepackage{amsmath}
16: \usepackage{amssymb}
17: \usepackage{amsfonts}
18: \usepackage{pifont}
19: \usepackage{colordvi}
20: \usepackage{bm}
21: \usepackage{bbm}
22: \usepackage{epsf}
23: \usepackage{epsfig}
24: \usepackage{times}
25: \usepackage{nicefrac}
26: \usepackage{colordvi}
27: \usepackage{color}
28: \usepackage{graphicx}
29: \usepackage{dcolumn}
30: %\newcolumntype{.}{D{x}{}{-1}}
31:
32: %
33: % Our definitions
34: %
35:
36: \def\sprm#1#2{ \left\langle #1 \left\vert \right. #2 \right\rangle }
37: \def\mem#1#2#3{ \left\langle #1 \left\vert #2 \right\vert #3 \right\rangle }
38:
39: \bibliographystyle{myprsty}
40:
41: \begin{document}
42:
43: \preprint{Version 3.1}
44:
45: \title{Quantum Dot Potentials: Symanzik Scaling,
46: Resurgent Expansions and Quantum Dynamics}
47:
48: \author{Andrey Surzhykov}
49: \affiliation{Max--Planck--Institut f\"ur Kernphysik,
50: Saupfercheckweg 1, 69117 Heidelberg, Germany}
51:
52: \author{Michael Lubasch}
53: \affiliation{Max--Planck--Institut f\"ur Kernphysik,
54: Saupfercheckweg 1, 69117 Heidelberg, Germany}
55:
56: \author{Jean Zinn--Justin}
57: \affiliation{DAPNIA, Commissariat \`{a} l'\'{E}nergie Atomique,
58: Centre de Saclay, 91191 Gif--Sur--Yvette, France}
59:
60: \author{Ulrich D.~Jentschura\footnote{e-mail: ulj@mpi-hd.mpg.de}$^{,}$}
61: \affiliation{Max--Planck--Institut f\"ur Kernphysik,
62: Saupfercheckweg 1, 69117 Heidelberg, Germany}
63:
64: \date{\today}
65:
66: \begin{abstract}
67: This article is concerned with a special class of the ``double--well--like''
68: potentials that occur naturally in the analysis of finite
69: quantum systems. Special attention is paid, in particular, to
70: the so--called Fokker--Planck potential, which has a particular property:
71: the perturbation series for the ground--state energy vanishes to all orders
72: in the coupling parameter, but the actual ground-state energy
73: is positive and dominated by instanton configurations of the form
74: $\exp(-a/g)$, where $a$ is the instanton action. The instanton
75: effects are most naturally taken into account within
76: the modified Bohr--Sommerfeld quantization conditions whose
77: expansion leads to the generalized perturbative expansions
78: (so-called resurgent expansions) for the energy eigenvalues of the
79: Fokker--Planck potential. Until now, these
80: resurgent expansions have been mainly applied for small values
81: of coupling parameter $g$, while much less attention has been paid to
82: the strong-coupling regime. In this contribution, we compare the
83: energy values, obtained by directly resumming
84: generalized Bohr--Sommerfeld quantization conditions, to the strong-coupling
85: expansion, for which we determine the first few expansion coefficients
86: in powers of $g^{-2/3}$. Detailed calculations are performed
87: for a wide range of coupling parameters $g$ and indicate a considerable
88: overlap between the regions of validity of the
89: weak-coupling resurgent series and
90: of the strong--coupling expansion. Apart from the analysis of the
91: energy spectrum of the Fokker--Planck Hamiltonian, we also briefly
92: discuss the computation of its eigenfunctions. These
93: eigenfunctions may be utilized for the numerical integration of the
94: (single-particle) time-dependent Schr\"odinger equation and, hence, for
95: studying the dynamical evolution of the
96: wavepackets in the double--well--like potentials.
97: \end{abstract}
98:
99: % 11.15.Bt General properties of perturbation theory
100: % 11.10.Jj Asymptotic problems and properties
101: % 68.65.Hb Quantum dots
102: % 73.21.La Quantum dots
103: % 03.67.-a Quantum information
104: % 03.67.Lx Quantum computation
105: % 85.25.Cp Josephson devices
106: \pacs{11.15.Bt, 11.10.Jj, 68.65.Hb, 73.21.La, 03.67.-a, 03.67.Lx, 85.25.Cp}
107:
108: \maketitle
109:
110: % \tableofcontents
111:
112: \section{Introduction}
113: \label{intro}
114:
115: In this paper, we study the all-order summation of instanton contributions
116: to the energy eigenvalues of anharmonic quantum mechanical oscillators
117: which involve (almost) degenerate minima.
118: The Euclidean path integral of quantum mechanical systems of
119: this kind with one space-
120: and one time-dimension is dominated by instanton configurations whose
121: action remains finite in the limit of large Euclidean (imaginary) times.
122: In order to find the energy eigenvalues, instanton configurations
123: have to be taken into account.
124: Modified quantization conditions have been conjectured for various
125: classes of potentials (for a review see
126: Refs.~\onlinecite{ZJJe2004i,ZJJe2004ii}),
127: and accurate numerical calculations have been
128: verified against analytic expansions in the regime of small
129: coupling~\cite{JeZJ2001,JeZJ2004plb}.
130: Indeed, for small coupling, the energy eigenvalues are dominated
131: by one-, two-, and three-instanton effects which correspond to
132: trajectories of the classical particle with
133: a small number of oscillations between the (almost) degenerate minima
134: of the potential. Here, we are concerned with the all-order
135: summation of the instanton contributions, which is applicable
136: to intermediate and strong coupling. Also, the transition
137: from small to strong coupling, and strong-coupling expansions,
138: will be discussed. Finally, we consider applications in
139: the quantum dynamical simulation of finite systems.
140:
141: The purpose of this paper thus is threefold:
142: first, to find generalized perturbative expansions
143: (so-called ``resurgent expansions'') for excited states
144: of certain classes of notoriously problematic~\cite{HeSi1978}
145: quantum mechanical potentials, second, to derive large-coupling
146: asymptotics for these potentials and to investigate
147: overlap regions between small- and large-coupling asymptotics,
148: and third, to outline applications of the considerations
149: for the quantum dynamical simulation of a single particle
150: in a double-well-like potential.
151: The first of these purposes is connected with mathematical physics,
152: the second is tied to Symanzik scaling in the
153: ``poor man's'' variant and therefore, to a basic implementation
154: of the renormalization group, and the third one is
155: rather application-oriented.
156:
157: Within the first and second aims of our investigation,
158: we also investigate the fundamental question
159: whether small-coupling perturbative expansions
160: can be continued analytically to the regime of
161: large coupling, if one includes instantons into the
162: formalism. Instantons can be considered
163: either on the level of a resurgent expansion,
164: augmented by an optimized (generalized) Borel-Pad\'{e} resummation,
165: or on the level of a generalized quantization condition,
166: which entails an all-order resummation of the
167: instanton expansion.
168:
169: The third application is mainly tied to the semiconductor
170: ``double quantum dot'' structures~\cite{LuEtAl1989,LoDV1998,HoEtAl2002,HuEtAl2005},
171: which are formed from two quantum dots coupled by quantum
172: mechanical tunneling. Nowadays, these
173: structures are generally accepted to belong to one of the most hopeful
174: candidates for the realization of quantum bits (qubits), because
175: a single electron state in a double-well potential obviously can
176: be localized in either of the two wells and, in that sense, represents a
177: two-quantum-state system needed for quantum computing.
178: Indeed, the theoretical analysis of the
179: structural and the dynamical properties of such (single-electron)
180: double quantum dots can be traced back to double-well-like
181: potentials. In this context, quantum dynamical calculations for a
182: tunneling of a single particle between the two localized
183: states nowadays attract special interest \cite{GrDiJuHa1991, Op1999} but
184: require a detailed knowledge on the eigenvalues and eigenfunctions
185: of the (double-well-potentials) potentials for a wide range
186: of coupling parameters.
187:
188: Another question may as well be asked: The effective
189: instanton-related expansion parameter,
190: which reads $\Xi_1(g) = \sqrt{2/\pi}\, {\rm e}^{-1/6g}/g$
191: for the first two excited states of the Fokker--Planck
192: potential (as discussed below), is nonperturbatively
193: small for $g \to 0$, but numerically not very
194: small for some very moderate $g$. Specifically,
195: $\Xi_1(g)$ reaches its maximum $\Xi_1(1/6) = 1.76115\dots$
196: already at a rather small coupling parameter $g = 1/6$.
197: So, one may ask how the ``instanton expansion''
198: in powers of $\Xi_1(g)$ should be resummed, in addition to the
199: perturbative expansions about each instanton.
200: This latter step has never been accomplished, and we pursue its completion
201: via a direct resummation of generalized quantization conditions.
202:
203: This paper is organized as follows.
204: In Sec.~\ref{frame}, basic definitions
205: related to the Fokker--Planck and the double-well
206: potential are recalled. Calculations are described
207: in Sec.~\ref{resumm}. Specifically, we consider
208: the resummation of the resurgent expansion
209: in Sec.~\ref{resurg}, the resummation of the
210: quantization condition in Sec.~\ref{resumq},
211: large-coupling asymptotics in Sec.~\ref{largec},
212: and quantum dynamic simulations in Sec.~\ref{quadyn}.
213: Conclusions are drawn in Sec.~\ref{conclu}.
214:
215: %
216: % Basic framework and numerical procedure
217: %
218: \section{Basic framework and numerical procedure}
219: \label{frame}
220:
221: %
222: % Basic formulation
223: %
224: \subsection{Basic formulation}
225: \label{basic}
226:
227: In this manuscript, we discuss the determination of the eigenvalues of the
228: one--dimension Fokker--Planck (FP) Hamiltonian
229: %
230: \begin{equation}
231: \label{hamfp}
232: H_{\rm FP} = -\frac{1}{2} \left( \frac{d}{dq} \right)^2 \, + \,
233: \frac{1}{2} q^2 \left( 1 - \sqrt{g}q \right)^2 \, + \, \sqrt{g} q \, - \,
234: \frac{1}{2} \,,
235: \end{equation}
236: %
237: where $g$ is a positive coupling constant.
238: For $g \, = \, 0$, Eq.~(\ref{hamfp}) represents
239: the Hamiltonian of the quantum harmonic oscillator whose eigenvalues are
240: given by the well known formula $E^{(K)} \, = \, K$, where
241: $K = 0, 1, 2,\dots$ is the ``principal'' quantum number. For
242: nonvanishing coupling, in contrast, no
243: closed-form analytic expressions have been derived so far,
244: and approximations have to be used (for a classification
245: of the Fokker--Planck Hamiltonian in terms of a SUSY
246: algebra, see App.~\ref{appa}). The usefulness
247: of the notation $K$ instead of $N$ will become clear
248: in the following. If one considers the operator
249: $V(g) = \sqrt{g} \, q - \sqrt{g} \, q^3 + g \, q^4/2$ in (\ref{hamfp})
250: as a perturbation and formally applies
251: Rayleigh--Schr\"odinger perturbative expansion
252: to the $K$th harmonic oscillator state, then one finds the following
253: result for the first terms,
254: %
255: \begin{equation}
256: \label{epert}
257: E_{\rm FP, pert}^{(K)}(g) = K - 3K^2 g -
258: \left(17 \, K^3 + \frac{5}{2} \, K \right) g^2 +
259: \mathcal{O}(g^3) \,.
260: \end{equation}
261: %
262: All coefficients up to order $g^{80}$
263: are available for download~\cite{JeHomeHD}.
264: This perturbation expansion~\cite{HeSi1978}
265: fails to reproduce the spectrum of the Hamiltonian
266: (\ref{hamfp}) even qualitatively. For instance, while the
267: true ground--state energy $E^{(K = 0)}_{\rm FP}$ is manifestly nonvanishing
268: and positive, the perturbation series (\ref{epert}),
269: for $K = 0$, vanishes identically to all orders in the coupling $g$
270: and is thus formally converging to a zero energy eigenvalue.
271: A generalization of perturbation theory is required, therefore,
272: in order to correctly describe the physical properties of the
273: Fokker--Planck Hamiltonian, including its energy spectrum.
274:
275: A complete description
276: of the eigenvalues of the Hamiltonian (\ref{hamfp})
277: has been proposed recently~\cite{JeZJ2004plb,ZJJe2004i,ZJJe2004ii} by
278: using a generalized perturbation series involving instanton
279: contributions. Since the concept of instantons in quantum mechanics has
280: been presented in a number of places
281: \cite{JeZJ2004plb,ZJJe2004i,ZJJe2004ii,ZJ1984jmp}, we may here restrict
282: ourselves to a rather short account of the basic formulas. In the
283: semi--classical framework, the eigenvalues of the Fokker--Planck
284: Hamiltonian can be found by solving the generalized Bohr--Sommerfeld
285: quantization condition \cite{JeZJ2004plb,ZJ1984jmp}
286: %
287: %
288: \begin{eqnarray}
289: \label{qcond}
290: \lefteqn{\frac{1}{\Gamma(-B_{\rm FP}(E, g)) \,\,
291: \Gamma(1-B_{\rm FP}(E, g))}} \nonumber\\[2ex]
292: & & + \, \left(-\frac{2}{g}\right)^{2B_{\rm FP}(E, g)}
293: \frac{{\rm e}^{-A_{\rm FP}(E, g)}}{2 \pi} \, = \, 0 \,.
294: \end{eqnarray}
295: %
296: %
297: In this expression, the functions $B_{\rm FP}(E, g)$ and
298: $A_{\rm FP}(E, g)$ determine the perturbative expansion and the
299: perturbative expansion about the instantons, correspondingly.
300:
301: The evaluation of these functions in terms of series in variables $E$ and $g$
302: has been described in detail elsewhere \cite{ZJJe2004i,ZJJe2004ii,ZJ1984jmp}
303: for rather general classes of
304: potentials. In the particular case of the Fokker--Planck potential, for
305: example, the function $B_{\rm FP}(E, g)$ has the following expansion
306: [see Eq.~(14a) of Ref.~\onlinecite{JeZJ2004plb}]:
307: %
308: \begin{eqnarray}
309: \label{B_expansion}
310: B_{\rm FP}(E, g) \, = E + 3E^2g +
311: \left(35E^3 + \frac{5}{2}E\right)g^2 + \mathcal{O}(g^3) \, .
312: \end{eqnarray}
313: %
314: The function
315: $B_{\rm FP}(E, g)$ alone
316: defines the perturbation expansion (\ref{epert}) which
317: can be easily found by inverting the equation $B_{\rm FP}(E, g) \, = \,
318: K$. The instanton contributions to the eigenvalues of the Fokker--Planck
319: Hamiltonian are described by the function
320: [see Eq.~(14b) of Ref.~\onlinecite{JeZJ2004plb}]:
321: %
322: \begin{eqnarray}
323: \label{A_expansion}
324: \lefteqn{A_{\rm FP}(E, g) \, = \, \frac{1}{3g} \, + \,
325: \left( 17E^2 +\frac{5}{6}\right) \, g } \nonumber\\[2ex]
326: & & + \, \left(227E^3 + \frac{55}{2}E\right)g^2 \, + \,
327: \mathcal{O}(g^3) \, .
328: \end{eqnarray}
329: %
330: Extensive numerical checks of the generalized quantization condition
331: (\ref{qcond}) and the expansions (\ref{B_expansion})--(\ref{A_expansion}) have
332: recently been performed for the ground state of the Fokker--Planck potential
333: in the weak coupling regime \cite{JeZJ2004plb}. However, to the best of our
334: knowledge, a numerical verification of these formulae
335: (i) for excited states and (ii) for the large values of $g$
336: is still missing. The numerical checks will be presented in Sec.~\ref{resumq}.
337:
338: While, of course, the present work is mainly devoted to the investigation of
339: the energies and the corresponding
340: wave functions of the Fokker--Planck potential, we will
341: also briefly recall the properties of the well--known double-well potential
342: which is characterized by the Hamiltonian
343: %
344: %
345: \begin{equation}
346: \label{hamdw}
347: H_{\rm dw} = -\frac{1}{2} \left( \frac{d}{dq} \right)^2 \, + \,
348: \frac{1}{2} q^2 \left( 1 - \sqrt{g}q \right)^2 \, .
349: \end{equation}
350: %
351: %
352: Moreover, for the analysis of the energy spectra of the Hamiltonians
353: (\ref{hamfp}) and (\ref{hamdw}) it is very convenient to introduce
354: the interpolating potential:
355: %
356: %
357: \begin{equation}
358: \label{hami}
359: H_{\rm I} = -\frac{1}{2} \left( \frac{d}{dq} \right)^2 \, + \,
360: \frac{1}{2} q^2 \left( 1 - \sqrt{g}q \right)^2 \, +
361: \eta \left( \sqrt{g} q - \frac{1}{2} \right)\, ,
362: \end{equation}
363: %
364: which corresponds to the double-well potential if $\eta = 0$,
365: whereas $\eta = 1$ gives the Fokker--Planck case.
366:
367: %
368: % Numerical calculation of eigenenergies
369: %
370: \subsection{Numerical calculation of eigenenergies}
371: \label{numer}
372:
373: In order to numerically calculate energy eigenvalues of the
374: Fokker--Planck and of the double--well potential,
375: it is sufficient to consider matrix elements of these
376: potentials in the basis of harmonic oscillator eigenfunctions,
377: and to perform matrix diagonalization in a large basis
378: spanned by harmonic oscillator eigenfunctions
379: (typically, a number of $\gtrsim 5000$ basis functions is sufficient
380: for all calculations reported in the current article).
381: One then observes the apparent convergence of the
382: eigenenergies as the size of the basis is increased.
383: This procedure is numerically stable
384: provided one uses quadrupole precision (128-bit, 32 decimal figure)
385: arithmetic.
386:
387: %
388: %
389: \begin{figure}[t]
390: %\vspace*{-1cm}
391: \begin{center}
392: \includegraphics[height=0.9\linewidth,angle=270]{Fig1.eps}
393: \end{center}
394: %\vspace*{-5cm}
395: \caption{\label{fig1}
396: (Color online) Eigenvalues $E_{\rm FP}^{(M = 0)}$ (left panel) and
397: $E_{\rm FP}^{(M = 1,2)}$ (right panel) of the Fokker--Planck Hamiltonian
398: as a function of the coupling parameter $g$. Results have been computed by the
399: diagonalization of the Fokker--Planck Hamiltonian
400: (\ref{hamfp}) in the basis of the harmonic
401: oscillator wavefunctions.}
402: \end{figure}
403: %
404: %
405:
406: Fokker--Planck energies of the lowest three eigenstates
407: [of the potential (\ref{hamfp})]
408: found by matrix diagonalization are displayed in Fig.~\ref{fig1}
409: as a function of the coupling $g$ (we use the notation
410: $M = 0,1,2$ in order to denote these three energy levels). As seen
411: from this Figure, the states $M = 1,2$ are degenerate
412: in the limit $g \to 0$. A similar energy level
413: splitting is well known for the symmetric double--well potential
414: \cite{ZJJe2004i,ZJJe2004ii,JeZJ2001} and may be explained in terms of
415: nonperturbative instanton contributions. In contrast to the
416: double--well case, the Fokker--Planck potential contains a linear
417: symmetry--breaking term [cf. Eq.~(\ref{hamfp})], and this term
418: might be expected to lift any degeneracy.
419: However, excited states can still develop a degeneracy
420: for $g \to 0$ in view of the (only perturbatively broken) parity
421: $\varepsilon \, = \, \pm 1$ of the quantum eigenstates.
422: The ground state of the Fokker--Planck potential, however,
423: is located in one of the wells and does not develop
424: any degeneracy due to parity (see also Fig.~\ref{fig9} below).
425:
426: It is interesting to investigate the adiabatic following
427: of eigenvalues for the interpolating potential (\ref{hami})
428: as a function of the parameter $\eta$ for fixed $g$.
429: This calculation (see Fig.~\ref{fig2}) reveals that the
430: identification of the double-well
431: energy eigenvalues~\cite{JeZJ2001,ZJJe2004i,ZJJe2004ii} with
432: quantum numbers $(N,\varepsilon)$ for the double-well
433: with the quantum number $M$
434: for the Fokker--Planck potential should proceed as follows:
435: %
436: \begin{subequations}
437: \label{ident}
438: \begin{align}
439: (N=0,+) \Leftrightarrow & \; M=0\,, \\
440: (N=0,-) \Leftrightarrow & \; M=1\,, \\
441: (N=1,+) \Leftrightarrow & \; M=2\,, \\
442: (N=1,-) \Leftrightarrow & \; M=3\,.
443: \end{align}
444: \end{subequations}
445: %
446: The general relation is $M = 2 N + (1 - \varepsilon)/2$.
447: However, the asymptotic behavior of the eigenenergies
448: for $g \to 0$ is different in the two cases:
449: %
450: \begin{subequations}
451: \begin{align}
452: \label{NtoN}
453: E_{\rm dw}^{(N,\varepsilon)}(g) \to & \; N + \frac12\,,
454: \qquad g \to 0 \,, \\
455: \label{MtoK}
456: E_{\rm FP}^{(M)}(g) \to & \;
457: [\mkern - 2.5 mu [ (M+1)/2 ] \mkern - 2.5 mu ] \,,
458: \qquad g \to 0 \,,
459: \end{align}
460: \end{subequations}
461: %
462: where $[\mkern - 2.5 mu [x] \mkern - 2.5 mu ]$ is the integral part of
463: $x$, i.e., the largest integer $m$ satisfying $m \le x$.
464: Equation (\ref{MtoK}) implies that the perturbative contribution
465: to the Fokker--Planck energy level with quantum number $M$ is
466: given by Eq.~(\ref{epert}) with
467: $K = [\mkern - 2.5 mu [ (M+1)/2 ] \mkern - 2.5 mu ]$.
468:
469: Apart from the degeneracy introduced by the instanton contributions, the
470: eigenvalues $E^{(M = 1,2)}_{\rm FP}(g)$ also have a
471: qualitatively different dependence on $g$ when compared to the
472: ground--state energy $E^{(M = 0)}_{\rm FP}(g)$. As seen from Fig.~\ref{fig1}, while the
473: energy $E^{(M = 0)}_{\rm FP}(g)$ increases monotonically as a function of the
474: coupling constant $g$, the energies $E^{(M = 1,2)}_{\rm FP}(g)$
475: have minima at
476: $g_0 \, \approx \, 0.07$ and $g_0 \, \approx \, 0.025$, respectively.
477:
478: %
479: %
480: \begin{figure}[ht]
481: %\vspace*{-1.0cm}
482: \begin{center}
483: \includegraphics[width=0.7\linewidth,angle=270]{Fig2.eps}
484: \end{center}
485: %\vspace*{1.0cm}
486: \caption{\label{fig2} (Color online)
487: Adiabatic following of the lowest four eigenvalues
488: of the interpolating potential (\ref{hami}) from $\eta = 0$
489: (double-well) to $\eta = 1$ (Fokker--Planck potential).
490: The value of the coupling parameter is held constant at
491: $g = 0.007$.}
492: \end{figure}
493: %
494: %
495:
496: %
497: % Resummations
498: %
499: \section{Resummations, Energy Eigenvalues and Quantum Dynamics}
500: \label{resumm}
501:
502: %
503: % Resummation of the resurgent expansion
504: %
505: \subsection{Resummation of the resurgent expansion}
506: \label{resurg}
507:
508: The generalized Bohr--Sommerfeld quantization condition (\ref{qcond}) together
509: with the expansions (\ref{B_expansion})---(\ref{A_expansion}) of the
510: $A_{\rm FP}$ and $B_{\rm FP}$ functions uniquely determines the eigenvalues
511: of the Fokker--Planck Hamiltonian. For the ground state,
512: the energy eigenvalue can be found by systematic expansion
513: of (\ref{qcond}) in powers of the two small parameters $\exp(-1/3g)$ and $g$,
514: whereas for excited states, the parameters are $\exp(-1/6g)$ and $g$.
515: Thus, for the Fokker--Planck potential, the particular form of the
516: expansion differs for the ground vs.~excited states.
517:
518: Explicitly, the ground-state energy ($M=0$) is given by the resurgent
519: expansion \cite{JeZJ2004plb}:
520: %
521: %
522: \begin{equation}
523: \label{eground}
524: E^{(0)}_{\rm FP}(g) =
525: \sum^{\infty}_{n=1} \left( {{\rm e}^{-1/3g} \over 2 \pi} \right)^{n} \,
526: \sum^{n-1}_{k=0} \left\{ \ln\left(-\frac{2}{g}\right) \right\}^k \,
527: \sum^{\infty}_{l=0} f^{(0)}_{nkl} \, g^{l} \, ,
528: \end{equation}
529: %
530: %
531: where the index $n$ denotes the order of the ``instanton contribution'':
532: $n = 1$ is a one-instanton, $n=2$ is a two-instanton, etc.
533: (as noted below, the one-instanton configuration involves a
534: back-tunneling of the particle to the lower well for the
535: ground state and thus has twice the action of the characteristic
536: one-instanton effect for excited states). Another subtle point which
537: should be recalled here is that the leading one-instanton term
538: involves a summation over all possible $n$--instanton configurations but
539: neglects instanton interactions \cite{ZJJe2004i}. As seen from
540: Eq.~(\ref{eground}), the evaluation
541: of the ground state energy within such a (first--order) approximation,
542: also referred as a ``dilute instanton gas'' approximation, requires the
543: knowledge of the $f^{(0)}_{10l}$ coefficients. Since these coefficients are
544: available for download~\cite{JeHomeHD}, we only recall the six leading
545: ones~\cite{JeZJ2004plb}:
546: %
547: %
548: \begin{eqnarray}
549: f^{(0)}_{100} &=& 1 \,, \quad
550: f^{(0)}_{101} = -\frac{5}{6} \,, \quad
551: f^{(0)}_{102} = -\frac{155}{72} \,, \nonumber \\[1ex]
552: f^{(0)}_{103} &=& -\frac{17315}{1296} \,, \quad
553: f^{(0)}_{104} = -\frac{3924815}{31104} \,, \\[1ex]
554: f^{(0)}_{105} &=& -\frac{294332125}{186624} \,, \quad
555: f^{(0)}_{106} = -\frac{163968231175}{6718464} \,. \nonumber
556: \end{eqnarray}
557: %
558: %
559: By inserting these coefficients in Eq.~(\ref{eground}), we are able
560: to perform now the numerical check of the validity of the one-instanton
561: expansion for the ground state at small coupling. In Table~\ref{table1},
562: for example, the energy $E^{(0)}_{\rm FP}(g)$ is displayed for coupling
563: parameters in the range 0.005 $\le g \le$ 0.03 and is compared to the
564: ``true'' eigenvalues as obtained from the diagonalization of the
565: Fokker--Planck potential in the basis of harmonic oscillator wavefunctions.
566: As seen from Table~\ref{table1}, the ground state energy is
567: dominated by the one-instanton effect for relatively small
568: values of the coupling parameter $g \le$ 0.01.
569: For stronger coupling, however,
570: large discrepancies between the ``true'' energies and the results of
571: resummation of Eq.~(\ref{eground}) at $n$ = 1 are found indicating
572: the importance of the higher--instanton terms which take into account
573: the instanton interactions. The evaluation of the higher--order corrections
574: ($n \ge$ 2) to the ground--state energy $E^{(0)}_{\rm FP}(g)$ is, however,
575: a very difficult task since it requires a double resummation of the resurgent
576: expansion, in powers of both $g$ and $\exp(-1/3g)$. In the
577: present work such a double summation based on sequentially
578: adding higher-order instanton terms will not be performed.
579: Still, for the sake of completeness, we here indicate the leading
580: two--instanton \cite{JeZJ2004plb} and three--instanton coefficients
581: for the Fokker--Planck ground state:
582: %
583: %
584: \begin{eqnarray}
585: f^{(0)}_{210} &=& 2 \,, \quad
586: f^{(0)}_{200} = 2 \gamma \,, \nonumber \\[1ex]
587: f^{(0)}_{211} &=& -\frac{10}{3} \,, \quad
588: f^{(0)}_{201} = -\frac{10}{3} \gamma - 3 \,, \\[1ex]
589: f^{(0)}_{310} &=& 8 \, \gamma \,, \quad
590: f^{(0)}_{300} = 6 \gamma^2 + \frac{\pi^2}{6}\,, \nonumber \\[1ex]
591: f^{(0)}_{311} &=& -\frac{80}{3} \gamma - 6 \,, \quad
592: f^{(0)}_{301} = -15 \gamma^2 - 12 \gamma -17 + \frac{5}{12} \pi^2 \,, \nonumber
593: \end{eqnarray}
594: %
595: %
596: where $\gamma = 0.577216\dots$ is Euler's constant.
597:
598: %
599: % Table 1
600: %
601: \begin{table}[t]
602: \caption{\label{table1} Ground--state energy of the Fokker--Plank
603: Hamiltonian. Results have been computed by the diagonalization of the
604: Hamiltonian in the basis of the harmonic oscillator wavefunctions ("true"
605: energy) and by resummation of Eq.~(\ref{eground}) within the one-instanton
606: approximation ($n$ = 1). The numerical uncertainty of the entries
607: in the right column is estimated on the basis of the apparent
608: convergence of results under an appropriate
609: increase of the number of $f^{(0)}_{10l}$ parameters, which corresponds
610: to the number of terms in the perturbative expansion about the leading
611: instanton. Numerical discrepancies between the left and right column
612: are due to higher-order instanton contributions, as described in the text.
613: We underline those decimal figures in the one-instanton results
614: which are equal to the corresponding ones in the complete
615: numerical solution.}
616: \vspace*{0.3cm}
617: \begin{scriptsize}
618: \begin{center}
619: \begin{tabular}{l@{\hspace*{0.5cm}}l@{\hspace*{0.5cm}}l}
620: \hline
621: \hline
622: \rule[-3mm]{0mm}{8mm}
623: $g$ & $E^{(M = 0)}_{\rm FP}$ (diagonalization)
624: & $n = 1$ term of Eq.~(\ref{eground}) \\
625: \hline
626: \rule[-3mm]{0mm}{8mm}
627: 0.005 & 1.766 107 332 563 $\times$ 10$^{-30}$ &
628: \underline{1.766 107 332 563} $\times$ 10$^{-30}$ \\
629: \rule[-3mm]{0mm}{8mm}
630: 0.010 & 5.267 473 259 637 $\times$ 10$^{-16}$ &
631: \underline{5.267 473 259 637} $\times$ 10$^{-16}$ \\
632: \rule[-3mm]{0mm}{8mm}
633: 0.015 & 3.508 587 565 372 $\times$ 10$^{-11}$ &
634: \underline{3.508 587 56}4 030 $\times$ 10$^{-11}$ \\
635: \rule[-3mm]{0mm}{8mm}
636: 0.020 & 9.033 155 571 641 $\times$ 10$^{-09}$ &
637: \underline{9.033 15}4 730 920 $\times$ 10$^{-09}$ \\
638: \rule[-3mm]{0mm}{8mm}
639: 0.025 & 2.519 767 018 258 $\times$ 10$^{-07}$ &
640: \underline{2.519 76}0 770 755(1) $\times$ 10$^{-07}$ \\
641: \rule[-3mm]{0mm}{8mm}
642: 0.030 & 2.313 302 179 961 $\times$ 10$^{-06}$ &
643: \underline{2.313} 251 574 075(2) $\times$ 10$^{-06}$ \\
644: \hline
645: \hline
646: \end{tabular}
647: \end{center}
648: \end{scriptsize}
649: \end{table}
650:
651: As seen from Eq.~(\ref{eground}), no splitting into levels with
652: positive and negative parity arises for the ground state of the
653: Fokker--Planck potential due to the linear symmetry--breaking term
654: in Eq.~(\ref{hamfp}). This term modifies the potential in such a way that the
655: leading, one-instanton ($n$ = 1) shift of the ground state energy
656: results from a back--tunneling (instanton--antiinstanton configuration)
657: of the particle to the lower well \cite{JeZJ2004plb}. For excited states,
658: in contrast, the one-instanton configuration is a trajectory which starts in
659: one well and ends in the other, restoring the broken symmetry.
660: Therefore, any excited state ($M >$ 0) of the Fokker--Planck Hamiltonian can
661: be characterized by its principal quantum number
662: %
663: %
664: \begin{equation}
665: \label{FP_M_to_K}
666: K = [\mkern - 2.5 mu [ (M+1)/2 ] \mkern - 2.5 mu ]
667: \end{equation}
668: %
669: %
670: and the parity
671: %
672: %
673: \begin{equation}
674: \label{FP_M_to_epsilon}
675: \varepsilon = 2 \left( 2 K - M - \frac12 \right) \, .
676: \end{equation}
677: %
678: %
679: In fact, this classification is very
680: similar to the double-well potential (\ref{hamdw}) except,
681: of course, the particular case of the ground state. It follows
682: naturally that the resurgent expansion for
683: the excited states of the Fokker--Planck potential is very close to the
684: analogous expansion for the double--well potential and reads
685: \cite{JeZJ2004plb}:
686: %
687: %
688: \begin{align}
689: \label{efpgen}
690: & E^{(M > 0)}_{\rm FP}(g) =
691: E^{(K, \varepsilon)}_{\rm FP}(g) =
692: \sum_{l = 0}^\infty E_{K, l} \, g^l
693: \nonumber\\[2ex]
694: & + \sum^{\infty}_{n=1} \left[-\varepsilon \, \Xi_K(g) \right]^n \,
695: \sum^{n-1}_{k=0} \left\{ \ln\left(-\frac{2}{g}\right) \right\}^k \,
696: \sum^{\infty}_{l=0} f^{(K)}_{nkl} \, g^{l} \, ,
697: \end{align}
698: %
699: %
700: where $E_{K, l}$ are perturbative coefficients and
701: $\Xi_K(g)$ is given by
702: %
703: %
704: \begin{equation}
705: \Xi_K(g) = \frac{2^{K-1/2}}{g^K\, \sqrt{ \pi \, K! \, (K-1)!}}\,
706: \rm{e}^{ - 1/6 g} \, .
707: \end{equation}
708: %
709: %
710: The power of $\Xi$ can again
711: be associated with the order of the instanton
712: ($K = 1$: one-instanton, $K = 2$ means two-instanton, etc.).
713: One should note that two intricacies are associated to
714: the precise meaning of the quantities that enter Eq.~(\ref{efpgen}):
715: %
716: \begin{itemize}
717: \item In analogy to the double-well potential, the
718: imaginary part which is generated by the resummation of the
719: perturbation series about the leading instanton
720: (the ``discontinuity'' of the distributional Borel sum
721: in the terminology of Ref.~\onlinecite{CaGrMa1986})
722: is compensated by an explicit imaginary part that stems from the
723: two-instanton effect [from the factor $\ln(-2/g)$].
724: Related questions have been discussed at length in
725: Refs.~\onlinecite{ZJJe2004i,ZJJe2004ii}.
726: %
727: \item In contrast to the ground--state energy (\ref{eground}), the leading
728: contribution to the
729: energies $E^{(\varepsilon,K > 0)}_{\rm FP}(g)$ for small coupling
730: arises from the perturbation expansion (\ref{epert}) which is manifestly
731: nonvanishing to all orders in $g$. However, since this perturbation expansion
732: is independent of the parity $\varepsilon$, the energy splitting of
733: the levels with the same principal quantum number $K$ is again dominated by the
734: one-instanton contribution ($n = 1$).
735: \end{itemize}
736: %
737: Similar to the ground state
738: (\ref{eground}), we may compute such a contribution and, hence, a
739: splitting of an arbitrary excited state $K$ by making use of
740: the $f^{(K)}_{10l}$ coefficients, which for $K > 0$ read
741: ($l = 0,1,2,3$):
742: %
743: %
744: \begin{subequations}
745: \begin{align}
746: f^{(K)}_{100} =& 1 \,, \qquad
747: f^{(K)}_{101} = -\frac{17}{2} K^2 - 6 K - \frac{5}{12} \,,
748: \\[1ex]
749: f^{(K)}_{102} =& \frac{289}{8} K^4
750: - \frac{23}{2} K^3 - \frac{1139}{24} K^2 - \frac{45}{4} K
751: - \frac{695}{288} \,,
752: \\[1ex]
753: f^{(K)}_{103} =& -\frac{4913}{48} K^6
754: + \frac{629}{2} K^5 + \frac{1637}{32} K^4 \nonumber\\[2ex]
755: & - \frac{1885}{3} K^3
756: - \frac{155825}{576} K^2 - \frac{3835}{24} K - \frac{68885}{10368} \,.
757: \end{align}
758: \end{subequations}
759: %
760: %
761: Results for $K \leq 28$ are available for download~\cite{JeHomeHD}.
762: In Table~\ref{table2}, for example, the splitting
763: $E^{(\varepsilon = -1,K = 1)}_{\rm FP}(g) -
764: E^{(\varepsilon = +1,K = 1)}_{\rm FP}(g)$ of the first
765: two excited states $M = 1,2$ due to the one-instanton effect
766: ($n$ = 1) is displayed as
767: a function of the coupling parameter $g$. Again, a comparison of the results
768: obtained by the resummation of Eq.~(\ref{efpgen}) and by the diagonalization of
769: the Fokker--Planck Hamiltonian indicates the importance of the
770: higher--instanton
771: effects ($n >$ 1) and, hence, the necessity of a double resummation of the
772: resurgent expansion, in powers of both $g$ and $\exp(-1/6g)$. Instead of
773: performing such a double summation explicitly, it is more
774: convenient to enter directly into the
775: quantization condition (\ref{qcond}), with resummed
776: quantities as defined by the
777: $A_{\rm FP}(E, g)$ and $B_{\rm FP}(E, g)$ functions. We discuss this
778: alternative approach in the next Section.
779:
780: %
781: % Table 2
782: %
783: \begin{table}[t]
784: \caption{\label{table2} Energy difference between the excited states
785: $E^{(M = 1, 2)}_{\rm FP}(g)$ of the Fokker--Planck
786: Hamiltonian. Results have been computed by
787: diagonalizing the Hamiltonian in the basis of the harmonic oscillator
788: wavefunctions (left column) and by resummation of Eq.~(\ref{efpgen}) within
789: the one-instanton approximation (right column). The numerical uncertainty is
790: estimated on the basis of the apparent
791: convergence of results under an appropriate
792: increase of the number of $f^{(K)}_{10l}$ parameters.
793: As in Table~\ref{table1},
794: we underline those decimal figures in the one-instanton results
795: which are equal to the corresponding ones in the complete
796: numerical solution.}
797: \vspace*{0.3cm}
798: \begin{scriptsize}
799: \begin{center}
800: \begin{tabular}{l@{\hspace*{0.5cm}}l@{\hspace*{0.5cm}}l}
801: \hline
802: \hline
803: \rule[-3mm]{0mm}{8mm}
804: $g$ & $E^{(M = 2)}_{\rm FP} - E^{(M = 1)}_{\rm FP}$ (diag.)
805: & $n = 1$ term of Eq.~(\ref{efpgen}) \\
806: \hline
807: \rule[-3mm]{0mm}{8mm}
808: 0.005 & 9.848 553 978 903 $\times$ 10$^{-13}$ &
809: \underline{9.848 553 978 903} $\times$ 10$^{-13}$ \\
810: \rule[-3mm]{0mm}{8mm}
811: 0.010 & 7.801 059 663 554 $\times$ 10$^{-06}$ &
812: \underline{7.801 059 65}9 99(1) $\times$ 10$^{-06}$ \\
813: \rule[-3mm]{0mm}{8mm}
814: 0.015 & 1.213 924 539 483 $\times$ 10$^{-03}$ &
815: \underline{1.213 9}1 452(2) $\times$ 10$^{-03}$ \\
816: \rule[-3mm]{0mm}{8mm}
817: 0.020 & 1.289 613 568 640 $\times$ 10$^{-02}$ &
818: \underline{1.28}8 765(1) $\times$ 10$^{-02}$ \\
819: \rule[-3mm]{0mm}{8mm}
820: 0.025 & 4.633 794 364 814 $\times$ 10$^{-02}$ &
821: \underline{4.6}11 6(1) $\times$ 10$^{-02}$ \\
822: \rule[-3mm]{0mm}{8mm}
823: 0.030 & 9.699 341 140 782 $\times$ 10$^{-02}$ &
824: \underline{9.6}10 2(6) $\times$ 10$^{-02}$ \\
825: \hline
826: \hline
827: \end{tabular}
828: \end{center}
829: \end{scriptsize}
830: \end{table}
831: %
832: %
833: %
834:
835:
836: %
837: % Resummation of the quantization condition
838: %
839: \subsection{Resummation of the quantization condition}
840: \label{resumq}
841:
842: The resurgent expansions (\ref{eground}) and (\ref{efpgen}) for the energies of
843: the ground and excited states of the Fokker--Planck Hamiltonian
844: follow as a direct consequence of the quantization condition (\ref{qcond}).
845: As seen from our calculations summarized in Tables \ref{table1} and
846: \ref{table2}, these expansions are very useful for small coupling,
847: but not of particular usefulness even for rather
848: moderate values of $g$, because of the necessity of their double resummation.
849: Here, we would like to investigate whether it is possible to
850: resum the divergent series that gives rise to $A_{\rm FP}(E, g)$ and
851: $B_{\rm FP}(E, g)$ directly and look for solutions of the
852: quantization condition (\ref{qcond}) without any
853: intermediate recourse to the resurgent
854: expansion. In fact, this approach currently
855: appears to be the only feasible way to evaluate the
856: multi--instanton expansion (in powers of $n$),
857: because the quantization condition incorporates all instanton orders.
858:
859: In order to introduce such a ``direct summation'' approach, we recall
860: that the solution of the generalized Bohr--Sommerfeld quantization condition
861: (\ref{qcond}) for a particular coupling parameter $g$ must provide the
862: energy spectrum of the Fokker--Planck Hamiltonian. In other words, if
863: one defines the left--hand side of the quantization condition
864: (\ref{qcond}) as a function of two variables $E$ and $g$:
865: %
866: \begin{align}
867: \label{Q_function}
868: Q(E, g) =& \,
869: \frac{1}{\Gamma(-B_{\rm FP}(E, g)) \, \Gamma(1-B_{\rm FP}(E, g))}
870: \nonumber\\[1ex]
871: & + \left(-\frac{2}{g}\right)^{2B_{\rm FP}(E, g)}
872: \frac{{\rm e}^{-A_{\rm FP}(E, g)}}{2 \pi} \,,
873: \end{align}
874: %
875: then the zeros of this function at fixed $g$ determine the energy spectrum of
876: (\ref{hamfp}):
877: %
878: \begin{equation}
879: \label{Q_zeros}
880: Q(E^{(M)}_{\rm FP}(g), g) \, = \, 0, \qquad M = 0, 1, 2, .... \, .
881: \end{equation}
882: %
883: A numerical analysis of the function $Q(E, g)$ can
884: be used, therefore, in order to examine
885: the validity and applicability of the generalized
886: quantization condition given by
887: Eqs.~(\ref{qcond})---(\ref{A_expansion})
888: for the case of strong coupling.
889:
890: \begin{figure}[t]
891: \begin{center}
892: \vspace*{-0.5cm}
893: \includegraphics[width=0.5\linewidth,angle=270]{Fig3.eps}
894: \end{center}
895: \vspace*{-0.7cm}
896: \caption{\label{fig3} (Color online)
897: Energy dependence of (the real part of) the
898: function $Q(E, g)$. Calculations have been performed for
899: fixed $g = 0.03$.}
900: \end{figure}
901:
902: As seen from Eq.~(\ref{Q_function}), any analysis of the function
903: $Q(E, g)$ can be traced back to the evaluation of the functions
904: $A_{\rm FP}(E, g)$ and $B_{\rm FP}(E, g)$ which constitute series in two
905: variables, namely
906: $E$ and $g$ [cf. Eqs.~(\ref{B_expansion})---(\ref{A_expansion})].
907: In order to compute these series, it is convenient to re--write the
908: functions $A_{\rm FP}(E, g) \equiv A_{\rm FP}(E, g x) |_{x=1}$ and
909: $B_{\rm FP}(E, g) \equiv B_{\rm FP}(E, g x) |_{x=1}$
910: as (formal) power series in terms of a variable $x$, taken at $x$ = 1
911: [cf. \S 8.5 of Ref.~\onlinecite{Ha1949}]:
912: %
913: \begin{eqnarray}
914: \label{a_x_series}
915: B_{\rm FP}(E, g) \, = \, \sum\limits_{n = 0}^{N_{max}}
916: \left. b^{(n)}_{\rm FP}(E, g) \, x^n \right|_{x=1}\, ,
917: \end{eqnarray}
918: %
919: %
920: \begin{eqnarray}
921: \label{b_x_series}
922: A_{\rm FP}(E, g) - \frac{1}{3 g} \, = \, \sum\limits_{n = 0}^{N_{max}}
923: \left. a^{(n)}_{\rm FP}(E, g) \, x^n \right|_{x=1}\, ,
924: \end{eqnarray}
925: %
926: %
927: where the coefficients
928: $a^{(n)}_{\rm FP}(E, g)$ and $b^{(n)}_{\rm FP}(E, g)$
929: are uniquely determined by Eqs.~(\ref{B_expansion}) and (\ref{A_expansion}):
930: $b^{(0)}_{\rm FP}(E, g) \, = \, E$,
931: $b^{(1)}_{\rm FP}(E, g) \, = \, 3E^2g$,
932: etc. In the
933: computations, the power series (\ref{a_x_series}) and (\ref{b_x_series})
934: allow one to use a unified computer algebra routine for the Borel--like
935: summations, which simply takes as input the variables $a^{(n)}_{\rm FP}(E, g)$
936: and $b^{(n)}_{\rm FP}(E, g)$, as a function of $g$ and returns the
937: value of the resummed series at $x=1$. Indeed, this routine can be universally
938: used for different values of $g$ and is therefore convenient for
939: further numerical computations which are discussed below.
940:
941: Making use of Eqs.~(\ref{a_x_series}) and (\ref{b_x_series}), we
942: may now perform a simultaneous summation of the perturbation series as
943: well as of the perturbation series about each of the instantons and find the
944: functions $B_{\rm FP}(E, g)$ and $A_{\rm FP}(E, g)$, correspondingly,
945: where we use the same notation for a function and its Borel sum.
946: There is a small subtlety because
947: for positive $g$, the power series (\ref{a_x_series}) and
948: (\ref{b_x_series}) are nonalternating and divergent and, hence, special
949: resummation techniques are required to calculate the
950: Borel sums. In our present
951: calculations, for example, we apply a generalized Borel--Pad\'e
952: method \cite{Je2000prd,JeSo2001}. The discussion of this method is
953: beyond the scope of the present work, and the reader is referred to
954: Refs.~\onlinecite{FrGrSi1985,Je2000prd,JeSo2001} for a more detailed discussion.
955: Because of their nonalternating property,
956: the perturbation series defining the functions
957: $A_{\rm FP}(E, g)$ and $B_{\rm FP}(E, g)$ are Borel summable
958: only in the distributional sense~\cite{CaGrMa1986}.
959: The evaluation of the Borel--Laplace integral thus requires an
960: integration along a contour which is tilted with respect to the real
961: axis (for details see
962: Refs.~\onlinecite{FrGrSi1985,Je2001pra,JeSo2001}
963: and the contours $C_{+1}$ and $C_{-1}$ in Ref.~\onlinecite{Je2000prd}).
964: The resummation of the divergent series
965: (\ref{a_x_series}) and (\ref{b_x_series}) may be carried out along each of
966: these contours, but it is important to characterize the perturbative and
967: instanton contributions in the same way, i.e. to deform the contours
968: for $B_{\rm FP}(E, g)$ and $A_{\rm FP}(E, g)$
969: either above or below the real axis, consistently. In the
970: terminology of Ref.~\onlinecite{CaGrMa1986}, one should
971: exclusively use either ``upper sums'' or ``lower sums,'' but
972: mixed prescriptions are forbidden. From a historical perspective, it is
973: interesting to remark that the possibility of deforming the
974: Borel integration contour had already been anticipated in
975: a remark near the end of Chap.~8 of the classic Ref.~\onlinecite{Ha1949}.
976:
977: %
978: %
979: \begin{table}[t]
980: \caption{\label{table3} Ground--state energy of the Fokker--Plank
981: Hamiltonian. Results have been computed by the diagonalization of the
982: Hamiltonian in the basis of the harmonic oscillator wavefunctions
983: (left column) and by solving Eq.~(\ref{Q_zeros}),
984: as indicated in the right column.}
985: \vspace*{0.3cm}
986: \begin{scriptsize}
987: \begin{center}
988: \begin{tabular}{l@{\hspace*{0.5cm}}l@{\hspace*{0.5cm}}l}
989: \hline
990: \hline
991: \rule[-3mm]{0mm}{8mm}
992: $g$ & $E^{(0)}_{\rm FP}$ (diagonalization)& Zero of ${\rm Re}[Q(E, g)]$ \\
993: \hline
994: \rule[-3mm]{0mm}{8mm}
995: 0.010 & 5.267 473 259 637 $\times$ 10$^{-16}$ &
996: \underline{5.267 473 259 637} $\times$ 10$^{-16}$ \\
997: \rule[-3mm]{0mm}{8mm}
998: 0.030 & 2.313 302 179 961 $\times$ 10$^{-06}$ &
999: \underline{2.313 302 17}(2) $\times$ 10$^{-06}$ \\
1000: \rule[-3mm]{0mm}{8mm}
1001: 0.070 & 1.267 755 797 982 $\times$ 10$^{-03}$ &
1002: \underline{1.267 7}4(6) $\times$ 10$^{-03}$ \\
1003: \rule[-3mm]{0mm}{8mm}
1004: 0.100 & 5.199 138 696 222 $\times$ 10$^{-03}$ &
1005: \underline{5.199} 3(2) $\times$ 10$^{-03}$ \\
1006: \rule[-3mm]{0mm}{8mm}
1007: 0.170 & 2.079 244 408 360 $\times$ 10$^{-02}$ &
1008: \underline{2.07}8(1) $\times$ 10$^{-02}$ \\
1009: \rule[-3mm]{0mm}{8mm}
1010: 0.300 & 5.318 357 438 655 $\times$ 10$^{-02}$ &
1011: \underline{5.3}23(9) $\times$ 10$^{-02}$ \\
1012: \hline
1013: \hline
1014: \end{tabular}
1015: \end{center}
1016: \end{scriptsize}
1017: \end{table}
1018: %
1019: %
1020:
1021: We are now in the position
1022: to analyze the properties of the function $Q(E, g)$ and,
1023: hence, to extract the energy spectrum of the Fokker--Planck Hamiltonian.
1024: As mentioned above, to perform such an analysis for any particular $g$ we have
1025: to (i) resum the (divergent) series for the functions $A_{\rm FP}(E, g)$ and
1026: $B_{\rm FP}(E, g)$ and (ii) insert the resulting generalized Borel sums
1027: into Eq.~(\ref{Q_function}). We may then interpret
1028: the $Q(E, g)$ as a function of
1029: $E$ (at fixed $g$) and (iii) numerically determine the zeros of this
1030: function which correspond to the energy values $E^{(M)}_{\rm FP}$ of the
1031: Fokker--Plank Hamiltonian, according to Eq.~(\ref{Q_zeros}).
1032: In Fig.~\ref{fig3}, for instance, we display the energy dependence of the
1033: real part of the function $Q(E, g)$ taken at $g = 0.03$.
1034: In the energy range $0 < E < 1$, this function has three zeros which
1035: obviously correspond to the ground $E^{(0)}_{\rm FP}$ and to the excited
1036: $E^{(1, 2)}_{\rm FP}$ states. The ground-state energy
1037: $E^{(0)}_{\rm FP}$ determined in such a way is presented in Table~\ref{table3}
1038: and compared to reference values obtained by the diagonalization of the
1039: Hamiltonian matrix in the basis of harmonic oscillator wavefunctions. Moreover,
1040: apart from the particular case of $g = 0.03$, we also display
1041: the energy $E^{(0)}_{\rm FP}$ for other coupling parameters spanning
1042: the range from $g = 0.01$ to $g = 0.3$ (see also Fig.~\ref{fig4}).
1043: This rather wide range of coupling parameters $g$ considered here allows us to
1044: investigate the behavior of the generalized Bohr--Sommerfeld quantization
1045: condition in the transition from weak to strong coupling.
1046: As seen from Table~\ref{table3}, the ground--state energy is well reproduced
1047: at $g = 0.01$ (up to 14 decimal digits).
1048: Alternatively, a highly accurate value of the ground
1049: state energy (at $g = 0.01$) can be obtained from the
1050: the one-instanton contribution to the resurgent expansion (\ref{eground})
1051: for $g < 0.01$, as indicated in Table~\ref{table1}.
1052: The accuracy of the one-instanton approximation is
1053: rapidly decreasing for higher $g$. For instance, at the moderate value of
1054: $g = 0.03$, the one-instanton term of the
1055: resurgent expansion (\ref{eground}) reproduces the ground-state
1056: energy only to four decimal digits (see the last row of
1057: Table~\ref{table1}), while a
1058: total of eight digits can be obtained from solving
1059: Eq.~(\ref{Q_zeros}), as indicated in the second row of Table~\ref{table3}.
1060: For even stronger coupling, one observes a much larger
1061: numerical uncertainty in the determination of the zeros of the
1062: function $Q(E, g)$, because the convergence of the
1063: generalized and optimized Borel--Pad\'{e} methods employed in
1064: the resummation of the
1065: $A_{\rm FP}(E, g)$ and $B_{\rm FP}(E, g)$ functions
1066: is empirically observed to reach fundamental
1067: limits for larger values of the coupling, which cannot be overcome
1068: by the use of multiprecision arithmetic and might indicate a
1069: fundamental limitation for the convergence of the transforms
1070: and are not due to numerical cancellations. It might be interesting
1071: to explore these limits also from a mathematical point of view.
1072: Specifically, we have determined the numerical uncertainty
1073: of the $A_{\rm FP}(E, g)$ and $B_{\rm FP}(E, g)$ functions
1074: on the basis of the apparent convergence of the Borel--Pad\'{e}
1075: approximants, integrated in the complex plane and accelerated
1076: according to Ref.~\onlinecite{JeSo2001}, using an optimal truncation
1077: of the order of the transforms. We found that as the order of
1078: the Borel--Pad\'{e} transformation was increased, the apparent
1079: convergence of the transforms stopped at around order 40 for
1080: $g = 0.03$ and higher. Despite these difficulties, the generalized
1081: Bohr--Sommerfeld quantization formula (\ref{Q_zeros})
1082: determines the ground--state energy of the
1083: Fokker--Planck potential with an accuracy of about 0.01 \%
1084: up to $g \leq 0.3$ (cf. Table~\ref{table3}).
1085:
1086: \begin{figure}[t]
1087: %\vspace*{-2.0cm}
1088: \begin{center}
1089: \includegraphics[height=0.9\linewidth,angle=270]{Fig4.eps}
1090: \end{center}
1091: %\vspace*{-4.5cm}
1092: \caption{\label{fig4} (Color online)
1093: Energy dependence of (the real part of) the
1094: function $Q(E, g)$. As in Fig.~\ref{fig3},
1095: the results of linear regression analysis of the
1096: function $Q(E, g)$ are depicted by the solid line. Calculations have been
1097: performed around the ``true'' energies of the ground state and for the
1098: different values of the coupling parameter: $g$ = 0.03 (left panel),
1099: $g = 0.1$ (middle panel) and $g = 0.3$ (right panel). The energy
1100: eigenvalues obtained using the displayed graphs determine the
1101: corresponding entries in the right column of Table~\ref{table3}.}
1102: \end{figure}
1103:
1104: Until now we have discussed the computation of the ground--state
1105: energy $E^{(0)}_{\rm FP}$ of the Fokker--Planck Hamiltonian. Of course, the
1106: function $Q(E, g)$ may also help to determine the energies of
1107: excited states. In contrast to the ground state, however, the
1108: numerical analysis of the function $Q(E, g)$ for excited states is more
1109: complicated due to bad convergence of the Borel sums
1110: for the $A_{\rm FP}(E, g)$ and $B_{\rm FP}(E, g)$ functions
1111: in the energy range relevant for the excited states.
1112: As seen from Fig.~\ref{fig5}, the convergence problems lead
1113: to relatively large numerical
1114: uncertainties for the numerical calculation of the function
1115: $Q(E, g)$ already for a relatively mild coupling parameter
1116: $g = 0.07$. We recall that this value of $g$ corresponds to the
1117: minimum of the energy $E_{\rm FP}^{(M = 1)}(g)$ as a
1118: function of $g$ and thus can be naturally identified
1119: as marking the transition from weak to strong coupling.
1120: As a result of the numerical
1121: uncertainties, the energy $E_{\rm FP}^{(M = 1)}(g = 0.07)$ of the
1122: first excited state may be reproduced only up to 2 decimal digits
1123: (see Table~\ref{table4}).
1124: For even larger values of parameter $g$, the maximal accuracy of
1125: calculations, based on Eq.~(\ref{Q_zeros}),
1126: is only a single significant digit, even though the
1127: double resummation of the instanton expansion, and of the
1128: perturbative expansion about each instanton, is implicitly
1129: contained in the cited Equation.
1130:
1131: \begin{figure}[t]
1132: \begin{center}
1133: %\vspace*{-1.5cm}
1134: \includegraphics[height=0.9\linewidth,angle=270]{Fig5.eps}
1135: \end{center}
1136: %\vspace*{-4.5cm}
1137: \caption{\label{fig5} (Color online)
1138: Energy dependence of (the real part of) the
1139: function $Q(E, g)$. The results for the zeroes of
1140: $Q(E, g)$ obtained by a quadratic regression
1141: analysis are depicted by solid lines. Calculations have been performed
1142: around the "true" energies of the first excited state with
1143: $K$ = 1, $\varepsilon$ = +1 ($M$ = 1)
1144: and for three different
1145: values of the coupling parameter: $g = 0.01$ (left panel),
1146: $g = 0.03$ (middle panel) and $g = 0.05$ (right panel).
1147: The energy eigenvalues obtained using the displayed graphs determine the
1148: corresponding entries in the right column of Table~\ref{table4}.}
1149: \end{figure}
1150:
1151: Supplementing the the specific Fokker--Planck energies of the $M = 0,1$ states
1152: presented in Tables~\ref{table3} and \ref{table4},
1153: we indicate in Fig.~\ref{fig6}
1154: the $g$--dependence of the energies $E^{(0)}_{\rm FP}$ and $E^{(1)}_{\rm FP}$
1155: as obtained from the analysis of the function $Q(E, g)$ and
1156: compare to reference values obtained from the
1157: diagonalization of the Hamiltonian matrix. As indicated in Fig.~\ref{fig6},
1158: the results obtained by a direct resummation of the
1159: quantization condition give a good quantitative picture
1160: of the ground--state and the first excited--state eigenvalues in the
1161: ranges $0 \le g \le 0.7$ and $0 \le g \le 0.3$, respectively.
1162: The behavior of the numerical uncertainty as a function of $g$
1163: is explicitly shown in Fig.~\ref{fig7}, where we plot the quantity
1164: %
1165: \begin{equation}
1166: \label{relative_error}
1167: \Delta(g) \, = \,
1168: \left| \frac{E_{\rm FP, \, resum}^{(M)}(g) -
1169: E_{\rm FP, \, diag}^{(M)}(g)}{E_{\rm FP, \, diag}^{(M)}(g)}
1170: \right| \, ,
1171: \end{equation}
1172: %
1173: with $E^{(M)}_{\rm FP, \, resum}(g)$ and $E_{\rm FP, \, diag}^{(M)}(g)$ being
1174: the eigenvalues as obtained from the direct
1175: resummation of the quantization condition given in
1176: Eq.~(\ref{Q_zeros}) and the
1177: diagonalization of the Hamiltonian (\ref{hamfp}),
1178: respectively (the latter values, which are numerically
1179: more accurate, are taken as the reference values). While the accuracy of
1180: resummations for the ground--state energy remains
1181: satisfactory even for (relatively) strong coupling, the relative error
1182: for the first excited is numerically much more significant.
1183:
1184: \begin{table}[t]
1185: \caption{\label{table4} The energy of the first excited state
1186: with $M=1$
1187: of the Fokker--Planck Hamiltonian. Results have been
1188: computed by the diagonalization of the Hamiltonian in the basis of the
1189: harmonic oscillator wavefunctions (exact energy) and by solving
1190: Eq.~(\ref{Q_zeros}).}
1191: \vspace*{0.3cm}
1192: \begin{scriptsize}
1193: \begin{center}
1194: \begin{tabular}{l@{\hspace*{0.5cm}}l@{\hspace*{0.5cm}}l}
1195: \hline
1196: \hline
1197: \rule[-3mm]{0mm}{8mm}
1198: $g$ & $E^{(M = 1)}_{\rm FP}$ (diagonalization) & zero of the Re[$Q(E, g)$] \\
1199: \hline
1200: \rule[-3mm]{0mm}{8mm}
1201: 0.010 & 9.677 074 461 352 $\times$ 10$^{-01}$ &
1202: \underline{9.677 074 461 352} $\times$ 10$^{-01}$ \\
1203: \rule[-3mm]{0mm}{8mm}
1204: 0.020 & 9.219 489 780 495 $\times$ 10$^{-01}$ &
1205: \underline{9.219 4}90(3) $\times$ 10$^{-01}$ \\
1206: \rule[-3mm]{0mm}{8mm}
1207: 0.030 & 8.354 795 860 905 $\times$ 10$^{-01}$ &
1208: \underline{8.354} 4(6) $\times$ 10$^{-01}$ \\
1209: \rule[-3mm]{0mm}{8mm}
1210: 0.070 & 6.828 548 309 058 $\times$ 10$^{-01}$ &
1211: \underline{6.8}33(4) $\times$ 10$^{-01}$ \\
1212: \rule[-3mm]{0mm}{8mm}
1213: 0.200 & 8.710 869 037 634 $\times$ 10$^{-01}$ &
1214: \underline{8.7}6(5) $\times$ 10$^{-01}$ \\
1215: \rule[-3mm]{0mm}{8mm}
1216: 0.250 & 9.508 936 793 119 $\times$ 10$^{-01}$ &
1217: \underline{9}.4(2) $\times$ 10$^{-01}$ \\
1218: \hline
1219: \hline
1220: \end{tabular}
1221: \end{center}
1222: \end{scriptsize}
1223: \end{table}
1224:
1225:
1226: %
1227: % Strong-coupling expansion
1228: %
1229: \subsection{Strong--coupling expansion}
1230: \label{largec}
1231:
1232: As discussed in the previous Section, a numerical procedure based on
1233: the generalized Bohr--Sommerfeld formulae may provide relatively accurate
1234: estimates of the ground as well as the (first
1235: two) excited--state energies for the
1236: coupling parameters in the range $0 \le g \le 0.3$.
1237: The question arises whether $g \approx 0.3$ can be considered as
1238: belonging to the strong coupling regime. Since the minima of the
1239: first two excited-state energies occur at $g = 0.07$ and $g = 0.025$,
1240: respectively, one might be tempted to answer the question affirmatively.
1241: However, one could devise a different
1242: criterion for the transition to the strong
1243: coupling regime. For instance, one might define the strong--coupling region
1244: as a region of an appropriately specified
1245: large-coupling asymptotic behavior of the eigenvalues
1246: $E_{\rm FP}^{(M)}(g)$.
1247:
1248: The large--coupling asymptotics of the Fokker--Planck potential thus
1249: represents a natural next aim in the current investigation.
1250: To this end, we apply a
1251: so--called Symanzik scaling $q \rightarrow g^{-1/6} q$ in Eq.~(\ref{hamfp})
1252: and rewrite the Fokker--Planck potential into another one
1253: with the same eigenvalues but a fundamentally different structure
1254: \cite{SiDi1970, We1996b}:
1255: %
1256: %
1257: \begin{eqnarray}
1258: \label{scaled_hamiltonian}
1259: H_{\rm FP} = g^{1/3}
1260: \left[ H_{\rm L} +
1261: \left( -q^3 - \frac{1}{2}\right)g^{-1/3} + \frac{q^2}{2} g^{-2/3}
1262: \right] \, ,
1263: \end{eqnarray}
1264: %
1265: %
1266: where the Hamiltonian $H_{\rm L}$ does not depend on $g$:
1267: %
1268: %
1269: \begin{eqnarray}
1270: \label{HL}
1271: H_{\rm L} = -\frac{1}{2} \left( \frac{d}{dq} \right)^2 +
1272: q + \frac{q^4}{2} \, .
1273: \end{eqnarray}
1274: %
1275: %
1276: We conclude that the $M$th eigenvalue of the Fokker--Planck Hamiltonian
1277: for the $g \rightarrow \infty$ is determined in leading order by the
1278: $M$th eigenvalue $E_{\rm L}^{(M)}$ of the Hamiltonian (\ref{HL}):
1279: %
1280: %
1281: \begin{equation}
1282: \label{leading_asymptotics}
1283: E_{\rm FP}^{(M)}(g) \approx g^{1/3} E_{\rm L}^{(M)} \, .
1284: \end{equation}
1285: %
1286: %
1287: Moreover, Eq.~(\ref{leading_asymptotics}) also indicates that
1288: the classifications of the levels $M = 0, 1, 2, \dots$ of the
1289: Fokker--Planck and the $H_{\rm L}$ Hamiltonians are obviously identical
1290: in the strong--coupling regime.
1291:
1292: \begin{figure}[t]
1293: %\vspace*{-2.0cm}
1294: \begin{center}
1295: \includegraphics[height=0.9\linewidth,angle=270]{Fig6.eps}
1296: \end{center}
1297: %\vspace*{-5.5cm}
1298: \caption{\label{fig6} (Color online) Eigenvalues $E^{(M = 0)}_{\rm FP}$ (left panel) and
1299: $E^{(M = 1)}_{\rm FP} \equiv E^{(K = 1, \varepsilon = +1)}_{\rm FP}$ (right panel)
1300: of the Fokker--Planck Hamiltonian
1301: as a functions of the coupling parameter $g$. Results have been computed by the
1302: diagonalization of the Fokker--Planck Hamiltonian in the basis of the harmonic
1303: oscillator wavefunctions (solid lines) and by solving Eq.~(\ref{Q_zeros}),
1304: where the latter correspond the the data points with the error bars.}
1305: \end{figure}
1306:
1307: Based on Eq.~(\ref{leading_asymptotics}), we now wish to compute the leading
1308: asymptotics of the ground $E_{\rm FP}^{(M = 0)}(g)$ and excited--state
1309: $E_{\rm FP}^{(M = 1, 2)}(g)$ energies of the Fokker--Planck
1310: Hamiltonian. This computation obviously requires information about the
1311: corresponding eigenvalues of the Hamiltonian $H_{\rm L}$. The
1312: energies $E_{\rm L}^{(M)}$ have again been determined by a
1313: diagonalization of the
1314: Hamiltonian matrix within a basis of up to 1000 harmonic oscillator
1315: wavefunctions and than utilized in Eq.~(\ref{leading_asymptotics}).
1316: As seen from Fig.~\ref{fig8}, the leading asymptotics of the eigenvalues
1317: $E_{\rm FP}^{(M = 0, 1, 2)}(g)$ (dotted line),
1318: calculated in such a way, significantly
1319: overestimate the energies of the Fokker--Planck Hamiltonian for the
1320: region $0 \le g \le 0.3$. Higher--order corrections
1321: to the large-coupling asymptotics are therefore required,
1322: in order to reproduce more accurately the asymptotics of the eigenvalues
1323: $E_{\rm FP}^{(M = 0, 1, 2)}(g)$. We observe that
1324: we may apply standard Rayleigh--Schr\"{o}dinger perturbation theory
1325: to Eq.~(\ref{scaled_hamiltonian}) and use the fact that
1326: the perturbative with respect to $H_{\rm L}$, which is
1327: $V(g) = \left( -q^3 - 1/2 \right)g^{-1/3} + q^2 g^{-2/3}/2$,
1328: remains Kato--bounded
1329: with respect to the unperturbed Hamiltonian $H_{\rm L}$ for large $g$.
1330: Within such an approach, a strong--coupling perturbation expansion can be
1331: written for each energy $E_{\rm FP}^{(M)}(g)$:
1332: %
1333: %
1334: \begin{equation}
1335: \label{strong_coupling_expansion}
1336: E_{\rm FP}^{(M)}(g) = g^{1/3} \sum\limits_{k = 0}^{\infty} L^{(M)}_k
1337: g^{-2k/3} \,
1338: \end{equation}
1339: %
1340: %
1341: where $L^{(M)}_0 \equiv E_{\rm L}^{(M)}$ and the higher perturbation
1342: coefficients $L^{(M)}_{k > 0}$ are calculated in the basis of the wavefunctions
1343: of the unperturbed Hamiltonian (\ref{HL}). The first six $L^{(M)}_{k}$
1344: coefficients for the ground $M$ = 0 as well as the first excited $M$ = 1, 2
1345: states are given in Table~\ref{table5}.
1346:
1347: \begin{figure}[t]
1348: \begin{center}
1349: %\vspace*{-0.3cm}
1350: \includegraphics[width=0.6\linewidth,angle=270]{Fig7.eps}
1351: \end{center}
1352: %\vspace*{-0.6cm}
1353: \caption{\label{fig7} (Color online) Relative error (\ref{relative_error}) of
1354: calculations of the eigenvalues as a function of coupling constant $g$.
1355: Results are presented for the ground ($M$ = 0) and the first excited
1356: ($M$ = 1) states of the Fokker--Planck potential.}
1357: \end{figure}
1358:
1359: In Fig.~\ref{fig8}, we implemented the first few
1360: expansion coefficients as listed in Table~\ref{table5}, in order to calculate
1361: strong-coupling asymptotics
1362: for the three lowest levels of the Fokker--Planck Hamiltonian
1363: (dashed lines). Obviously, the
1364: partial sum of the expansion (\ref{strong_coupling_expansion}) as defined by
1365: only six coefficients provides much better agreement with
1366: numerically determined energy eigenvalues
1367: than the leading asymptotics (\ref{leading_asymptotics}). For instance,
1368: the energy of the first excited state $M = 1$ is very well described by the
1369: (first six terms of the) strong--coupling expansion already at $g = 0.07$
1370: (the agreement is better than one percent).
1371: This point, which
1372: also corresponds to the minimum of the function $E_{\rm FP}^{(M = 1)}(g)$,
1373: can thus naturally be identified as the transition region
1374: between the regimes of weak and strong coupling. By using such a
1375: definition for the transitory regime, we
1376: may finally conclude that the resummation of the Bohr--Sommerfeld quantization
1377: formulae (\ref{qcond})--(\ref{A_expansion}),
1378: using the condition (\ref{Q_zeros}), may provide reasonable estimates
1379: for the energy levels $E_{\rm FP}^{(M)}(g)$ even in a limited
1380: subregion of the strong-coupling regime.
1381:
1382:
1383: %
1384: % Quantum dynamics
1385: %
1386: \subsection{Quantum dynamics}
1387: \label{quadyn}
1388:
1389: In previous Sections of the current paper,
1390: we have presented a systematic study of the energy
1391: spectrum of the Fokker--Planck potential. In particular, two methods for
1392: computing the eigenvalues $E_{\rm FP}^{(M)}(g)$
1393: have been discussed in detail: (i)
1394: a ``brute--force'' method which is based on the diagonalization of the
1395: Hamiltonian matrix and (ii) a generalized perturbative approach which
1396: accounts for the instanton effects.
1397: While, of course, the precise computation of the energy levels is
1398: a very important
1399: task, the complete description of the properties of the particular
1400: Hamiltonian also requires an access to its wavefunctions
1401: $\Phi^{(M)}_{\rm FP}(q)$. Most naturally, these wavefunctions may be obtained
1402: together with the eigenvalues $E_{\rm FP}^{(M)}(g)$ by
1403: matrix diagonalization.
1404: In our calculations, the basis of the standard harmonic
1405: oscillator wavefunction $\{ \phi_n(q) \}_{n=0}^{\infty}$ is used for
1406: the construction of a Hamiltonian matrix
1407: $\mem{\phi_n}{H_{\rm FP}}{\phi_m}$. The
1408: Fokker--Planck eigenfunctions are thus given by:
1409: %
1410: %
1411: \begin{equation}
1412: \label{FP_wavefunctions}
1413: \Phi^{(M)}_{\rm FP}(q) =
1414: \sum\limits_{n = 0}^{\infty} c^{(M)}_n \phi_n(q) \, ,
1415: \end{equation}
1416: %
1417: %
1418: where the coefficients $c^{(M)}_n$ are found by the diagonalization procedure.
1419: In Fig.~\ref{fig9} we display, for example, the wavefunctions
1420: (\ref{FP_wavefunctions}) of the ground $M$ = 0 and the first excited
1421: $M = 1, 2$ states as calculated for a coupling parameter $g = 0.05$.
1422: As is evident from Fig.~\ref{fig9}, the
1423: symmetry--breaking term of the Hamiltonian
1424: (\ref{hamfp}) leads to a ground--state wavefunction
1425: $\Phi^{(M = 0)}_{\rm FP}(q)$ (dashed line) which is neither a symmetric nor
1426: antisymmetric combination of the wavefunctions of the right and left wells,
1427: but localized in the lower well. For the first
1428: excited states, in contrast, the symmetry is partially restored, and we may
1429: attribute the wavefunctions $\Phi^{(M = 1)}_{\rm FP}(q)$ (dashed and dotted
1430: line) and $\Phi^{(M = 2)}_{\rm FP}(q)$ (dotted line) to states with
1431: positive ($\varepsilon$ = +1) and negative ($\varepsilon$ = -1) parities
1432: [see also Eqs.~(\ref{FP_M_to_K})---(\ref{FP_M_to_epsilon})].
1433:
1434: %
1435: %
1436: %\begin{widetext}
1437: \begin{center}
1438: \begin{table}
1439: \begin{center}
1440: %\begin{minipage}{8.6cm}
1441: \caption{\label{table5} $L^{(M)}_{k}$ coefficients of the strong--coupling
1442: perturbation expansion (\ref{strong_coupling_expansion}) of the eigenvalues
1443: $E_{\rm FP}^{(M = 0, 1, 2)}$ of the Fokker--Planck Hamiltonian.}
1444: \begin{tabular}{c@{\hspace{0.6cm}}r@{\hspace{0.3cm}}r@{\hspace{0.3cm}}r}
1445: \hline
1446: \hline
1447: %\rule[-3mm]{0mm}{8mm}
1448: $j$ &
1449: \multicolumn{1}{c}{$M=0$} &
1450: \multicolumn{1}{c}{$M=1$} &
1451: \multicolumn{1}{c}{$M=2$} \\
1452: \hline
1453: 0 & $0.281\,067\,805$ & $1.854\,587\,292$ & $3.686\,419\,624$ \\
1454: 1 & $-0.132\,985\,313$ & $-0.209\,853\,650$ & $-0.307\,985\,031$ \\
1455: 2 & $0.021\,367\,333$ & $0.028\,174\,214$ & $0.025\,765\,237$ \\
1456: 3 & $-0.000\,876\,935$ & $0.000\,220\,875$ & $-0.000\,099\,791$ \\
1457: 4 & $-0.000\,060\,335$ & $0.000\,031\,789$ & $-0.000\,002\,135$ \\
1458: 5 & $-0.000\,001\,557$ & $0.000\,000\,739$ & $-0.000\,000\,385$ \\
1459: \hline
1460: \hline
1461: \end{tabular}
1462: %\end{minipage}
1463: \end{center}
1464: \end{table}
1465: \end{center}
1466: %\end{widetext}
1467: %
1468: %
1469:
1470: Until now, we have only discussed the evaluation of the eigenstates
1471: and eigenenergies
1472: of the Fokker--Planck Hamiltonian. In theoretical studies of
1473: double quantum-dot nanostructures \cite{GrDiJuHa1991, Op1999, GrHa1998, TsOp2004}
1474: and of the quantum tunneling phenomena in atomic physics
1475: \cite{LiBa1990, LiBa1992}, a large number of problems arise,
1476: in which the time-propagation of some (specially prepared) wavepacket in
1477: double-well-like potentials has to be considered. The methods
1478: for such a time-propagation, such as the well-known Crank-Nicolson method,
1479: the split-operator technique, and approaches based on the Floquet
1480: formalism and many others, are discussed in detail
1481: in the literature (see, e.g.,
1482: Refs.~\onlinecite{GrDiJuHa1991, GrHa1998,CrNi1947, FlMoFe1976, PrKeKn1997}).
1483: As a supplement to our previous considerations,
1484: we will now consider the time-propagation of an initial wavepacket in a
1485: double-well-like potential, recalling the adiabatic
1486: approach as one of the most simple and best-known techniques
1487: (see, e.g., the book~\onlinecite{Te2003} and references therein)
1488: for the integration
1489: of the (time--dependent) single particle Schr\"odinger equation.
1490: Within such a technique, in which we can naturally make use of the
1491: results previously derived for the eigenstates
1492: and eigenenergies, the propagation of a wavepacket $\Psi(q, t)$ in
1493: the (time--independent) Fokker--Planck potential (\ref{hamfp}) is
1494: given by:
1495: %
1496: %
1497: \begin{equation}
1498: \label{propagation_FP}
1499: \Psi(q, t) = \sum\limits_{M=0}^{\infty}
1500: b^{(M)} \,
1501: \exp\left(-{\rm i}\, E^{(M)}_{\rm FP} t\right) \,
1502: \Phi^{(M)}_{\rm FP}(q) \,.
1503: \end{equation}
1504: %
1505: %
1506: Here the coefficients $b^{(M)} = \sprm{\Phi^{(M)}_{\rm FP}}{\Psi(t=0)}$
1507: determine the decomposition of the initial wavepacket (at $t$ = 0) in the basis
1508: of the eigenfunctions (\ref{FP_wavefunctions}).
1509:
1510: Equation~(\ref{propagation_FP}) provides an exact solution for
1511: the wavefunction $\Psi(q, t)$ at an arbitrary moment of time only
1512: in the limit of an infinitely large basis of harmonic oscillator
1513: $\{ \phi_n(q) \}_{n=0}^{\infty}$ and Fokker--Planck
1514: $\{ \Phi^{(M)}_{\rm FP}(q) \}_{M=0}^{\infty}$ wavefunctions. For computational
1515: reasons, however, summations over basis functions have to be
1516: restricted to a finite number. In our calculations,
1517: basis sets of 300--1000 wavefunctions have been applied depending on the
1518: parameters of the initial wavepacket and the coupling parameter $g$. The
1519: actual size of the basis has been chosen according to the numerical checks of
1520: the Ehrenfest theorem or the normalization of the wavepacket.
1521:
1522: \begin{figure}[t]
1523: %\vspace*{-2.0cm}
1524: \begin{center}
1525: \includegraphics[width=0.8\linewidth,angle=270]{Fig8.eps}
1526: \end{center}
1527: %\vspace*{-0.5cm}
1528: \caption{\label{fig8} (Color online)
1529: Exact values (solid line) for the ground state
1530: and the two lowest excited energy levels
1531: for the Fokker--Planck potential as a function of $g$, together with the
1532: leading asymptotics (dotted line) and the
1533: partial sum of the strong-coupling expansion (dashed line) as defined by the
1534: first six nonvanishing terms in powers of $ g^{-2k/3}$
1535: [see Eq.~(\ref{strong_coupling_expansion})], which are
1536: listed in Table~\ref{table3}. Note that the leading strong-coupling
1537: asymptotics alone cannot satisfactorily
1538: describe the true energy eigenvalues for
1539: moderate and small coupling. By contrast, the partial sum
1540: of the first six nonvanishing terms
1541: of the strong-coupling asymptotics yields numerical values
1542: which are indistinguishable from the true eigenvalues on the level
1543: of the line thickness used in the plots, down to rather small
1544: values of the coupling (dashed vs.~solid lines).}
1545: \end{figure}
1546:
1547: Within the adiabatic approximation, which is valid for
1548: slowly varying potentials~\cite{Te2003,CN}, we may divide the
1549: time evolution of the potential into small intervals $\Delta t$ and assume that
1550: for every $k$th interval the Hamiltonian is time--independent and the
1551: propagation of the wavepacket is governed again by Eq.~(\ref{propagation_FP})
1552: where, of course, eigenvalues $E^{(M)}_{\rm FP}$ and
1553: eigenfunctions $\Phi^{(M)}_{\rm FP}(q)$ should be replaced with the eigenvalues
1554: $E^{(M)}_{k}$ and eigenfunctions $\Phi^{(M)}_{k}(q)$ of the
1555: Hamiltonian $H_k \equiv H(t_k)$. We have applied this adiabatic
1556: time--propagation method,
1557: whose variations are well known from the
1558: literature \cite{HwPe1977,Ve2000,Su2003,Te2003} and which is
1559: equivalent to an exponentiation of the instantaneous
1560: Hamiltonian for each time interval $\Delta t$,
1561: to investigate the evolution of an (initially) Gaussian wavepacket in
1562: a time--dependent potential (\ref{hami}) which oscillates
1563: sinusoidally between the Fokker--Planck and the
1564: double-well cases. Since the animated results of this simulation are
1565: available for download\cite{JeHomeHD}, we just present a
1566: small series of snapshots in Fig.~\ref{fig10}.
1567: As seen from these pictures, the wavepacket, which
1568: is initially located in the right well, performs
1569: oscillations between the wells.
1570: These oscillations are controlled by the temporal change of
1571: the potential (\ref{hami}). We have checked empirically that the adiabatic
1572: approximation employed here does not represent an obstacle for an accurate
1573: time evolution in even rapidly oscillating potentials,
1574: because of the calculational efficiency of the other steps in our time
1575: propagation algorithm (notably, the diagonalization including the
1576: determination of eigenvectors can be implemented
1577: in a computationally very favorable way on modern computers).
1578: It is thus possible to perform quantum dynamical simulations
1579: in potentials which oscillate between two limiting forms with
1580: two fundamentally different characteristic ground-state configurations,
1581: each of which is governed by instantons, though in a different way.
1582: A generalization of our approach to two-dimensional potentials
1583: appears to be feasible and is currently being studied.
1584:
1585: \begin{figure}[t]
1586: %\vspace*{-1.5cm}
1587: \begin{center}
1588: \includegraphics[width=0.5\linewidth,angle=270]{Fig9.eps}
1589: \end{center}
1590: %\vspace*{-0.6cm}
1591: \caption{\label{fig9} (Color online)
1592: The wavefunctions $\Phi^{(M)}_{\rm FP}(q)$ of the ground $M$ = 0 and the first
1593: excited $M$ = 1, 2 states of the Fokker--Planck Hamiltonian calculated at
1594: coupling parameter $g$ = 0.05. The base lines for the plots of the wave
1595: functions correspond to their energies $E_{\rm FP}^{(M)}(g)$.}
1596: \end{figure}
1597:
1598: %
1599: % Conclusions
1600: %
1601: \section{Conclusions}
1602: \label{conclu}
1603:
1604: In this paper,
1605: we have investigated the Fokker--Planck [Eq.~(\ref{hamfp})]
1606: and the double-well [Eq.~(\ref{hamdw})] potential from the point of
1607: view of large-order perturbation theory (resurgent expansion and
1608: generalized quantization condition), in order to map out the regimes
1609: of validity of the instanton-related resurgent expansion for the
1610: lowest energy levels, and in order to explore the possibility of
1611: reaching the strong-coupling regime via a direct resummation
1612: of the $A_{\rm FP}$ and $B_{\rm FP}$ functions
1613: given in Eqs.~(\ref{B_expansion}) and (\ref{A_expansion}), which enter the
1614: generalized quantization condition (\ref{qcond}).
1615: The latter approach entails a complete double resummation
1616: of the resurgent expansions (\ref{eground}) and
1617: (\ref{efpgen}) both in powers of the
1618: instanton coupling $\exp(-1/6g)$ and in powers of the coupling $g$
1619: (perturbative expansion about each instanton). The quest has been
1620: to explore the applicability of resummed expansions for medium
1621: and large coupling parameters, in the transitory regime to large coupling.
1622:
1623: It is quite natural to identify the transition region for the
1624: first excited state as defined
1625: by the minimum of the energy level $E_{\rm FP}^{(M=1)}(g)$ as a
1626: function of $g$ (see the right panel of Fig.~\ref{fig6}),
1627: which occurs near $g \approx 0.07$.
1628: As is evident from Figs.~\ref{fig6} and~\ref{fig8}, it is possible to reach
1629: convergence for both the resummed secular equation (\ref{Q_zeros})
1630: as well as convergence of the strong-coupling
1631: expansion~(\ref{strong_coupling_expansion})
1632: in a somewhat restricted overlap region $0.04 \lesssim g \lesssim 0.3$.
1633: The question whether it is possible to use the resummed instanton
1634: expansion for large coupling, cannot universally be answered affirmatively,
1635: although it is reassuring to find at
1636: least a restricted region of overlap. In order to interpret
1637: the overlap, one should remember that for
1638: typical Borel summable series
1639: as they originate in various contexts in field theory,
1640: it is much easier to perform resummations at moderate and even large
1641: coupling parameters than for the instanton-related case considered
1642: here. An example is the 30-loop resummation
1643: of the anomalous dimension $\gamma$ function of six-dimensional
1644: $\phi^3$-theories and of Yukawa model theories~\cite{BrKr2000},
1645: which lead to excellent convergence for couplings as large
1646: as $g = 10$ and higher~\cite{JeSo2001}.
1647:
1648: %
1649: %
1650: \begin{figure}[t]
1651: %\vspace*{-1.5cm}
1652: \begin{center}
1653: \includegraphics[width=0.7\linewidth,angle=270]{Fig10.eps}
1654: \end{center}
1655: %\vspace*{-0.8cm}
1656: \caption{\label{fig10} (Color online)
1657: Time propagation of the (initially) Gaussian wavepacket in the time--dependent
1658: potential (\ref{hami}) with $\eta = \sin(t/60)$ which oscillates between the
1659: double--well and Fokker--Planck potentials. The base line
1660: for the plot of the wave function corresponds to its instantaneous
1661: energy.}
1662: \end{figure}
1663: %
1664: %
1665:
1666: For the latter case, it is even possible to obtain, numerically,
1667: the strong-coupling asymptotics on the basis of a resummed
1668: weak-coupling perturbation theory (see, e.g., Ref.~\onlinecite{Su2001}
1669: for a remarkable realization of this idea in an
1670: extremely nontrivial context). The general notion is that any perturbation
1671: (a potential in the case of quantum mechanics and
1672: an interaction Lagrangian in the case of field theory)
1673: determines the large-order behavior of perturbation series
1674: describing a specific physical quantity.
1675: However, the potential or interaction Lagrangian
1676: also determine the large-coupling expansion for the
1677: physical quantity under investigation.
1678: This means that there is a connection between the large-coupling expansion
1679: and the large-order behavior of the perturbation series generated
1680: in each theory, and this correspondence can be exploited in order
1681: to infer strong-coupling asymptotics even in cases where only
1682: a few perturbative terms are known~\cite{Su2001}.
1683: According to our numerical
1684: results, the corresponding calculation of strong-coupling asymptotics
1685: on the basis of only a few perturbative terms would be much more difficult in
1686: cases where instantons are present, even if like in our case,
1687: additional information is present in the form of the generalized quantization
1688: condition (\ref{qcond}).
1689:
1690: We deliberately refrain from speculating about further implications of this
1691: observation and continue with a summary of the application-oriented
1692: results gained in the current investigation.
1693: Potentials of the double-well type are important for a number
1694: of application--oriented calculations, including
1695: Josephson junction qubits~\cite{BeEtAl2003, WeEtAl2005},
1696: inversion doubling in molecular physics \cite{Hu1927,Xu2005},
1697: semiconductor double
1698: quantum dots~\cite{LoDV1998, HoEtAl2002, HuEtAl2005},
1699: as well as, in a wider context, Bose--Einstein condensates in
1700: multi--well traps~\cite{StCeMo2004, AlStCe2005, StAlCe2006, AlEtAl2005}.
1701: For the latter case,
1702: a number of theoretical works have been performed recently in
1703: order to explore the time propagation of (one particle)
1704: wavepackets in driven double--well potentials
1705: \cite{GrDiJuHa1991,Op1999,TsOp2004}. In Sec.~\ref{numer}
1706: we discuss a numerical procedure for an accurate description of
1707: energy levels and of the corresponding wave functions, which can thus be
1708: used in order to construct basis sets for an accurate quantum dynamical time
1709: evolution of wave packets in both static
1710: as well as time--dependent potentials. This well-known adiabatic technique
1711: for the integration of the (single particle) time-dependent
1712: Schr\"odinger equation is briefly recalled in Sec.~\ref{quadyn}.
1713: We illustrate this technique by a calculation for a potential
1714: which oscillates between the Fokker--Planck and the double-well
1715: cases, governed by a time-dependent
1716: interpolating parameter $\eta = \cos(\omega t)$
1717: as given in Eq.~(\ref{hami}). Similar calculations can be done for
1718: cases where the potential admits resonances as in the
1719: case of a cubic anharmonic oscillator. In this case, the
1720: method of complex scaling leads to a basis of states which can be used
1721: in order to start quantum dynamical simulations.
1722: Related work is currently in progress.
1723:
1724: %
1725: % Acknowledgments
1726: %
1727: \section{Acknowledgments}
1728:
1729: U.D.J.~acknowledges support from the Deutsche
1730: Forschungsgemeinschaft (Heisenberg program).
1731: M.L.~is grateful to Max--Planck--Institute for Nuclear Physics
1732: for the stimulating atmosphere during a guest researcher
1733: appointment, on the occasion of which the current work was
1734: completed.
1735:
1736: \appendix
1737:
1738: %
1739: % Supersymmetry and the Fokker--Planck Potential
1740: %
1741: \section{Supersymmetry and the Fokker--Planck Potential}
1742: \label{appa}
1743:
1744: This appendix is meant to provide a brief identification
1745: of the Fokker--Planck potential (\ref{hamfp}) in terms
1746: of a supersymmetric (SUSY) algebra~\cite{CoKhSu1995,KaSo1997}.
1747: We define the operators
1748: %
1749: \begin{equation}
1750: B^\pm = \frac{1}{\sqrt{2}} \,
1751: \left( W(q) \mp \frac{d}{dq} \right)
1752: \end{equation}
1753: %
1754: with $W(q) = q \, (1- \sqrt{g} \, q)$ and the SUSY Hamiltonian
1755: %
1756: \begin{equation}
1757: H_{\rm SUSY} =
1758: \left( \begin{array}{cc} B^+ B^- & 0 \\ 0 & B^- B^+ \end{array} \right) =
1759: \left( \begin{array}{cc} H_1 & 0 \\ 0 & H_2 \end{array} \right) \,.
1760: \end{equation}
1761: %
1762: with
1763: %
1764: \begin{subequations}
1765: \begin{align}
1766: H_1 =& -\frac{1}{2} \left( \frac{d}{dq} \right)^2 +
1767: \frac{1}{2} q^2 \left( 1 - \sqrt{g}q \right)^2 +
1768: \sqrt{g} q - \frac{1}{2} \,, \\
1769: %
1770: H_2 =& -\frac{1}{2} \left( \frac{d}{dq} \right)^2 +
1771: \frac{1}{2} q^2 \left( 1 - \sqrt{g}q \right)^2 -
1772: \sqrt{g} q + \frac{1}{2} \,.
1773: \end{align}
1774: \end{subequations}
1775: %
1776: Notice that $W(q)$ finds a natural interpretation as a
1777: ``superpotential'' in the sense of Refs.~\onlinecite{CoKhSu1995,KaSo1997}.
1778: The Hamiltonians $H_1$ and $H_2$ are ``superpartners.''
1779: They are related to each other by a simple reflection and
1780: translation, $q \to 1/\sqrt{g} - q$, and have the same
1781: spectra. In that sense, one may say that the Fokker--Planck
1782: potential is its own superpartner. The Fokker--Planck potential
1783: therefore constitutes
1784: a case of ``broken supersymmetry'' with zero Witten index
1785: [see, e.g., Eq.~(2.88) of Ref.~\onlinecite{KaSo1997}].
1786: The construction of the supersymmetric partner thus does not
1787: help in the analysis of the Fokker--Planck Hamiltonian. One is
1788: forced into the instanton-inspired analysis presented in the current
1789: study.
1790:
1791: As a last remark, we recall that the Fokker--Planck potential
1792: %
1793: \begin{equation}
1794: V_{\rm FP}(q) = \frac{1}{2} q^2 \left( 1 - \sqrt{g}q \right)^2 +
1795: \sqrt{g} q - \frac{1}{2} = \frac12\, \left[W^2(q) - W'(q) \right]
1796: \end{equation}
1797: %
1798: could be assumed to admit a zero eigenvalue. However,
1799: as is evident from the
1800: discussion following Eq.~(7.28) of Ref.~\onlinecite{ZJJe2004ii},
1801: the corresponding eigenfunction is not normalizable,
1802: and thus cannot be interpreted as a physical state vector.
1803: Neither the Fokker--Planck potential nor its isospectral
1804: supersymmetric partner admit a zero eigenvalue.
1805:
1806: \begin{thebibliography}{10}
1807:
1808: \bibitem{ZJJe2004i}
1809: J. Zinn-Justin and U.~D. Jentschura, Ann. Phys. (N.Y.) {\bf 313}, 197 (2004).
1810:
1811: \bibitem{ZJJe2004ii}
1812: J. Zinn-Justin and U.~D. Jentschura, Ann. Phys. (N.Y.) {\bf 313}, 269 (2004).
1813:
1814: \bibitem{JeZJ2001}
1815: U.~D. Jentschura and J. Zinn-Justin, J. Phys. A {\bf 34}, L253 (2001).
1816:
1817: \bibitem{JeZJ2004plb}
1818: U.~D. Jentschura and J. Zinn-Justin, Phys. Lett. B {\bf 596}, 138 (2004).
1819:
1820: \bibitem{HeSi1978}
1821: I.~W. Herbst and B. Simon, Phys. Rev. Lett. {\bf 41}, 67 (1978).
1822:
1823: \bibitem{LuEtAl1989}
1824: M. Luban, J.~H. Luscombe, M.~A. Reed, and D.~L. Pursey, Appl. Phys. Lett. {\bf
1825: 54}, 1997 (1989).
1826:
1827: \bibitem{LoDV1998}
1828: D. Loss and D.~P. DiVincenzo, Phys. Rev. A {\bf 57}, 120 (1998).
1829:
1830: \bibitem{HoEtAl2002}
1831: A.~W. Holleitner, R.~H. Blick, A.~K. H\"uttel, K. Eberl, and J.~P. Kotthaus,
1832: Science {\bf 297}, 70 (2002).
1833:
1834: \bibitem{HuEtAl2005}
1835: A.~K. H\"uttel, S. Ludwig, H. Lorenz, K. Eberl, and J.~P. Kotthaus, Phys. Rev.
1836: B {\bf 72}, R081310 (2005).
1837:
1838: \bibitem{GrDiJuHa1991}
1839: F. Grossmann, T. Dittrich, P. Jung, and P. H\"anggi, Phys. Rev. Lett. {\bf 67},
1840: 516 (1991).
1841:
1842: \bibitem{Op1999}
1843: L.~A. Openov, Phys. Rev. B {\bf 60}, 8798 (1999).
1844:
1845: \bibitem{JeHomeHD}
1846: See the URL http://www.mpi-hd.mpg.de/\~{}ulj.
1847:
1848: \bibitem{ZJ1984jmp}
1849: J. Zinn-Justin, J. Math. Phys. {\bf 25}, 549 (1984).
1850:
1851: \bibitem{CaGrMa1986}
1852: E. Caliceti, V. Grecchi, and M. Maioli, Commun. Math. Phys. {\bf 104}, 163
1853: (1986).
1854:
1855: \bibitem{Ha1949}
1856: G.~H. Hardy, {\em Divergent Series} (Clarendon Press, Oxford, UK, 1949).
1857:
1858: \bibitem{Je2000prd}
1859: U.~D. Jentschura, Phys. Rev. D {\bf 62}, 076001 (2000).
1860:
1861: \bibitem{JeSo2001}
1862: U.~D. Jentschura and G. Soff, J. Phys. A {\bf 34}, 1451 (2001).
1863:
1864: \bibitem{FrGrSi1985}
1865: V. Franceschini, V. Grecchi, and H.~J. Silverstone, Phys. Rev. A {\bf 32},
1866: 1338 (1985).
1867:
1868: \bibitem{Je2001pra}
1869: U.~D. Jentschura, Phys. Rev. A {\bf 64}, 013403 (2001).
1870:
1871: \bibitem{SiDi1970}
1872: B. Simon and A. Dicke, Ann. Phys. (N.Y.) {\bf 58}, 76 (1970).
1873:
1874: \bibitem{We1996b}
1875: E.~J. Weniger, Ann. Phys. (N.Y.) {\bf 246}, 133 (1996).
1876:
1877: \bibitem{GrHa1998}
1878: M. Grifoni and P. H\"anggi, Phys. Rep. {\bf 304}, 229 (1998).
1879:
1880: \bibitem{TsOp2004}
1881: A.~V. Tsukanov and L.~A. Openov, Semiconductors {\bf 38}, 91 (2004).
1882:
1883: \bibitem{LiBa1990}
1884: W.~A. Lin and L.~E. Ballentine, Phys. Rev. Lett. {\bf 65}, 2927 (1990).
1885:
1886: \bibitem{LiBa1992}
1887: W.~A. Lin and L.~E. Ballentine, Phys. Rev. A {\bf 45}, 3637 (1992).
1888:
1889: \bibitem{CrNi1947}
1890: J. Crank and P. Nicolson, Proc. Camb. Phil. Soc. {\bf 43}, 50 (1947).
1891:
1892: \bibitem{FlMoFe1976}
1893: J.~A. Fleck, J.~R. Morris, and M.~D. Feit, Appl. Phys. A {\bf 10}, 129
1894: (1992).
1895:
1896: \bibitem{PrKeKn1997}
1897: M. Protopapas, C.~H. Keitel, and P.~L. Knight, Rep. Prog. Phys. {\bf 60}, 389
1898: (1997).
1899:
1900: \bibitem{Te2003}
1901: S. Teufel, {\em Adiabatic Perturbation Theory in Quantum Dynamics --- Lecture
1902: Notes in Mathematica Vol. 1821} (Springer, Berlin, Heidelberg, New York,
1903: 2003).
1904:
1905: \bibitem{CN}
1906: We note that our adiabatic approach, for slowly varying potentials, is limited
1907: essentially by the incompleteness of the basis used in each time propagation
1908: step, while the principal restriction of the familiar Crank--Nicolson time
1909: propagation algorithm (see Ref.~\cite{CrNi1947}) lies in the truncation of
1910: the exponential of the instantaneous Hamiltonian. Of course, for the
1911: one-dimensional problems studied here, an implementation of the
1912: Crank--Nicolson scheme also leads to a feasible alternative approach to the
1913: investigation of the quantum dynamics. In any case, the continuous monitoring
1914: of the Ehrenfest theorem and of the conservation of the normalization of the
1915: wavefunction provide important cross-checks for the computational consistency
1916: of the time propagation.
1917:
1918: \bibitem{HwPe1977}
1919: J.~T. Hwang and P. Pechukas, J. Comp. Phys. {\bf 67}, 4640 (1977).
1920:
1921: \bibitem{Ve2000}
1922: E.~P. Velicheva, Phys. At. Nucl. {\bf 63}, 661 (2000).
1923:
1924: \bibitem{Su2003}
1925: A.~A. Suzko, Phys. Lett. A {\bf 308}, 267 (2003).
1926:
1927: \bibitem{BrKr2000}
1928: D. Broadhurst and D. Kreimer, Phys. Lett. B {\bf 475}, 63 (2000).
1929:
1930: \bibitem{Su2001}
1931: I.~M. Suslov, Pis'ma v. Zh. \'{E}ksp. Teor. Fiz. {\bf 74}, 211 (2001), [JETP
1932: {\bf 74}, 191 (2001)].
1933:
1934: \bibitem{BeEtAl2003}
1935: A.~J. Berkley, H. Xu, R.~C. Ramos, M.~A. Gubrud, F.~W. Strauch, P.~R. Johnson,
1936: J.~R. Anderson, A.~J. Dragt, C.~J. Lobb, and F.~C. Wellstood, Science {\bf
1937: 300}, 1548 (2003).
1938:
1939: \bibitem{WeEtAl2005}
1940: L.~F. Wei, Y.~X. Liu, and F. Nori, Phys. Rev. B {\bf 71}, 134506 (2005).
1941:
1942: \bibitem{Hu1927}
1943: F. Hund, Z. Phys. {\bf 43}, 803 (1927).
1944:
1945: \bibitem{Xu2005}
1946: C.~N. Xuan, Chem. Phys. Lett. {\bf 406}, 415 (2005).
1947:
1948: \bibitem{StCeMo2004}
1949: A.~I. Streltsov, L.~S. Cederbaum, and N. Moiseyev, Phys. Rev. A {\bf 70},
1950: 053607 (2004).
1951:
1952: \bibitem{AlStCe2005}
1953: O.~E. Alon, A.~I. Streltsov, and L.~S. Cederbaum, Phys. Lett. A {\bf 347}, 88
1954: (2005).
1955:
1956: \bibitem{StAlCe2006}
1957: A.~I. Streltsov, O.~E. Alon, and L.~S. Cederbaum, Phys. Rev. A {\bf 73},
1958: 063626 (2006).
1959:
1960: \bibitem{AlEtAl2005}
1961: M. Albiez, R. Gati, J. F\"olling, S. Hunsmann, M. Cristiani, and M.~K.
1962: Oberthaler, Phys. Rev. Lett. {\bf 95}, 010402 (2005).
1963:
1964: \bibitem{CoKhSu1995}
1965: F. Cooper, A. Khare, and U. Sukhatme, Phys. Rep. {\bf 251}, 267 (1995).
1966:
1967: \bibitem{KaSo1997}
1968: H. Kalka and G. Soff, {\em Supersymmetrie} (B. G. Teubner, Stuttgart, 1997).
1969:
1970: \end{thebibliography}
1971: \end{document}
1972:
1973: