quant-ph0210033/PR.tex
1: \documentstyle[graphicx,psfig]{oldelsart}
2: \newcommand{\leqsi}{\; {\scriptstyle {< \atop \sim}} \;}
3: \newcommand{\geqsi}{\; {\scriptstyle {> \atop \sim}} \;}
4: \newcommand{\vs}[1]{\vspace*{#1}}
5: \newcommand{\prL}{Phys.\ Rev.\ Lett.\ }
6: \newcommand{\pr}{Phys.\ Rev.\ }
7: \newcommand{\jpb}{J.\ Phys.\ B: Atom.\ Mol.\ Opt.\ Phys.}
8: \newcommand{\epl}{Europhys.\ Lett.\ }
9: \newcommand{\epj}{Eur.\ Phys. J.\ }
10: \newcommand{\ch}{CHAOS\ }
11: \newcommand{\ap}{Adv.\ Phys.\ }
12: \newcommand{\phr}{Phys.\ Rep.\ }
13: \newcommand{\amop}{Adv.\ At.\ Mol.\ Opt.\ Phys.\ }
14: \newcommand{\natw}{Die Naturwissenschaften\ }
15: \newcommand{\pla}{Physics Letters A\ }
16: \newcommand{\anp}{Ann.\ Phys.\ }
17: \newcommand{\spu}{Sov.\ Phys.\ Usp.\ }
18: \newcommand{\josab}{J.\ Opt.\ Soc.\ Am.\ B\ }
19: \newcommand{\zpd}{Z.\ Phys.\ D\ }
20: \newcommand{\apl}{Appl.\ Phys.\ }
21: \newcommand{\csf}{Chaos, Solitons and Fractals\ }
22: \newcommand{\psc}{Physica Scripta\ }
23: \newcommand{\pha}{Physica A\ }
24: \newcommand{\jpa}{J.\ Phys.\ A\ }
25: \newcommand{\sptp}{Sov.\ Phys.\ Tech.\ Phys.\ }
26: \newcommand{\jpf}{J.\ Phys.\ France\ }
27: \newcommand{\phd}{Physica D\ }
28: \newcommand{\nc}{Il Nuovo Cimento\ }
29: \newcommand{\rmp}{Rev.\ Mod.\ Phys.\ }
30: \newcommand{\jpc}{J.\ Phys.\ C\ }
31: \newcommand{\ajp}{Am.\ J.\ Phys.\ }
32: \newcommand{\jcp}{J.\ Chem.\ Phys.\ }
33: \newcommand{\zp}{Z.\ Phys.\ }
34: \def\bea{\begin{eqnarray}}
35: \def\eea{\end{eqnarray}}
36: \def\be{\begin{equation}}
37: \def\ee{\end{equation}}
38: % harmless do not remove - needed for polish names with french
39: %version of latex
40: \def\c#1{\setbox0=\hbox{#1}\ifdim\ht0=1ex\accent24 #1%
41:   \else{\ooalign{\hidewidth\char24\hidewidth\crcr\unhbox0}}\fi}
42: %\tightenlines
43: %\psdraft
44: \begin{document}
45: \date{}
46: 
47: \begin{frontmatter}
48: 
49: \title{Non-dispersive wave packets in periodically driven quantum systems
50: \vskip -1truecm}
51: 
52: \author{Andreas Buchleitner$^1$, Dominique Delande $^{1,2}$,
53:  and Jakub Zakrzewski$^3$}
54: 
55: \address{$^1$Max-Planck-Institut f\"ur Physik komplexer Systeme,
56:         Dresden, Germany }
57: \address{$^2$Laboratoire Kastler-Brossel, Tour 12, \'Etage 1, Universit\'e Pierre et
58: Marie Curie,\\
59: 4 Place Jussieu, 75005 Paris, France
60: }
61:  \address{$^3$Instytut Fizyki imienia Mariana Smoluchowskiego, Uniwersytet
62: Jagiello\'nski,
63:  Reymonta 4, PL-30-059 Krak\'ow, Poland
64: }
65: 
66: \tableofcontents
67: 
68: \begin{abstract}
69: With the exception of the harmonic oscillator, quantum wave-packets
70: usually spread as time evolves. This is due to the
71: non-linear character of the classical equations of motion which
72: makes the various components of the wave-packet evolve at various
73: frequencies. We show here that, using the nonlinear 
74: resonance between an internal frequency of a system and an external 
75: periodic driving, it is possible to overcome this spreading
76: and build non-dispersive (or non-spreading)
77: wave-packets which are well localized and follow a classical
78: periodic orbit without spreading. From the quantum mechanical
79: point of view, the non-dispersive wave-packets are time periodic
80: eigenstates of the Floquet Hamiltonian, localized in the nonlinear
81: resonance island.
82: 
83: We discuss the general mechanism which produces the non-dispersive wave-packets,
84: with emphasis on simple realization in the electronic motion
85: of a Rydberg electron driven by a microwave field.
86: We show the robustness of such wavepackets  for a model 
87: one-dimensional as well as for realistic three dimensional atoms.
88: We consider their  essential properties such as
89: the stability versus ionization, the characteristic
90: energy spectrum and long lifetimes. The requirements
91: for experiments aimed at observing such non-dispersive wave-packets are
92: also considered.
93: 
94: 
95: The analysis is extended to situations in which the driving frequency is
96: a multiple of the internal atomic frequency. Such a case allows us to discuss
97: non-dispersive states composed of several, macroscopically separated
98: wave-packets communicating among themselves by tunneling. Similarly we
99: briefly discuss other closely related phenomena in atomic and molecular
100: physics as well as possible further extensions of the theory.    
101: \end{abstract}
102: 
103: \begin{keyword}
104: wave-packet, dispersion, spreading, coherent states, Rydberg atoms,
105: non-linear resonance, atom-field interaction
106: \PACS{
107: 05.45.Mt, 03.65.Sq, 32.80.Qk, 32.80.Rm, 42.50.Hz}
108: \end{keyword}
109: 
110: \end{frontmatter}
111: 
112: 
113: \section{Introduction}
114: \subsection{What is a wave packet?}
115: \label{INTRO1}
116: It is commonly accepted that, for macroscopic systems like comets, cars, 
117: cats, and dogs \cite{englert95}, quantum objects
118: behave like classical ones. 
119: Throughout this report, we will understand by
120: ``quantum objects'' 
121: physical systems governed by the Schr\"odinger
122: equation: the system can then be entirely described by its state 
123: $|\psi\rangle$, which, mathematically speaking,
124: is just a vector in Hilbert space.
125: We will furthermore restrict ourselves to the dynamics 
126: of a single, spin-less particle, 
127: such as to have an immediate representation of $|\psi\rangle$ in configuration 
128: and momentum space by the wave functions 
129: $\langle\vec{r}|\psi\rangle =\psi(\vec{r})$ 
130: and $\langle\vec{p}|\psi\rangle =\psi(\vec{p})$, respectively. 
131: The same object is in classical mechanics described by its phase space coordinates 
132: $\vec{r}$ and $\vec{p}$ (or variants thereof), 
133: and our central concern will be to understand
134: how faithfully we can mimic the classical time evolution of $\vec{r}$ and $\vec{p}$ by a 
135: {\em single} quantum state $|\psi\rangle$, in the {\em microscopic} realm.
136: 
137: Whereas classical dynamics are described by Hamilton's equations of motion, which determine
138: the values 
139: of $\vec{r}$ and $\vec{p}$ at any time, given some initial condition $(\vec{r_0},\vec{p_0})$, 
140: the quantum evolution is described by the Schr\"odinger equation,
141: which propagates the wave function. Hence, it is suggestive to
142: associate a classical particle with a quantum state $|\psi\rangle$
143: which is optimally localized around the classical particle's phase
144: space position, at any time $t$.              
145: However, quantum mechanics imposes a fundamental limit
146: on 
147: localization, expressed by 
148: Heisenberg's uncertainty relation
149: \begin{equation}
150: \Delta z\ .\ \Delta p \ge \frac{\hbar}{2},
151: \label{heisenberg}
152: \end{equation}
153: where $\Delta z$ and $\Delta p$ are the uncertainties 
154: (i.e., square roots of the variances) of the 
155: probability distributions of $z$ and its conjugate momentum $p$ in
156: state $|\psi\rangle$, 
157: respectively 
158: (similar relations hold for 
159: other choices of canonically conjugate coordinates). Consequently, the
160: best we can hope for is a quantum state localized with a finite width
161: $(\Delta z,\Delta p)$ around the particle's classical position
162: $(z,p)$, with $\Delta z$ and $\Delta p$ much smaller than the typical
163: scales of the classical trajectory. This, however, would satisfy our
164: aim of constructing a quantum state that mimics the classical motion,
165: provided $|\psi\rangle$ keeps track of the classical time evolution
166: of $z$ and $p$, and $\Delta z$ and $\Delta p$ remain small as time
167: proceeds. After all, also classical bodies follow their center of
168: mass trajectory even if they have a finite volume. Quantum states
169: which exhibit these properties at least on a finite time scale are
170: called ``wave-packets'', simply due to their localization properties
171: in phase space \cite{schroe26}.
172: 
173: More formally, a localized solution of a wave equation like the
174: Schr\"odinger equation can be conceived as a linear superposition of
175: plane waves (eigenstates of the momentum operator) or of any other
176: suitable basis states. From a purely technical point of view, such a
177: superposition may be seen as a {\em packet of waves}, hence, a
178: wave-packet. Note, however, that any strongly localized object is a
179: wave-packet in this formal sense, though not all superpositions of
180: plane waves qualify as localized objects. In addition, this formal
181: definition quite obviously depends on the basis used for the
182: decomposition. Therefore, the only sensible definition of a
183: wave-packet can be through its localization properties in phase space,
184: as outlined above. 
185: 
186: What can we say about the localization properties of a quantum state
187: $|\psi\rangle$ as time evolves? For simplicity, let us assume that
188: the Hamiltonian describing the dynamics has the time-independent form 
189: \begin{equation}
190: H = \frac{{\vec p}^2}{2m} + V({\vec r}),
191: \end{equation} 
192: with $V({\vec r})$ 
193: some potential.
194: The 
195: time evolution of $|\psi\rangle$ is then described by the 
196: Schr\"odinger equation 
197: \begin{equation}
198: \left[ -\frac{{\hbar}^2}{2m} \Delta + V({\vec r})  \right]\psi({\vec r},t) 
199: = i \hbar \frac{\partial \psi({\vec r},t)}{\partial t}.
200: \end{equation}
201: The expectation values of position and momentum in this state are
202: given by 
203: \begin{eqnarray}
204: <{\vec r}(t)> = \langle \psi(t)|\ {\vec r}\ |\psi(t) \rangle , \\ 
205: <{\vec p}(t)> = \langle \psi(t)|\ {\vec p}\ |\psi(t) \rangle ,
206: \end{eqnarray}
207: with time evolution 
208: \begin{eqnarray}
209: &&\frac{d<{\vec r}>}{dt} = \frac{1}{i\hbar}\ <[{\vec r},H]> = 
210: \frac{<{\vec p}>}{m},\\
211: &&\frac{d<{\vec p}>}{dt} = \frac{1}{i\hbar}\ <[{\vec p},H]> = 
212: - <\nabla V({\vec r})>,
213: \label{ehrenfest1}
214: \end{eqnarray}
215: and $[.,.]$ 
216: the commutator. 
217: These 
218: are almost the classical equations of motion generated by $H$, apart
219: from the right hand side of eq.~(\ref{ehrenfest1}), and apply for {\em
220: any} $|\psi\rangle$, irrespective of its localization
221: properties. If we additionally assume $|\psi\rangle$ to be
222: localized within a spatial region where $\nabla V(\vec{r})$ is
223: essentially constant, we have $<\nabla V({\vec r})> \simeq \nabla
224: V(<{\vec r}>)$, and therefore
225: \begin{eqnarray}
226: \frac{d<{\vec r}>}{dt} & = & \frac{<{\vec p}>}{m},\\
227: \frac{d<{\vec p}>}{dt} & \simeq & - \nabla V(<{\vec r}>),
228: \label{ehrenfest2}
229: \end{eqnarray}
230: precisely identical to the classical equations of motion. This is
231: nothing but 
232: Ehrenfest's theorem and tells us that the quantum expectation values
233: of $\vec{r}$ and $\vec{p}$ of an initially localized wave-packet
234: evolve according to the classical dynamics, as long as
235: $|\psi\rangle$ remains localized within a range where $\nabla
236: V(\vec{r})$ is approximately constant. However, these equations do not
237: yet give us any clue on the time evolution of the uncertainties
238: $(\Delta z,\Delta p)$ (and of those in the remaining degrees of
239: freedom), and, consequently, neither on the time scales on which they
240: are reliable. 
241: 
242: On the other hand, given a localized wave-packet at time $t=0$, a
243: decomposition 
244: \begin{equation}
245: |\psi (t=0)\rangle = \sum_n c_n \ |\phi_n\rangle
246: \label{decompo}
247: \end{equation}
248: with coefficients 
249: \begin{equation}
250: c_n = \langle \phi_n | \psi(t=0) \rangle
251: \label{coeffs}
252: \end{equation}
253: in unperturbed energy eigenstates 
254: \begin{equation}
255: H |\phi_n\rangle = E_n |\phi_n\rangle,\ n=1,2,\ldots,
256: \end{equation}
257: tells us immediately that 
258: \begin{equation}
259: |\psi (t)\rangle =  
260: \sum_n c_n \exp \left( -i \frac{E_n t}{\hbar}\right) \ |\phi_n\rangle 
261: \label{psit}
262: \end{equation}
263: cannot be stationary, {\em except} for 
264: \begin{equation}
265: |\psi (t=0)\rangle = |\phi_j\rangle ,
266: \label{single}
267: \end{equation}
268: for some suitable $j$. 
269: 
270: The eigenstates $|\phi_j\rangle$ are typically delocalized over a 
271: large part of phase space (for example, over
272: a classical trajectory, see section~\ref{SQ}), and thus
273: are not wave-packets. There is however, an exception:
274: in the vicinity of a (stable) fixed point of the classical
275: dynamics (defined \cite{lichtenberg83} as a point in phase space 
276: where the 
277: time derivatives 
278: of positions and momenta vanish simultaneously), there exist localized 
279: eigenstates, see section~\ref{GM}.
280: 
281: For a particle moving in a one-dimensional, binding potential bounded
282: from below, there is a stable fixed
283: point at any potential minimum. 
284: The quantum mechanical ground state of this system is 
285: localized near
286: the fixed point at the global minimum of the potential and is a
287: wave-packet, though a very special one: 
288: it does
289: not evolve in time. Note that there is no need for the potential to be 
290: harmonic, any potential minimum will do.
291: The same argument can be used for a one-dimensional binding potential 
292: whose origin 
293: moves with uniform velocity.
294: The problem can be 
295: reduced to the previous one  by transforming 
296: to the moving frame where 
297: the potential is
298: stationary. 
299: Back in the laboratory frame, 
300: the ground state of the particle in the 
301: moving frame will appear
302: as a wave-packet which moves 
303: at uniform velocity. Obviously, expanding the wave-packet in a
304: stationary 
305: basis
306: in the laboratory 
307: frame will result in an awfully complicated decomposition, with
308: time-dependent coefficients, and this example clearly 
309: illustrates
310: the importance of the proper choice of the referential.\footnote{In passing, 
311: note
312: that such a situation is actually realized in particle accelerators: 
313: electromagnetic
314: fields are applied to the particles, such 
315: that these are trapped at some fixed 
316: point 
317: (preferably stable) in
318: an accelerated frame \cite{lichtenberg83}.}
319: 
320: If eq.~(\ref{single}) is not fulfilled, the initial localization of
321: $|\psi(t=0)\rangle$ (which is equivalent to an appropriate choice
322: of the $c_n$ in eq.~(\ref{coeffs})) will progressively deteriorate as
323: time evolves, the wave-packet will {\em spread}, 
324: due to the accumulation of relative phases of the
325: different contributions to the sum in eq.~(\ref{psit}). Whereas the
326: classical dynamics in a one-dimensional binding potential $V(\vec{r})$ are described
327: by periodic orbits (at any energy), the quantum dynamics are in
328: general {\em not} periodic. A return to the initial state is {\em only}
329: possible if {\em all} the phases $\exp(-i E_n t/\hbar)$ simultaneously take
330: the same value. This implies that all the energy levels $E_n$ 
331: (with $c_n\neq 0$) are
332: equally spaced, or that all the 
333: level spacings are integer multiples of some quantity.
334: In practice, this is realized only for the harmonic oscillator 
335: (in any dimension), and for tops or rotors where the Hamiltonian
336: is proportional to some component of an angular 
337: momentum variable. Another possibility is to use the linear Stark effect
338: in the hydrogen atom which produces manifolds of equally spaced energy levels.
339: However, experimental imperfections (higher order Stark effect and
340: effect of the ionic core on non-hydrogenic Rydberg atoms) break the
341: equality of the spacings and consequently lead to dispersion \cite{raman97}.
342:  
343: The equality of consecutive spacings has a simple classical interpretation: 
344: since all classical trajectories are
345: periodic with the {\em same} period, the system is exactly back in its initial state
346: after an integer number of periods. In other words, in those special cases, 
347: there is no wave-packet
348: spreading at long times.
349: 
350: However, for more generic systems, the energy levels are not equally spaced,
351: neither are the spacings simply related, and a wave-packet {\em will} spread. 
352: For a 
353: one-dimensional,
354: time-independent system, it is even possible to estimate the time after which 
355: the wave-packet
356: has significantly spread (this phenomenon is also known as the ``collapse'' 
357: of the
358: wave-packet \cite{yeazell90,alber91}). 
359: This is done by expanding the various energies $E_n$ around
360: the ``central'' energy $E_{n_0}$ of the wave-packet:
361: \begin{equation}
362: E_n \simeq E_{n_0} + (n-n_0) \frac{dE_n}{dn}(n_0) + \frac{(n-n_0)^2}{2} 
363: \frac{d^2E_n}{dn^2}(n_0).
364: \label{esec}
365: \end{equation}
366: 
367: The wave-packet being initially localized, its energy is more or less
368: well defined and only a relatively small number $\Delta n \ll n_0$
369: of the coefficients $c_n$ have significant values.
370: At short times, the contribution of the second order term in eq.~(\ref{esec})
371: to the evolution can be neglected. Within this approximation, the important energy levels
372: can be considered as equally spaced, and one obtains a periodic motion of the
373: wave-packet, with period:
374: \begin{equation}
375: T_{\rm recurrence} = \frac{2\pi \hbar}{\frac{dE_n}{dn}(n_0)}.
376: \label{trec}
377: \end{equation}
378: In the standard semiclassical 
379: WKB approximation (discussed in section \ref{wkb}) \cite{landau2}, this is
380: nothing but the classical period of the motion at energy $E_{n_0}$, 
381: and one recovers the similarity
382: between the quantum motion of the wave-packet and the classical motion of a
383: particle.
384: 
385: At longer times, the contributions of the various eigenstates to the dynamics 
386: of the
387: wave-packet will come out of phase because of the second order term in 
388: eq.~(\ref{esec}),
389: resulting in spreading and collapse of the wave-packet. A rough estimate of 
390: the collapse time
391: is thus when the relevant phases have changed by $2\pi.$
392: One obtains:
393: \begin{equation}
394: T_{\rm collapse} \simeq \frac{1}{(\Delta n)^2}\ \frac{2 \hbar}{\frac{d^2E_n}{dn^2}(n_0)}.
395: \label{tcol}
396: \end{equation}
397: Using the standard WKB approximation, one can show that this expression 
398: actually
399: corresponds to the time needed
400: for the corresponding classical phase space density to 
401: significantly
402: spread under the influence of the {\em classical} evolution. 
403: 
404: At still longer times, a pure quantum phenomenon appears, due to the 
405: discrete
406: nature of the energy spectrum. Since the $(n-n_0)^2$ factors in 
407: eq.~(\ref{esec}) are all
408: integers, the second order contributions to the
409: phase are all integer multiples of the phase of the $n-n_0=1$ term. If the 
410: latter is an
411: integer multiple of $2\pi$, {\em all} the second order contributions 
412: will rephase,
413: inducing a revival of the wave-packet in its original shape. A refined 
414: estimation of the revival time actually shows that this analysis 
415: overestimates the
416: revival time by a factor two.\footnote{It must also be noted that, at simple rational multiples 
417: (such as 1/3, 1/2, 2/3) of the revival time, one observes ``fractional revivals'' 
418: \cite{parker86,alber86,averbukh89,yeazell91},
419: where only part of the
420: various amplitudes which contribute to eq.~(\ref{psit}) rephase. This
421: generates a wave-function split
422: into several individual wave-packets, localized at different positions along 
423: the
424: classical orbit.}
425:  The correct result is
426: \cite{parker86,alber86,averbukh89,yeazell91}:
427: \begin{equation}
428: T_{\rm revival} = \frac{2\pi \hbar}{\frac{d^2E_n}{dn^2}(n_0)}.
429: \label{trev}
430: \end{equation}
431: 
432: Based on the very elementary considerations above, we can so far
433: draw the following conclusions:
434: \begin{itemize}
435: \item An initially localized wave-packet will follow the classical
436: equations of motion for a finite time $t\sim T_{\rm recurrence}$;
437: \item its localization properties {\em cannot} be stationary as time
438: evolves;
439: \item in general, the initial quasi-classical motion is followed by collapse
440: and revival, with the corresponding time scales $T_{\rm
441: recurrence}<T_{\rm collapse}<T_{\rm revival}$.
442: \end{itemize}
443: 
444: In the sequel of this report, we will show how under very general
445: conditions it is indeed possible to create wave-packets as single
446: eigenstates of quantum systems, i.e., as localized ground states in an
447: appropriately defined reference frame.
448: The most suitable framework is to consider
449: quantum evolution in classical phase space, that provides a picture
450: which is {\em independent} of the choice of the basis and allows for
451: an immediate comparison with  the classical Hamiltonian flow. In
452: addition, such a picture motivates a semiclassical interpretation,
453: which we will expand upon in sec.~\ref{SQ}. The appropriate technical
454: tool for a phase space description are quasiprobability distributions
455: \cite{hillery84} 
456: as the Wigner representation
457: of the state $|\psi(t)\rangle$,
458: \begin{equation}
459: W({\vec r},{\vec p}) = \frac{1}{(2\pi \hbar)^f}
460: \ \int \psi^*\left({\vec r}+ \frac{\vec x}{2}\right)  
461: \psi \left({\vec r}- \frac{\vec x}{2}\right)
462: \exp{\left( i \frac{{\vec x}.{\vec p}}{\hbar}\right)}
463: \ {\rm d}^f{\vec x},
464: \end{equation}
465: where $f$ is the number of degrees of freedom. 
466: 
467: The Wigner density $W({\vec r},{\vec p})$ is real, but not necessarily
468: positive \cite{hillery84,moyal49}.
469: Its  
470: time-evolution follows from the Schr\"odinger equation \cite{hillery84}:
471: \begin{equation}
472: \hbar \frac{\partial W({\vec r},{\vec p},t)}{\partial t} =
473: -2 H({\vec r},{\vec p},t) \sin \left(\frac{ \hbar \Lambda}{2}\right )\ 
474: W({\vec r},{\vec p},t),
475: \label{dwsdt}
476: \end{equation}
477: where 
478: \begin{equation}
479: \Lambda = \overleftarrow{\nabla_{\vec p}} \overrightarrow{\nabla_{\vec r}} 
480: - \overleftarrow{\nabla_{\vec r}} \overrightarrow{\nabla_{\vec p}}, 
481: \end{equation}
482: and the arrows indicate in which direction the derivatives act.
483: Eq.~(\protect\ref{dwsdt}) can serve to
484: motivate the semiclassical approach. Indeed,
485: the $\sin$ function can be expanded in a Taylor series, i.e. a power 
486: expansion in $\hbar.$ 
487: At lowest non--vanishing order, only terms linear in $\Lambda$ 
488: contribute and
489: one obtains:
490: \begin{equation}
491: \frac{\partial W({\vec r},{\vec p},t)}{\partial t} = \{H, W({\vec r},{\vec p},t)\},
492: \label{liouville}
493: \end{equation}
494: where $\{.,.\}$ denotes the classical Poisson bracket \cite{lichtenberg83}:
495: \footnote{We choose here the most common definition of the Poisson bracket.
496: Note, however, that 
497: some
498: authors \protect{\cite{landau1,haake90}} use the opposite sign!}
499: \begin{equation}
500: \{f,g\} = \sum_{i=1...f}{\frac{\partial f}{\partial r_i} 
501: \frac{\partial g}{\partial p_i} - \frac{\partial f}{\partial p_i} 
502: \frac{\partial g}{\partial r_i} }.
503: \label{poisson_brackets}
504: \end{equation}
505: Eq.~(\ref{liouville}) is nothing but 
506: the classical Liouville equation \cite{lichtenberg83}
507: which describes the  classical evolution of a phase 
508: space density.
509: Hence, in the ``semiclassical limit'' $\hbar \rightarrow 0,$ the
510: Wigner density evolves classically.
511: Corrections of higher power in $\hbar$ can be calculated systematically.
512: For example, the next order is $\hbar^3 H\Lambda^3W/24$ in eq.~(\ref{dwsdt}), 
513: and
514: generates terms which contain third order 
515: derivatives (in either position and/or
516: momentum) of the
517: Hamiltonian. Therefore, for a Hamiltonian of maximal degree two
518: in 
519: position and/or momentum, all higher order terms in eq.~(\ref{dwsdt}) vanish 
520: and the
521: Wigner distribution follows the classical evolution for an arbitrary 
522: initial
523: phase space density, 
524: and for arbitrarily long times. The harmonic oscillator 
525: is an example
526: of such a 
527: system \cite{schroe26,glauber63}, in agreement with our discussion of
528: eq.~(\ref{esec}) above.
529: 
530: 
531: Now, once again, why does a wave-packet spread? At first sight, it could be thought that
532: this is due to the higher order terms in eq.~(\ref{dwsdt}), and thus of
533: quantum origin. This is not true and spreading of a wave-packet
534: has a purely classical origin, as illustrated 
535: by the following example. Let us consider
536: a one-dimensional, free particle (i.e. no potential), initially described by a
537: Gaussian wave-function with average position $z_0$,
538: average momentum $p_0>0$, and spatial width $\sigma$:
539: \begin{equation}
540: \psi(z,t=0) = \frac{1}{\pi^{1/4} \sqrt{\sigma}}
541: \exp \left( i \frac {p_0 z}{\hbar}
542: - \frac{(z-z_0)^2}{2\sigma^2}\right ).
543: \label{wp_gaussian}
544: \end{equation}
545: The corresponding Wigner distribution is a Gaussian in phase space:
546: \begin{equation}
547: W(z,p,t=0)=\frac{1}{\pi \hbar} \ \exp \left( - \frac {(z-z_0)^2}{\sigma^2} 
548: -\frac {\sigma^2 (p-p_0)^2}{\hbar^2} \right).
549: \label{wigner_gaussian}
550: \end{equation}
551: As the Hamiltonian is quadratic in the momentum, without potential, 
552: this distribution
553: evolves precisely alike the equivalent 
554: classical phase space density. Hence, the part of the 
555: wave-packet
556: with $p>p_0$ will evolve 
557: with a 
558: larger velocity than the part with $p<p_0.$ 
559: Even if both parts are initially localized close to $z_0$, 
560: the  
561: contribution of different velocity classes 
562: implies that their distance will increase
563: without bound at long times. The wave-packet will therefore 
564: {\em spread}, because
565: the various {\em classical} trajectories have different velocities. Spreading is thus
566: a completely classical phenomenon. 
567: 
568: This can be seen quantitatively by calculating the exact quantum evolution.
569: One obtains
570: \begin{equation}
571: W(z,p,t)=\frac{1}{\pi \hbar} \exp \left( - \frac {(z-z_0-pt/m)^2}{\sigma^2} 
572: -\frac {\sigma^2 (p-p_0)^2}{\hbar^2} \right)
573: \end{equation}
574: for the Wigner distribution, and 
575: \begin{equation}
576: \psi(z,t) = \frac{1}{\pi^{1/4}}\ 
577: \frac{{\rm e}^{i\phi}}{\left(\sigma^2+\frac{\hbar^2t^2}{m^2\sigma^2}\right)^{1/4}}
578: \ \exp \left\{ i \frac{p_0 z}{\hbar}- \frac{\left(z-z_0-\frac{p_0t}{m}\right)^2}{2\sigma^2+\frac{2i\hbar t}{m}}\right\}
579: \end{equation}
580: for the wave-function 
581: (${\rm e}^{i\phi}$ is an irrelevant, complicated
582: phase factor).
583: The former is 
584: represented in fig.~\ref{classical_spreading}, together with the
585: evolution of a swarm of classical particles with an initial
586: phase space density identical to the one of the initial quantum
587: wave-packet. Since the quantum evolution follows exactly the classical one,
588: the phase space volume of the wave-packet is preserved. However, the
589: Wigner distribution is progressively stretched along the $z$ axis.
590: This results in a less and less localized wave-packet, with
591: \begin{equation}
592: \left\{
593: \begin{array}{l}
594: \displaystyle
595: \Delta z (t) = \frac{\sigma}{\sqrt{2}} \sqrt{1+\frac{\hbar^2t^2}{m^2\sigma^4}},\\
596: \displaystyle
597: \Delta p (t) = \frac{\hbar}{\sigma \sqrt{2}}.
598: \end{array}
599: \right.
600: \end{equation}
601: The product $\Delta z\Delta p,$ initially minimum ($\hbar/2)$, continuously increases
602: and localization is eventually lost.
603: 
604: \psfull
605: \begin{figure}
606: \centerline{\psfig{figure=bdzf01.ps,width=8cm}}
607: \caption{Evolution of the Wigner density of a wave-packet for a free
608: particle moving in a one-dimensional configuration space, compared 
609: to the classical 
610: evolution of a swarm
611: of classical particles with the same initial probability density. 
612: While the uncertainty in momentum does not vary
613: with time, the Wigner density stretches along the position coordinate
614: and loses its initial minimum uncertainty character. This implies
615: spreading of the wave-packet. This has a purely classical origin, as shown
616: by the classical evolution of the swarm of particles, which closely follow the
617: quantum evolution. The contour of the Wigner density is chosen to contain
618: 86\% of the probability.}
619: \label{classical_spreading}
620: \end{figure}
621: 
622: \subsection{Gaussian wave-packets -- Coherent states}
623: \label{coherent_states}
624: 
625: We have already realized above that,
626: for the harmonic oscillator, the second derivative $d^2E_n/dn^2$ in 
627: eq.~(\ref{esec}) 
628: vanishes
629: identically, and
630: a wave-packet does not spread, undergoing periodic motion. 
631: For this specific 
632: system,
633: one can define a restricted class of wave-packets, which are minimum
634: uncertainty states
635: (i.e., $\Delta z\Delta p = \hbar/2)$, and remain minimal under
636: time-evolution \cite{schroe26}. 
637: Nowadays, these states are known as ``coherent'' states of the harmonic oscillator
638: \cite{glauber63}, and are frequently employed in the analysis of simple quantum
639: systems such as the quantized electromagnetic field \cite{mandel90,cct92}.
640: They have Gaussian wave-functions, see fig.~\ref{coherent_state}(a), 
641: given by eq.~(\ref{wp_gaussian}), and characterized by an average position
642: $z_0$, an average momentum $p_0$, and a spatial width
643: \begin{equation}
644: \sigma = \sqrt{\frac{\hbar}{m\omega}},
645: \label{cohsigma}
646: \end{equation}
647: where $\omega$ is the classical eigenfrequency of the harmonic oscillator. 
648: The corresponding
649: Wigner distribution, 
650: eq.~(\ref{wigner_gaussian}), also
651: has 
652: Gaussian shape. The properties
653: of coherent states are widely discussed in the litterature, see
654: \cite{cct73,liboff80}.
655: 
656: In the ``naturally scaled'', dimensionless coordinates 
657: $z\sqrt{m\omega/\hbar}$ and $p/\sqrt{m\omega \hbar}$,
658: the classical trajectories of the harmonic oscillator 
659: are circles, and the Wigner distribution is an
660: isotropic Gaussian centered at $(z_0,p_0),$ see 
661: fig.~\ref{coherent_state}(b). 
662: Under time evolution, which follows precisely the classical
663: dynamics, its isotropic Gaussian shape is preserved.
664: 
665: \psfull
666: \begin{figure}
667: \centerline{\psfig{figure=bdzf02a.eps,width=7cm,angle=-90}
668: \psfig{figure=bdzf02b.eps,width=7cm,angle=-90}}
669: \caption{(a): Coherent state of the harmonic oscillator 
670: \protect\cite{schroe26}. The
671: probability density and the modulus of the wave-function
672: (solid line) have a Gaussian distribution. The real (dashed line)
673: and the imaginary (dotted line) part of the wave-function show, in addition, 
674: oscillations which reflect
675: the non-vanishing momentum. Whereas the envelope of the wave-function
676: preserves its shape under time evolution, the frequency of the oscillations
677: exhibits the same 
678: time dependence as the momentum of the corresponding classical particle. 
679: (b): Contour of the corresponding Wigner distribution which shows
680: localization in both position and momentum. The isovalue for the contour
681: is chosen to enclose 86\% of the total probability.} 
682: \label{coherent_state} 
683: \end{figure}
684: 
685: An important point when discussing wave-packets is to avoid the
686: confusion between localized wave-packets (as defined above) and minimum 
687: uncertainty
688: (coherent or squeezed \cite{cct92}) states. The latter are
689: just a very restricted class of localized states. They are the best ones in 
690: the sense
691: that they have optimum localization. On the other hand, as soon as dynamics 
692: is considered,
693: they have nice properties only for harmonic oscillators. In generic systems, 
694: they
695: spread exactly like other wave-packets.
696: Considering only coherent states as good semiclassical analogs of classical
697: particles is in our opinion 
698: a too formal point of view. Whether
699: the product $\Delta z\Delta p$ is exactly $\hbar/2$ or slightly
700: larger is certainly of secondary relevance for the semiclassical 
701: character of the wave-packet. What counts is that, in the semiclassical limit
702: $\hbar \rightarrow 0,$ the wave-packet is asymptotically perfectly
703: localized in all directions of phase space. This was Schr\"odinger's original 
704: concern, without reference to the actual value of $\Delta z\Delta p$ 
705: \cite{schroe26}.
706: 
707: Finally, for future applications, let us define the so called Husimi representation
708: of the quantum wave-function~\cite{husimi40}. It is the squared projection of a given
709: quantum state over a set of coherent states. 
710: Let us denote the gaussian wavefunction of eq.~(\ref{wp_gaussian}) 
711: (with $\sigma$ given by eq.~(\ref{cohsigma})) as 
712: $|{\mathrm Coh}(z_0,p_0)\rangle$. Then the
713: Husimi representation of $\psi(z)$ is defined as
714: \begin{equation}
715: {\mathrm Hus}(z,p)=\frac{1}{\pi}|\langle {\mathrm Coh}(z,p)|\psi\rangle|^2,
716: \label{husimi_def}
717: \end{equation} 
718: where the factor $1/\pi$ is due to the resolution of unity in the coherent states basis
719: \cite{cct73} and is often omitted (to confuse the reader). Alternatively,
720: the Husimi function may be looked upon as a Wigner function convoluted with
721: a Gaussian~\cite{hillery84}.
722: 
723: \subsection{A simple example: the one-dimensional hydrogen atom}
724: \label{INTRO2}
725: 
726: We now illustrate the ideas discussed in the preceding sections, using the
727: specific example of a one-dimensional hydrogen atom.
728: This object is both, representative of generic systems, and useful for
729: atomic systems to be discussed later in this paper.
730: We choose the simplest hydrogen atom: we neglect all relativistic, spin and QED
731: effects, and assume that the nucleus is infinitely massive. The Hamiltonian 
732: reads:
733: \begin{equation}
734: H = \frac{p^2}{2m} - \frac{e^2}{z},
735: \label{ham_h_1d}
736: \end{equation}
737: where $m$ is the mass of the electron, $e^2=q^2/4\pi\epsilon_0$, with $q$
738: the elementary charge, and
739: $z$ is restricted to the positive real axis. The validity of this
740: model as compared to the real 3D atom will be discussed in sec.~\ref{LIN}.
741: 
742: Here and in the rest of this paper, we will use atomic units, defined by 
743: $m$, $e^2$
744: and $\hbar$. The unit of length is
745: the Bohr radius $a_0=\hbar^2/me^2=5.2917\times
746:  10^{-11}\ {\rm m}$, the unit of time
747: is 
748: $\hbar^3/me^4 = 2.4189\times 10^{-17}\ {\rm s}$, the unit of 
749: energy is the 
750: Hartree
751: $me^4/\hbar^2 = 27.2\ {\rm eV},$ {\em twice} the ionization energy of the
752: hydrogen atom, and the unit of frequency is 
753: $me^4/2\pi\hbar^3 = 6.5796\times 10^{16}\ {\rm Hz}$ \cite{bethe77}.
754: 
755: With these premises, the energy levels are:
756: \footnote{The present analysis
757: is restricted to bound states of the atom. Continuum (i.e., scattering) 
758: states also
759: exist  but usually do not significantly contribute to the
760: wave-packet dynamics. If needed, they can be incorporated without any
761: fundamental difficulty~\cite{goldberger50,faisal87}.}
762: \begin{equation}
763: E_n= - \frac{1}{2n^2},\ \ \ \ \ \ {\rm for}\ n\geq 1.
764: \label{energy_levels}
765: \end{equation}
766: Clearly, the levels are not equally spaced, and therefore (see eq.~(\ref{esec})) 
767: any wave-packet
768: will spread. 
769: 
770: Fig.~\ref{wp_h1d} shows the evolution of a wave-packet
771: built from a linear combination of eigenstates of $H$, using a Gaussian
772: distribution of the coefficients $c_n$ in eq.~(\ref{psit}). The distribution
773: is  
774: centered at $n_0=60$, with
775: a width $\Delta n = 1.8$ for the $|c_n|^2.$ 
776: The calculation is done numerically, but is simple 
777: in the hydrogen atom since
778: all ingredients -- energy levels and eigenstates -- are known analytically. 
779: At time $t=0$, the wave-packet is localized at the outer turning point
780: (roughly at a distance $2n_0^2$ from the origin), and has 
781: zero initial momentum; its shape is roughly Gaussian.
782: After a quarter of a classical Kepler 
783: period $T_{\rm recurrence}$, it is significantly closer to the nucleus,
784: with negative velocity, following the classical trajectory. 
785: After half a period, it has reached the nucleus (it is essentially 
786: localized near
787: the origin). However, interference fringes are clearly visible: they 
788: originate from
789: the interference between the head of the wave-packet, which has already 
790: been reflected
791: off the nucleus, and its tail, which has not yet reached the nucleus.
792: After 3/4 of a period, the interference 
793: fringes have disappeared, and the wave-packet propagates
794: to the right. It has already spread significantly. After one period,
795: it is close to its initial position, but no more as well localized as 
796: initially.
797: This recurrence time is given by
798: eqs.~(\ref{trec}) and (\ref{energy_levels}):
799: \begin{equation}
800: T_{\rm recurrence} =2\pi n_0^3.
801: \label{hrecur}
802: \end{equation}
803: 
804: \psfull
805: \begin{figure}
806: \centerline{\psfig{figure=bdzf03.ps,width=14cm,angle=-90}}
807: \caption{Time evolution of an initially localized
808: wave-packet in the one-dimensional hydrogen atom,
809: eq.~(\protect\ref{ham_h_1d}). The wave-packet is constructed 
810: as a linear superposition of energy eigenstates of $H$, with a Gaussian 
811: distribution (centered at $n_0=60$, with a 
812: width $\Delta n=1.8$ of the 
813: $|c_n|^2$) of the coefficients $c_n$ in eq.~(\protect\ref{decompo}). Time $t$ is
814: measured in units of the classical Kepler period $T_{\rm recurrence}$,
815: eq.~(\protect\ref{hrecur}). Note the quasiclassical approach of the wave-packet 
816: to the nucleus, during the first half period (top left), with the
817: appearance of interference fringes as the particle is accelerated towards the
818: Coulomb center. After one period (top right) the wave-packet almost resumes
819: its initial shape at the outer turning point of the classical motion, but
820: exhibits considerable dispersion (collapse) 
821: after few  Kepler cycles (bottom
822: left). Leaving a little more time to the quantum evolution, we observe a
823: non-classical revival after approx. $20$ Kepler cycles (bottom right).
824: Recurrence, collapse and revival times are very well predicted by
825: eqs.~(\protect\ref{hrecur}), (\protect\ref{hcol}) and (\protect\ref{hreviv}).}
826: \label{wp_h1d} 
827: \end{figure}
828: 
829: After few periods, the wave-packet has considerably spread and is now
830: completely delocalized along
831: the classical trajectory. The time for the collapse of the
832: wave-packet is well predicted by eq.~(\ref{tcol}):
833: \begin{eqnarray}
834: T_{\rm collapse} & \simeq & \frac{2 n_0^4}{3(\Delta n)^2} = 
835: \frac{n_0}{3 \pi (\Delta n)^2}\times T_{\rm recurrence}\nonumber \\
836: & = & 1.96\times T_{\rm recurrence},\ 
837: {\rm for}\
838: n_0=60\ {\rm and}\ \Delta n=1.8.
839: \label{hcol}
840: \end{eqnarray}
841: Finally, after 20 periods, the
842: wave-packet revives with a shape similar to its initial state. Again, this 
843: revival
844: time is in good agreement with the theoretical prediction, 
845: eq.~(\ref{trev}):
846: \begin{eqnarray}
847: T_{\rm revival} & = & \frac{2 \pi n_0^4}{3} = \frac{n_0}{3}\times T_{\rm
848: recurrence} \nonumber \\ 
849: & = & 20\times T_{\rm recurrence},\ {\rm for}\
850: n_0=60\ {\rm and}\ \Delta n=1.8.
851: \label{hreviv}
852: \end{eqnarray}
853: At longer times, the wave-packet continues to alternate between collapses 
854: and revivals.
855: In fig.~\ref{dpdx}, we show the temporal evolution 
856: of the product $\Delta z \Delta p.$
857: It is initially close to the Heisenberg limit (minimum value) $\hbar/2$, 
858: and oscillates at the
859: frequency of the classical motion with a global increase. 
860: When the wave-packet has completely
861: spread, the uncertainty product is roughly constant, with apparently 
862: erratic fluctuations. 
863: At the revival time, the uncertainty undergoes again rather orderly
864: oscillations of a relatively large magnitute reaching, at minima,
865: values close to $\hbar$. That is a manifestation of
866:  its partial relocalization.
867: 
868: \psfull
869: \begin{figure}
870: \centerline{\psfig{figure=bdzf04.eps,width=8cm,angle=-90}}
871: \caption{Time evolution of the uncertainty product $\Delta z\Delta p$ 
872: (in units of $\hbar$) of the wave-packet shown in fig.~\protect\ref{wp_h1d}. 
873: Starting out from minimum uncertainty, $\Delta z\Delta p\simeq
874: \hbar/2$ (the Heisenberg limit, eq.~(\protect\ref{heisenberg})), the
875: wave-packet exhibits some transient spreading on the time scale of a 
876: Kepler period $T_{\rm recurrence}$, thus reflecting the classical motion
877: (compare top left of fig.~\protect\ref{wp_h1d}), collapses on a time scale of
878: few Kepler cycles (manifest in the damping of the oscillations of $\Delta
879: z\Delta p$ during the first five classical periods), shows a fractional
880: revival around $t\simeq 10\times T_{\rm recurrence}$, and a full revival at 
881: $t\simeq 20\times T_{\rm recurrence}$. Note that, nontheless, even at the full
882: revival the contrast of the oscillations of the uncertainty product is reduced
883: as compared to the initial stage of the evolution, as a consequence of 
884: higher-order corrections which are neglected in eq.~(\protect\ref{esec}).}
885: \label{dpdx} 
886: \end{figure}
887:  
888: For the three-dimensional hydrogen atom, the energy spectrum is exactly the same as
889: in one dimension. This implies that the 
890: temporal dynamics is built from exactly the same
891: frequencies; thus, the 3D dynamics is essentially the same 
892: as the 1D dynamics.\footnote{In a generic,
893: multidimensional, integrable
894: system, there are several different classical 
895: frequencies along the various degrees
896: of freedom. Hence, only partial revivals of the 
897: wave-packet at various times
898: are observed. The 3D hydrogen atom is {\em not} generic, 
899: because the three frequencies
900: are degenerate, which opens the possibility of 
901: a {\em complete} revival, {\em simultaneously}
902: along all three coordinates.}
903: Indeed, collapses and revivals of the wave-packet were also observed, under
904: various experimental conditions, in the laboratory
905: \cite{yeazell90,yeazell91,yeazell88,marmet94}.
906: Fig.~\ref{wp_3d} shows the evolution of a 
907: minimum uncertainty 
908: wave-packet of the 3D atom, initially localized
909: on a circular Kepler orbit of the electron. It is built as a linear
910: combination of circular hydrogenic states (i.e., states with
911: maximum angular and magnetic quantum numbers $L=M=n-1$), using the
912: same Gaussian distribution of the coefficients as in fig.~\ref{wp_h1d}.
913: As expected, the wave-packet 
914: spreads along the
915: circular trajectory (but not transversally to it) and eventually
916: re-establishes its
917: initial shape after $T_{\rm revival}.$ Figure~\ref{swarm_3d_disp} 
918: shows the corresponding 
919: evolution of a swarm of classical particles, 
920: for the same initial phase space density.
921: The spreading 
922: of the classical distribution and of the quantum wave-packet
923: proceeds very similarly, 
924: whereas the
925: revival is completely absent in the classical evolution, which once more
926: illustrates 
927: its purely quantum origin.
928: 
929: \psfull
930: \begin{figure}
931: \centerline{\psfig{figure=bdzf05an.ps,width=6cm,angle=0}
932: \psfig{figure=bdzf05bn.ps,width=6cm,angle=0}}
933: 
934: \smallskip
935: 
936: \centerline{\psfig{figure=bdzf05cn.ps,width=6cm,angle=0}
937: \psfig{figure=bdzf05dn.ps,width=6cm,angle=0}}
938: 
939: \smallskip
940: 
941: \centerline{\psfig{figure=bdzf05en.ps,width=5cm,angle=0}}
942: \caption{Time evolution of a wave-packet launched along a
943: circular Kepler trajectory, with the same Gaussian weights $c_n$,
944: eq.~(\protect\ref{coeffs}), as employed for the one-dimensional 
945: example displayed
946: in fig.~\protect\ref{wp_h1d}, i.e., centered around the principal
947: quantum number $n_0=60$.
948: Since the relative
949: phases accumulated during the time evolution only depend on $n_0$ -- see
950: eq.~(\protect\ref{psit}) -- we observe precisely the same behaviour as in the
951: one-dimensional case: classical propagation 
952: at short times (top), followed by spreading and collapse
953: (middle), and revival (bottom). The snapshots of the
954: wave function are taken at times (in units of $T_{\rm recurrence}$) 
955: $t=0$ (top left), $t=0.5$ (top right), 
956: $t=1$ (middle left), $t=12.5$ (middle right), and $t=19.45$ (bottom).
957: The cube size is 10000 Bohr radii, centered on the nucleus (marked with 
958: a cross). The radius of the circular wave-packet trajectory 
959: equals approx.
960: 3600 Bohr radii.}
961: \label{wp_3d} 
962: \end{figure}
963: 
964: \psfull
965: \begin{figure}
966: \centerline{\psfig{figure=bdzf06a.eps,width=6cm,angle=-90}
967: \psfig{figure=bdzf06b.eps,width=6cm,angle=-90}}
968: \caption{Classical time evolution of a Gaussian (in spherical coordinates) 
969: phase space density fitted to the minimum uncertainty wave-packet of
970: fig.~\protect\ref{wp_3d} at time $t=0$ (left). As time evolves, the classical 
971: phase space density spreads along the circular Kepler orbit ($t=12.5$, right),
972: but exhibits no revival. Hence, wave-packet spreading is of purely classical
973: origin, 
974: only the revival is 
975: a quantum feature.  
976: The cube size is 10000 Bohr radii, centered on the nucleus (marked with 
977: a cross).  The radius of the circular wave-packet trajectory 
978: equals approx.
979: 3600 Bohr radii.
980: }
981: \label{swarm_3d_disp} 
982: \end{figure}
983: 
984: Finally, let us notice that collapse and revival of a 3D wave-packet 
985: depend
986: on the principal quantum number $n_0$ only -- see eqs.~(\ref{hcol}) and 
987: (\ref{hreviv}) -- and
988: are independent
989: of other parameters which characterize 
990: the classical motion, such as the eccentricity and the orientation
991: of the classical elliptical trajectory. This establishes that a 3D 
992: wave-packet 
993: with low average
994: angular momentum (and, a fortiori, a 1D wave-packet as shown in
995: fig.~\ref{wp_h1d}) -- which deeply explores the non-linearity of the 
996: Coulomb force -- does not
997: disperse faster than a circular wave-packet which essentially feels
998: a constant force. Hence, arguments on the non-linear character
999: of the interaction should be used 
1000: with some caution.
1001: 
1002: There have been several experimental realizations of electronic wave-packets
1003: in atoms \cite{raman97,yeazell90,yeazell91,yeazell88,mallalieu94,weinacht99},
1004: either along the pure radial coordinate or even along angular coordinates too.
1005: However, all these wave-packets dispersed rather quickly. 
1006: 
1007: \subsection{How to overcome dispersion}
1008: \label{INTRO3}
1009: 
1010: 
1011: Soon after the discovery of quantum mechanics, the spreading of 
1012: wave-packets was realized and attempts were made to overcome 
1013: it~\cite{schroe26}.
1014: From eq.~(\ref{psit}), it is however clear that this is only possible
1015: if the populated energy levels are equally spaced. In practice, this condition
1016: is 
1017: only met for the harmonic oscillator (or simple tops and rotors). In 
1018: any other system, the anharmonicity of the
1019: energy ladder will 
1020: induce dispersion. Hence, the situation
1021: seems hopeless.
1022:  
1023: Surprisingly, it is
1024: classical 
1025: mechanics which provides us with a possible solution. Indeed, as discussed
1026: above, a quantum wave-packet spreads exactly as the corresponding
1027: swarm of classical particles. Hence, dispersion can be overcome if all 
1028: classical trajectories
1029: behave similarly in the long time limit. In other words, if an initial 
1030: volume of phase
1031: space remains well localized under time evolution, it is reasonable to expect
1032: that a wave-packet built on this initial volume will not spread either. 
1033: The simplest
1034: example is to consider a stable fixed point: 
1035: by definition \cite{lichtenberg83}, 
1036: every initial condition in its vicinity
1037: will forever remain close to it. The corresponding wave-packet indeed does 
1038: not spread
1039: at long times \ldots though
1040: this is of limited interest, as it is 
1041: simply at rest!
1042: 
1043: Another possibility is to use a set of classical trajectories which all
1044: exhibit the
1045: same periodic motion, {\it with the same period for all trajectories}. This
1046: condition, however, is too restrictive, since it leads us back 
1047: to the harmonic oscillator.
1048: Though, we 
1049: can slightly relax this constraint by allowing classical trajectories
1050: which are not strictly periodic but quasi-periodic and staying forever
1051: in the vicinity of a well defined periodic orbit: A wave-packet built
1052: on such orbits should evolve along the classical periodic orbit
1053: while keeping a finite dispersion around it.
1054: 
1055: It happens that there is a simple possibility 
1056: to generate such 
1057: classical trajectories {\em locked} on a periodic orbit, which is to
1058: drive the system by an external periodic driving. The general 
1059: theory of nonlinear dynamical systems (described in section \ref{CD}) 
1060: \cite{lichtenberg83,ottb} shows 
1061: that when a nonlinear system (the internal frequency of which depends on 
1062: the initial conditions) is subject to an external
1063: periodic driving, a {\em phase locking} phenomenon -- known as
1064: a nonlinear resonance -- takes place. For initial conditions where the
1065: internal frequency is close to the driving frequency (quasi-resonant
1066: trajectories), the effect of the coupling is to force the motion towards the
1067: external frequency. In other words, trajectories which, in the absence
1068: of the coupling, would oscillate at a frequency slightly lower than
1069: the driving are pushed forward by the nonlinear coupling, and 
1070: trajectories
1071: with slightly larger frequency are pulled backward. In a certain region
1072: of phase space -- 
1073: termed 
1074: ``nonlinear resonance island'' -- all trajectories
1075: are trapped, and locked on the external driving. At the center of the resonance
1076: island, there is a stable periodic orbit 
1077: which precisely evolves with the driving frequency.
1078: If the driving is a small perturbation, this periodic orbit
1079: is just the periodic orbit which, in the absence of driving, has exactly
1080: the driving frequency. All the trajectories in the resonance island are
1081: winding around the central orbit with their phases locked on the
1082: external driving. The crucial point for our purposes 
1083: is that the resonance island occupies
1084: a finite volume of phase space, i.e., it traps all trajectories in a window
1085: of internal frequencies centered around the driving frequency. The size of this
1086: frequency window {\em increases} 
1087: with the amplitude of the system-driving coupling, and, as we shall see in
1088: section~\ref{GM}, can be made large enough to support wave-packet
1089: eigenstates of the corresponding quantum system.
1090: 
1091: The classical trapping mechanism is illustrated
1092: in fig.~\ref{wp_clas} which shows a swarm of classical particles launched
1093: along a circular Kepler orbit of a
1094: three-dimensional hydrogen atom exposed to a resonant, circularly polarized
1095: microwave field: the effect of the microwave field is to lock the particles
1096: in the vicinity of a circular trajectory. Note that also the phase along
1097: the classical circular trajectory is locked: the particles
1098: are grouped in the direction of the microwave field and follow
1099: its circular motion without any drift. There is a striking difference
1100: with the situation shown previously in fig.~\ref{swarm_3d_disp}, where the
1101: cloud of particles rapidly spreads in the absence of the microwave field
1102: (the same swarm of initial conditions is used in the two figures). In
1103: figure \ref{wp_clas}, there
1104: are few particles (about 10\%) in the swarm which are not phase 
1105: locked with the microwave field. This is due to the finite 
1106: subvolume of phase space which is effectively phase locked. 
1107: Particles in the tail
1108: of the initial Gaussian distribution may not be trapped 
1109: \cite{lee97}. 
1110: 
1111: 
1112: \psfull
1113: \begin{figure}
1114: \centerline{\psfig{figure=bdzf07.eps,width=6cm,angle=-90}}
1115: \caption{The initially Gaussian distributed swarm of classical particles,
1116: shown in fig.~\protect\ref{swarm_3d_disp}, after evolution
1117: in the presence of the Coulomb field of the nucleus, with a
1118: resonant, circularly polarized microwave
1119: field in the plane of the circular Kepler trajectory added. 
1120: The nonlinear resonance between the unperturbed Kepler motion and
1121: the driving field locks the phase of the particles on the phase of the driving
1122: field. As opposed to the free (classical) evolution depicted in 
1123: fig.~\protect\ref{swarm_3d_disp}, 
1124: the classical distribution does {\em not} exhibit
1125: dispersion along the orbit, except for few particles launched from the
1126: tail of the initial Gaussian distribution, 
1127: which are not trapped by the principal resonance island.}
1128: \label{wp_clas} 
1129: \end{figure}
1130: 
1131: Although
1132: the microwave field applied in fig.~\ref{wp_clas} 
1133: amounts to less than 5\% of the Coulomb field
1134: along the classical trajectory, it is sufficient to
1135: synchronize the classical motion. The same phenomenon 
1136: carries over to quantum mechanics, and 
1137: allows the creation of non-dispersive wave-packets, as will
1138: be explained in detail in section \ref{QD}.
1139: 
1140: \subsection{The interest of non-dispersive wave-packets}
1141: \label{INTRO4}
1142: 
1143: Schr\"odinger dreamt of the possibility of building quantum wave-packets 
1144: following classical trajectories \cite{schroe26}. He succeded for the harmonic 
1145: oscillator, but failed for other systems \cite{schroe68}. 
1146: It was then believed that
1147: wave-packets must spread if the system is nonlinear, and this is
1148: correct for time-independent systems.
1149: However, this is not true {\em in general}, and we have seen in the
1150: previous section that clever use of the non-linearity may, 
1151: on the contrary, {\em stabilize} a wave-packet and preserve it
1152: from spreading. Such non-dispersive wave-packets are thus
1153: a realization of Schr\"odinger's dream. 
1154: 
1155: One has to emphasize strongly that they are {\em not} some variant
1156: of the coherent states of the harmonic oscillator. They are
1157: of intrinsically completely different origin. Paradoxically,
1158: they exist only if there is some non-linearity, i.e. some
1159: unharmonicity,  in the classical system. They have
1160: some resemblance with classical solitons which are localized
1161: solutions of a non-linear equation that propagate without
1162: spreading. However, they are {\em not} solitons, as they are
1163: solutions of the {\em linear} Schr\"odinger equation. 
1164: They are simply new objects. 
1165: 
1166: Non-dispersive wave-packets in atomic systems were identified
1167: for the hydrogen atom exposed
1168: to a linearly polarized \cite{delande94,abu95d} and 
1169: circularly polarized \cite{ibb94}  microwave fields quite independently
1170: and using different physical pictures. The former approach associated
1171: the wave-packets with single Floquet states localized in the vicinity of
1172: the periodic orbit corresponding to atom-microwave nonlinear resonance.
1173: The latter treatment relied on the fact  that a transformation
1174: to a frame corotating with the microwave field removes
1175: the explicit time-dependence of the Hamiltonian for the
1176: circular polarization (see section~\ref{CP}). The states localized 
1177: near the equilibria of the rotating system
1178: were baptized ``Trojan wave-packets" to stress the analogy
1179: of the stability mechanism with Trojan asteroids. Such an approach
1180: is, however, 
1181: restricted to a narrow class of systems where the time-dependence
1182: can be removed and lacks the identification of the non-linear
1183: resonance as the relevant mechanism. We thus prefer to use
1184: in this review the more general term ``non-dispersive wave-packets"
1185: noting also that in several other papers ``non-spreading wave-packets''
1186: appear equally often.  
1187: 
1188: 
1189: Apart from their possible practical applications (for example,
1190: for the purpose of quantum control of atomic or molecular fragmentation 
1191: processes
1192: \cite{weinacht99}, or for information 
1193: storage \cite{ahn00,meyer00,kwiat00,bucksbaum00} in a confined volume 
1194: of (phase) space
1195: for long times), they show the fruitful character of classical
1196: nonlinear dynamics. Indeed, here the nonlinearity is not 
1197: a nuisance to be minimized, but rather 
1198: the essential ingredient. From complex nonlinear dynamics,
1199: a simple object is born. The existence of such non-dispersive wave-packets
1200: it extremely difficult to understand (let alone to predict) 
1201: from quantum mechanics
1202: and the Schr\"odinger equation alone. The classical nonlinear dynamics
1203: point of view is by far more illuminating and predictive. It is 
1204: the classical mechanics inside which led Berman and Zaslavsky \cite{berman77}
1205: to the pioneering discussion of states associated with the classical
1206: resonance island, the subsequent studies \cite{henkel92,holthaus94,holthaus95}
1207: further identified such states for driven one-dimensional systems using the
1208: Mathieu approach without, however, discussing the wave-packet aspects of
1209: the states. 
1210: The best proof of the importance of the classical mechanics inside
1211: is that the non-dispersive wave-packets  
1212: could have been discovered for a very long
1213: time (immediately after the formulation of the Schr\"odinger equation),
1214: but were actually identified only during the last ten years 
1215: \cite{abuth,delande94,abu95d,ibb94,farrelly95a,ibb95,farrelly95,kalinski95a,kalinski95b,delande95,kuba95a,ibb96,eberly96,kalinski96a,brunello96,kalinski96b,kalinski97,kalinski98,ibb97a,ibb97b,delande97,kuba97a,cerjan97,kuba97b,abu96,kuba98,kuba98a,abu98a,delande98,sachath,hornbergerda,hornberger98,sacha98a,sacha98b,sacha99a,yeazell00}, after
1216: the recent major developments of nonlinear dynamics. 
1217: 
1218: 
1219: 
1220: \section{Semiclassical quantization}
1221: \label{SQ}
1222: 
1223: In this section, we briefly recall the basic results
1224: on the semiclassical quantization of Hamiltonian systems, 
1225: which we will 
1226: need
1227: for
1228: the construction of non-spreading wave-packets in classical phase space, as well as
1229: to 
1230: understand
1231: their properties. This section does not contain
1232: any original material.
1233: 
1234: \subsection{WKB quantization}
1235: 
1236: \label{wkb}
1237: 
1238: For a one-dimensional, bounded, time-independent system, the Hamilton function
1239: (equivalent to the total energy)
1240: is a classical constant of motion, 
1241: and the  
1242: dynamics are periodic.
1243: It is possible to define canonically conjugate action-angle variables
1244: $(I,\theta)$, such that the Hamilton function depends 
1245: on the action 
1246: alone. The usual definition of the action along a periodic orbit (p.o.)
1247: writes:
1248: \begin{equation}
1249: I=\frac{1}{2\pi}\ \oint_{\mathrm p.o.}{p\ dz},
1250: \label{action_def}
1251: \end{equation}
1252: where $p$ is the momentum along the trajectory. 
1253: 
1254: The WKB (for Wentzel, Kramers, and Brillouin) method \cite{Berry_WKB}
1255: allows to construct 
1256: an approximate solution of the Schr\"odinger
1257: equation, in terms of the classical action-angle variables and of
1258: Planck's constant $\hbar:$
1259: \begin{equation}
1260: \psi(z) = \frac{1}{\sqrt{p}}\ \exp{\left(\frac{i}{\hbar}
1261: \int{p\ dz}\right)}
1262: \label{psi_wkb}
1263: \end{equation}
1264: as an integral along the classical trajectory.
1265: This construction is possible
1266: if and only if the phase accumulated along a period of the orbit is an integer
1267: multiple of $2\pi$. This means that  the quantized states are those where
1268: the action variable $I$ is an integer multiple of $\hbar.$ This simple
1269: picture has to be slightly amended because the semiclassical
1270: WKB approximation for the wave-function breaks down at the turning points
1271: of the classical motion,
1272: where the velocity of the classical particle 
1273: vanishes and, consequently, the
1274: expression~(\ref{psi_wkb}) diverges. This failure can be repaired \cite{Berry_WKB}
1275: by adding an additional phase $\pi/2$ for each turning point. This
1276: leads to the final quantization condition
1277: \begin{equation}
1278: I=\frac{1}{2\pi}\ \oint_{\mathrm p.o.}{p\ dz} = 
1279: \left(n+\frac{\mu}{4}\right) \hbar,
1280: \label{WKB}
1281: \end{equation}
1282: where $n$ is a non-negative integer and $\mu$ -- the ``Maslov index'' --
1283: counts the number of turning points along the periodic orbit
1284: ($\mu =2$ for a simple 1D periodic orbit).
1285: 
1286: Thus, the WKB recipe is extremely simple: when the classical Hamilton
1287: function $H(I)$
1288: is expressed in terms of the action $I$, the semiclassical energy
1289: levels are obtained by calculating $H(I)$ for the quantized values of 
1290: $I$:
1291: \begin{equation}
1292: E_n = H\left(I=\left(n+\frac{\mu}{4}\right)\hbar\right).
1293: \label{semen}
1294: \end{equation}
1295: 
1296: Finally, as a consequence of eqs.~(\ref{WKB}),(\ref{semen}), 
1297: the spacing between two
1298: consecutive semiclassical energy 
1299: levels is simply related to the classical frequency $\Omega$ of the motion,
1300: \begin{equation}
1301: E_{n+1}-E_n \simeq \frac{d E_n}{d n} = \hbar \frac{d H}{d I} = \hbar
1302: \Omega(I),
1303: \label{cprinc}
1304: \end{equation}
1305: a result which 
1306: establishes the immediate correspondence between a resonant 
1307: transition between two quantum mechanical eigenstates in the semiclassical
1308: regime, and resonant driving of the associated classical trajectory.
1309: 
1310: In the vicinity of a fixed (equilibrium) point, the Hamilton function 
1311: can be expanded
1312: at second order (the first order terms
1313: are zero, by definition of the fixed point), 
1314: leading to the ``harmonic approximation''. If the fixed point is
1315: stable, the semiclassical WKB quantization of the harmonic approximation gives 
1316: exactly the quantum result, although the semiclassical wave-function,
1317: eq.~(\ref{psi_wkb}), is incorrect. This remarkable feature is not true
1318: for an unstable fixed point (where classical trajectories escape far
1319: from the fixed point), and the WKB approximation fails in this case.    
1320: 
1321: \subsection{EBK quantization}
1322: \label{EBKsec}
1323: 
1324: For a multi-dimensional system, 
1325: it is a much more complicated task to extract the quantum mechanical 
1326: eigenenergies 
1327: from the classical dynamics of a Hamiltonian system.
1328: The problem can be solved for {\em integrable\/} systems,
1329: where there are as many constants of motion as degrees of 
1330: freedom~\cite{ozorio88}.
1331: This is known as EBK (for Einstein, Brillouin, and Keller) quantization
1332: \cite{einstein17}, and is a simple extension
1333: of the WKB quantization 
1334: scheme.
1335: Let us choose two degrees of freedom for simplicity, the extension
1336: to higher dimensions being straightforward. If the system
1337: is integrable, the Liouville-Arnold theorem~\cite{lichtenberg83}
1338: assures the existence of two pairs of canonically conjugate
1339: action-angle variables, $(I_1,\theta_1)$ and $(I_2,\theta_2)$, 
1340: such that the classical Hamilton function depends only on the actions:
1341: \begin{equation}
1342: H = H(I_1,I_2).
1343: \end{equation}
1344: The classical motion is periodic along each angle (the actions being constants
1345: of the motion)
1346: with frequencies
1347: \begin{eqnarray}
1348: \Omega_1 & = & \frac{\partial H}{\partial I_1},\\
1349: \Omega_2 & = & \frac{\partial H}{\partial I_2}.
1350: \end{eqnarray}
1351: In the generic case, these two frequencies are incommensurate, such that the
1352: full motion in the four-dimensional phase space 
1353: is quasi-periodic, and densely fills the
1354: so-called ``invariant torus" defined by the constant values $I_1$ and $I_2$.
1355: The semiclassical wave-function is constructed similarly to the WKB 
1356: wave-function.
1357: Turning points are now replaced by caustics \cite{ottb,einstein17,bornwolf} 
1358: of the classical
1359: motion (where the projection of the invariant torus on configuration
1360: space is singular), but the conclusions are essentially
1361: identical. The single-valued character of the wave-function requires the
1362: following quantization of the actions:
1363: \begin{eqnarray}
1364: I_1 & = & \left(n_1 + \frac{\mu_1}{4}\right) \hbar,\\
1365: I_2 & = & \left(n_2 + \frac{\mu_2}{4}\right) \hbar,
1366: \label{EBK}
1367: \end{eqnarray}
1368: where $n_1,n_2$ are two non-negative integers, and
1369: $\mu_1,\mu_2$ the Maslov indices (counting the number of caustics encountered
1370: on the torus)
1371: along the $\theta_1,\theta_2$ directions.
1372: Once again, 
1373: the semiclassical energy levels (which now depend on two quantum
1374: numbers) are obtained by substitution of these quantized values into the classical Hamilton
1375: function.
1376: 
1377: An alternative formulation of the EBK criterium 
1378: is possible using the original position-momentum 
1379: coordinates. Indeed, eq.~(\ref{EBK}) just expresses that, along any closed
1380: loop on the invariant torus, the phase accumulated 
1381: by the wave-function is an integer
1382: multiple of $2\pi$ (modulo the Maslov contribution). 
1383: Using the canonical invariance of the total action
1384: \cite{ozorio88}, 
1385: the EBK quantization conditions can be written as:
1386: \begin{equation}
1387: \frac{1}{2\pi} \oint_{{\mathrm closed\ path}\ \gamma_i}{\vec{p}. \d \vec{r}} = 
1388: \left(n_i + \frac{\mu_i}{4}\right) \hbar
1389: \label{ebkgen}
1390: \end{equation}
1391: where the integral has to be taken along two topologically 
1392: independent closed paths 
1393: $(\gamma_1,\gamma_2)$ on the
1394: invariant torus.   
1395: 
1396: Note that, as opposed to the WKB procedure in a 1D situation, 
1397: the EBK quantization uses the invariant tori of the classical
1398: dynamics, {\em not\/} the trajectories themselves.
1399: When there is a stable periodic orbit, it is 
1400: surrounded by invariant tori. 
1401: The smallest quantized torus around the stable orbit is associated with
1402: a quantum number equal to zero for the motion transverse to the orbit: 
1403: it defines 
1404: a narrow tube around the orbit, whose projection on configuration
1405: space will be 
1406: localized in the immediate vicinity
1407: of the orbit. Thus, the corresponding wave-function
1408: will also be localized close to this narrow tube, i.e.,
1409: along the stable periodic orbit in configuration space.
1410: Transversely to the orbit, the wave-function (or the Wigner function)
1411: will essentially look like the ground state of an harmonic oscillator,
1412: i.e. like a Gaussian. 
1413: 
1414: Finally, let us note that it is also possible to develop an analogous EBK
1415: scheme for periodically time dependent Hamiltonians \cite{breuer91}
1416: using the notion of an extended phase space \cite{lichtenberg83}.
1417: Such an approach will be extensively used in the next section, so it is 
1418: discussed in detail there.
1419: 
1420: 
1421: \subsection{Scars}
1422: 
1423: \label{scars}
1424: 
1425: When a periodic orbit is unstable, there is no torus closely surrounding it.
1426: However, it often happens that quantum eigenstates exhibit an increased
1427: probability density in the vicinity of unstable periodic orbits. This scarring
1428: phenomenon  is nowadays relatively well understood, and the interested
1429: reader may 
1430: consult references \cite{heller89,stoeckmann99}.
1431: 
1432: Similarly, some quantum states have
1433: an increased probability density 
1434: in the vicinity of an unstable equilibrium point
1435: \cite{abu95d,jensen89b,leopold94}.
1436: This localization is only partial. Indeed, since a quantum eigenstate
1437: is a stationary structure, some probability density
1438: {\em must} localize along the unstable directions of the classical Hamiltonian
1439: flow \cite{ozorio88}, and the localization cannot be perfect.
1440: This is in sharp contrast with stable equilibrium points and stable
1441: periodic orbits which -- see above -- optimally support
1442: localized eigenstates.
1443: 
1444: Note that there is, however, a big difference between scarring and localization
1445: in the vicinity of an unstable fixed point. The latter phenomenon
1446: is of purely classical origin. Indeed, close to an 
1447: equilibrium point, the velocity goes to zero and the particle consequently 
1448: spends
1449: more time close to the equilibrium point than further away from it. The quantum
1450: eigenfunctions have the same property: the probability density is large
1451: near the equilibrium point. This trivial enhancement of the
1452: probability density is already well known 
1453: for a one-dimensional
1454: system where the WKB wave-function, eq.~(\ref{psi_wkb}),
1455: diverges when the momentum
1456: tends to zero. 
1457: The localization effect
1458: near an unstable point is just the quantum manifestation of the
1459: critical slowing down of the classical particle~\cite{delande97}.
1460: 
1461: \section{Non-dispersive wave-packets and their realization in various 
1462: atomic systems}
1463: 
1464: \subsection{General model -- nonlinear resonances}
1465: \label{GM}
1466: 
1467: In this section, we present the general theory of non-dispersive wave-packets.
1468: As explained in section \ref{INTRO3}, the basic ingredients
1469: for building a non-dispersive wave-packet are a non-linear
1470: dynamical system and an external periodic driving which is
1471: resonant with an internal frequency of the dynamical system.
1472: We present here a very general theory starting out from classical
1473: mechanics which provides us with the most suggestive 
1474: approach to non-linear resonances. In a second step, we choose a pure
1475: quantum approach 
1476: giving essentially the same physics. 
1477: 
1478: We use a one-dimensional model, which displays all the interesting
1479: features of non-linear resonances. 
1480: While the direct link between classical nonlinear resonances, the corresponding
1481: Floquet states, and non-dispersive wave-packets has been identified only 
1482: recently \cite{abuth,delande94,abu95d,delande95,abu96} some 
1483: aspects of the developments presented below
1484: may be found in earlier studies  
1485: \cite{berman77,zaslavsky81,henkel92,holthaus95}.
1486: 
1487: Several complications not included in the simple one-dimensional
1488: model are important features of ``real systems". 
1489: They are discussed at a later stage in this paper:
1490: \begin{itemize}
1491: \item the effect of additional degrees
1492: of freedom, in sections \ref{LIN3D}-\ref{EP};
1493: \item higher 
1494: nonlinear resonances (where the driving frequency is a multiple
1495: of the internal frequency), in section \ref{HOR};
1496: \item an unbounded phase space, leading to the decay of 
1497: non-dispersive wave-packets (as ``open quantum systems''), 
1498: in section \ref{ION};
1499: \item sources of  ``decoherence'', such as
1500: spontaneous emission of atomic wave-packets, in section \ref{SPO};
1501: \item deviations from 
1502: temporal periodicity, in section \ref{PTP}.
1503: \end{itemize}
1504: 
1505: 
1506: In 
1507: particular cases, an apparently simpler approach is also possible
1508: (such as the use of the rotating frame for a Rydberg atom exposed to a
1509: circularly polarized electromagnetic field, see section \ref{CP}).
1510: Despite all its advantages, it may be quite specific and 
1511: too restricted to reveal nonlinear resonances 
1512: as the actual cause of the
1513: phenomenon. Here, we seek the most general description.
1514: 
1515: \subsubsection{Classical dynamics}
1516: \label{CD}
1517: 
1518: Let us start from a time-independent, bounded, one-dimensional system
1519: described by the Hamilton function
1520: $H_0(p,z)$.
1521: Since energy is conserved, the motion is confined to a
1522: one-dimensional manifold in two-dimensional phase space.
1523: Except for energies which define a fixed point of the
1524: Hamiltonian dynamics (such that $\partial H_0/\partial z = 0$
1525: and $\partial H_0/\partial p = 0;$ these fixed points generically
1526: only exist at some 
1527: isolated values of energy, for example at $E=0$
1528: for the harmonic oscillator), the motion is periodic
1529: in time, and the phase space trajectory is a simple closed loop.
1530: 
1531: It is always possible to find a set of canonically
1532: conjugate phase space coordinates adapted to the dynamics of the system.
1533: These are the action-angle coordinates $(I,\theta)$,
1534: whose existence is guaranteed
1535: by the Liouville-Arnold theorem \cite{lichtenberg83}, with:
1536: \begin{eqnarray}
1537: &&0 \leq I, \\
1538: &&0 \leq \theta \leq 2\pi, \label{thetaint} \\
1539: &&\left\{ \theta , I \right\} = 1,
1540: \end{eqnarray}
1541: and $\{ .,.\}$ the usual Poisson brackets, eq.~(\ref{poisson_brackets}).
1542: 
1543: A fundamental property is that the Hamilton function in these coordinates
1544: depends on $I$ alone, not on $\theta:$
1545: \begin{equation}
1546: H_0 = H_0(I).
1547: \end{equation}
1548: As a consequence of 
1549: Hamilton's
1550: equations of motion, $I$ is a
1551: constant of motion, and 
1552: \begin{equation}
1553: \theta =\Omega t +\theta_0
1554: \label{thetatime}
1555: \end{equation}
1556: evolves linearly in time,
1557: with the angular velocity
1558: \begin{equation}
1559: \Omega(I) = \frac{\partial H_0}{\partial I}(I),
1560: \label{internal_omega}
1561: \end{equation}
1562: which depends on the 
1563: action $I$.
1564: The period of the motion at a given value of $I$ reads
1565: \begin{equation}
1566: T = \frac{2\pi}{\Omega(I)}.
1567: \end{equation}
1568: 
1569: In simple words, the action $I$ is nothing but the properly
1570: ``rescaled" total energy, and the angle $\theta$ just measures how time
1571: evolves
1572: along the (periodic) orbits. 
1573: In a one-dimensional system, the action variable
1574: can be expressed as an
1575: integral along the orbit, see eq.~(\ref{action_def}).
1576: 
1577: 
1578: Suppose now that the system is exposed to a periodic driving force, such
1579: that the Hamilton function, in the original coordinates, writes
1580: \begin{equation}
1581: \label{ham}
1582: H = H_0(p,z) + \lambda V(p,z) \cos \omega t,
1583: \label{h_gen}
1584: \end{equation}
1585: with $\omega$ the frequency of the periodic drive and
1586: $\lambda$ some small parameter which determines
1587: the strength of the perturbation.
1588: For simplicity, we choose a single cosine function to define the
1589: periodic driving.  For a more
1590: complicated dependence on time \cite{ringot00}, 
1591: it is enough to expand it in a Fourier
1592: series, see section \ref{EP}.
1593: The equations become slightly more complicated, but the physics
1594: is essentially identical.
1595: 
1596: We now express the perturbation $V(p,z)$ in action-angle coordinates.
1597: Since $\theta$ is $2\pi-$periodic, eq.~(\ref{thetaint}), we obtain a Fourier series:
1598: \begin{equation}
1599: V(I,\theta) = \sum_{m=-\infty}^{+\infty}
1600: {V_m(I)\ \exp (im\theta)}.
1601: \label{fourier_components}
1602: \end{equation}
1603: Note that, as $\theta$ evolves linearly with time $t$ for the unperturbed
1604: motion (and therefore parametrizes an unperturbed periodic orbit),
1605: the $V_m$ can also be seen as the Fourier
1606: components of $V(t)$ evaluated along the classical, unperturbed trajectory.
1607: Furthermore, since the Hamilton function is real, $V_{-m}=V_m^*.$
1608: Again for the sake of simplicity, we will assume that both are
1609: real and thus equal.
1610: The general case can be studied as well, at the price of slightly
1611: more complicated formulas.
1612: 
1613: Plugging eq.~(\ref{fourier_components}) in eq.~(\ref{h_gen}) results in the
1614: following Hamilton function in action-angle coordinates,
1615: \begin{equation}
1616: H = H_0(I) + \lambda  \sum_{m=-\infty}^{+\infty}
1617: {V_m(I)\ \exp (im\theta)} \cos \omega t,
1618: \end{equation}
1619: which (assuming $V_{-m}=V_m$ -- see above) can be rewritten as
1620: \begin{equation}
1621: H = H_0(I) + \lambda \sum_{m=-\infty}^{+\infty}
1622: {V_m(I)\ \cos (m\theta-\omega t)}.
1623: \label{hamfou}
1624: \end{equation}
1625: For $\lambda$ sufficiently small, the phase space trajectories of the perturbed dynamics 
1626: will remain close to the unperturbed ones (for short times).
1627: This means
1628: that $m\theta -\omega t$ evolves approximately linearly in time
1629: as 
1630: $(m\Omega - \omega)t$ (see eq.~(\ref{thetatime})),
1631: while $I$ is slowly varying.
1632: It is therefore reasonable to expect that all the terms
1633: $ V_m(I)\ \cos (m\theta-\omega t)$ will oscillate rapidly and average out
1634: to zero,
1635: leading to an effective approximate Hamiltonian identical to the unperturbed
1636: one.
1637: Of course, this approach is too simple. Indeed, close to a  
1638: ``resonance'', 
1639: where $(s\Omega - \omega)$ is small, the various terms  
1640: $V_m\ \cos (m\theta-\omega t)$ oscillate, except for the $m=s$ term which
1641: may evolve very {\em slowly} and affect the dynamics considerably.
1642: For simplicity, we restrict the present analysis to the principal
1643: resonance such that $\Omega \simeq \omega$.  The extension
1644: to higher  resonances (with $s\Omega\simeq\omega$) is discussed
1645: in section \ref{HOR}.
1646: 
1647: 
1648: Our preceding remark is the basis of the ``secular approximation''
1649: \cite{lichtenberg83,cct92}.
1650:  The
1651: guiding idea
1652: is to perform a canonical change of coordinates involving
1653: the slowly varying variable $\theta -\omega t.$ Because of the explicit
1654: time dependence, this requires first 
1655: the passage to an extended phase space,
1656: which comprises time as an additional coordinate. The Hamilton function
1657: in extended phase space is defined by
1658: \begin{equation}
1659: {\cal H} = P_t + H,
1660: \label{hamext}
1661: \end{equation}
1662: with $P_t$ the momentum canonically conjugate to the new coordinate
1663:  - time $t$. The physical time $t$ is now parametrized by
1664: some
1665: fictitious time, say $\xi$. However,
1666: \begin{equation}
1667: \frac{\partial{\cal H}}{\partial P_t}=\frac{dt}{d\xi}=1,
1668: \end{equation}
1669:  i.e., $t$ and
1670: $\xi$ are essentially identical.
1671: ${\cal H}$, being independent of $\xi$,
1672:  is conserved as  $\xi$ evolves.
1673: The requested transformation to slowly varying variables $\theta -\omega t$ 
1674: reads:\footnote{This
1675: {\em canonical}
1676: change of coordinates is often refered to as
1677: ``passing to the rotating frame". It should however be emphasized that
1678: this
1679: suggests the correct picture {\em only}
1680: in phase space
1681: spanned by the action-angle coordinates $(I,\theta)$. In the original
1682: coordinates $(p,z)$, the transformation is usually very complicated, and
1683: only rarely
1684: a standard rotation in configuration space (see also section \ref{CP}).}
1685: \begin{eqnarray}
1686: &&\hat{\theta} = \theta - \omega t, \label{rotframe_a} \\
1687: &&\hat{I} = I,\label{rotframe_b} \\
1688: &&\hat{P}_t = P_t + \omega I,
1689: \label{rotframe_c}
1690: \end{eqnarray}
1691:  which transforms ${\cal H}$ into
1692: \begin{equation}
1693:  \hat{\cal H} = \hat{P}_t + H_0(\hat{I}) -\omega \hat{I}+ \lambda
1694: \sum_{m=-\infty}^{+\infty}
1695: {V_m(\hat{I})\ \cos (m\hat{\theta}+(m-1)\omega t)}.
1696: \label{floq_sc_gen}
1697: \end{equation}
1698: This Hamilton function does not
1699: involve any approximation yet. Only in the next step
1700: we average $\hat{\cal H}$ over the fast variable $t,$ i.e., over one period
1701: of the external driving. This has the effect of canceling all oscillating
1702: terms in the sum, except
1703: the resonant one, defined by $m=1$.
1704: Consequently, we are left with the approximate, 
1705: ``secular'' Hamilton
1706: function:
1707: \begin{equation}
1708: {\cal H}_{\mathrm sec} = \hat{P}_t + H_0(\hat{I}) -\omega \hat{I}+
1709: \lambda
1710: V_1(\hat{I}) \cos \hat{\theta}.
1711: \label{hsec_ap}
1712: \end{equation}
1713: 
1714: The secular Hamilton function no longer depends on time. Hence,
1715: $\hat{P}_t$ is a constant
1716: of motion and we are left with an integrable Hamiltonian system living in a
1717: two-dimensional phase space, spanned by $(\hat{I},\hat{\theta})$.
1718: The above averaging procedure is valid at first order in $\lambda$.
1719: Higher order expansions, using, e.g., the Lie algebraic
1720: transformation method~\cite{lichtenberg83}, 
1721: are possible.\footnote{An example is given
1722: in \cite{abu97}, in a slightly different situation,
1723: where the perturbation is not resonant with the
1724: internal frequency.} Basically, the interesting physical phenomena
1725: are already present at lowest non-vanishing order, to which we will
1726: restrain in the following.
1727: 
1728: The dynamics
1729: generated by the secular Hamilton function
1730: is rather simple. At order zero in $\lambda,$
1731: $\hat{I}$ is constant and $\hat{\theta}$ evolves linearly with time.
1732: As we can read from eq.~(\ref{hsec_ap}),  a
1733: continuous family (parametrized by the value of
1734: $0\leq \hat{\theta}<2\pi$)
1735: of fixed points exists
1736: if $d\hat{\theta}/dt=\partial{\cal H}_{\mathrm
1737: sec}/\partial\hat{I}$ vanishes,
1738: i.e., at actions $\hat{I}_1$ such that
1739: \begin{equation}
1740: \Omega(\hat{I}_1) = \frac{\partial H_0}{\partial I}(\hat{I}_1) = \omega.
1741: \label{resonance_condition}
1742: \end{equation}
1743: Thus, unperturbed trajectories that are {\em resonant} with the
1744: external drive are fixed points
1745: of the unperturbed secular
1746: dynamics. This is precisely why slowly
1747: varying variables
1748: are introduced.
1749: 
1750: Typically, eq.~(\ref{resonance_condition}) has only isolated solutions --
1751: we will
1752: assume that in the following. Such is the case when
1753: $\partial^2 H_0/\partial I^2$ does not vanish -- excluding
1754: the
1755: pathological situation of the harmonic oscillator, where all trajectories
1756: are simultaneously resonant.
1757: Hence,
1758: if $\partial^2 H_0/\partial I^2$ is positive, the line 
1759: ($\hat{I}=\hat{I}_1,0\leq \hat{\theta}<2\pi$, parametrized by $\hat\theta$)
1760: is a minimum of the unperturbed secular Hamilton function ${\cal H}_{\mathrm
1761: sec}$;
1762: if $\partial^2 H_0/\partial I^2$
1763: is negative, it is a maximum.
1764: 
1765: At first order in $\lambda$, the fixed points of the secular
1766: Hamiltonian should have an action close to $\hat{I_1}.$ Hence, it is
1767: reasonable to perform
1768: a power expansion of the unperturbed Hamiltonian in the vicinity of
1769: $\hat{I}=\hat{I_1}.$
1770: We obtain the following approximate  Hamiltonian:
1771: \begin{equation}
1772: {\cal H}_{\mathrm pend} = \hat{P}_t + H_0(\hat{I_1}) -\omega\hat{I}_1 + \frac{1}{2}
1773: H^{''}_0(\hat{I_1})\ (\hat{I}-\hat{I_1})^2 + \lambda
1774: V_1(\hat{I_1}) \cos \hat{\theta},
1775: \label{eqpend}
1776: \end{equation}
1777: with:
1778: \begin{equation}
1779: H^{''}_0 = \frac{\partial^2 H_0}{\partial I^2}.
1780: \label{second_derivative}
1781: \end{equation}
1782: Consistently at lowest order in $\lambda,$ it is not necessary 
1783: to take into
1784: account the dependence of $V_1$ on $\hat{I}.$
1785: 
1786: As already anticipated by the label, ${\cal H}_{\mathrm pend}$ defined
1787: in eq.~(\ref{eqpend}) describes a usual, one-dimensional
1788: pendulum: $\hat\theta$ represents the angle of the pendulum 
1789: with the vertical axis,
1790: $\hat{I}-\hat{I_1}$ its angular velocity, $1/H^{''}_0(\hat{I_1})$
1791: its momentum of inertia
1792: and $\lambda V_1(\hat{I_1})$ the gravitational field.
1793: This equivalence of the secular Hamilton function with that of a pendulum,
1794:  in the
1795: vicinity of the resonant action $\hat{I}_1$, is extremely useful to gain
1796: some physical insight
1797: in the dynamics of any Hamiltonian system close to a resonance. In particular,
1798: it will render our
1799: analysis of non-dispersive wave-packets rather simple.
1800: 
1801: Figure~\ref{pendulum} shows the isovalue lines of
1802: ${\cal H}_{\mathrm pend}$ in the $(\hat{I},\hat{\theta})$ plane, i.e. the
1803: classical phase space trajectories in the presence of the
1804: resonant perturbation. In
1805: the absence of the resonant perturbation, these should be horizontal 
1806: straight lines at
1807: constant  $\hat{I}.$
1808: We observe that the effect of the resonant perturbation is mainly
1809: to create a new structure, called the ``resonance island", located around
1810: the resonant action $\hat{I}_1.$ 
1811: 
1812: \psfull
1813: \begin{figure}
1814: \centerline{\psfig{figure=bdzf08.eps,width=10cm,angle=-90}}
1815: \caption{Isovalues of the Hamiltonian ${\cal H}_{\mathrm pend}$ of a
1816: pendulum in a gravitational field. This Hamiltonian is a good
1817: approximation for the motion of a periodically driven system when
1818: the driving frequency is resonant with the internal frequency
1819: of the system. The stable equilibrium point of the pendulum is surrounded
1820: by an island of librational motion (shaded region). 
1821: This defines the resonance island of the periodically driven system, where
1822: the internal motion is {\em locked} on the external driving. This 
1823: nonlinear phase-locking
1824: phenomenon is essential for the existence of non-dispersive wave-packets.}
1825: \label{pendulum}
1826: \end{figure}
1827: 
1828: To characterize this structure, let us examine the
1829: fixed points of the Hamiltonian~(\ref{eqpend}). They are easily calculated
1830: (imposing $\partial{\cal H}_{\mathrm pend}/\partial\hat{I}=\partial{\cal
1831: H}_{\mathrm pend}/\partial\hat\theta=0$), and
1832: located at
1833: \begin{equation}
1834: \hat{I}=\hat{I_1},\ \ \ \hat{\theta}=0,\ \ {\mathrm with\ energy}\
1835: H_0(\hat{I_1}) -\omega\hat{I}_1 + \lambda V_1(\hat{I_1}),
1836: \label{stablep}
1837: \end{equation}
1838: and
1839: \begin{equation}
1840: \hat{I}=\hat{I_1},\ \ \ \hat{\theta}=\pi,\ \ {\mathrm with\ energy}\
1841: H_0(\hat{I_1}) -\omega\hat{I}_1- \lambda V_1(\hat{I_1}),
1842: \label{resoact}
1843: \end{equation}
1844: respectively.
1845: 
1846: If $H^{''}_0(\hat{I_1})$ (and thus the ``kinetic energy'' part
1847: in ${\cal H}_{\mathrm pend}$)
1848: is positive, the minimum of the potential
1849:  $\lambda V_1(\hat{I_1})\cos\hat{\theta}$ corresponds to a global
1850:  minimum of ${\cal H}_{\mathrm pend}$, and thus to a stable
1851: equilibrium point. The maximum of $\lambda V_1(\hat{I_1})\cos\hat{\theta}$
1852: is a saddle point of ${\cal H}_{\mathrm pend}$, 
1853: and thus 
1854: represents an unstable equilibrium point,
1855: as the standard intuition suggests. For $H^{''}_0(\hat{I}_1)<0$
1856: the situation is reversed -- and less intuitive 
1857: since the
1858: ``kinetic energy'' is negative -- and the {\em maximum}
1859: of ${\cal H}_{\mathrm pend}$ is now a stable equilibrium point,
1860: as the reader may easily check by
1861: standard linear stability analysis in the vicinity of the
1862: fixed point. 
1863: Thus, in compact form, if
1864: $\lambda V_1(\hat{I_1}) H^{''}_0(\hat{I_1})$ is positive,
1865: $\hat{\theta}=\pi$ is
1866: a stable equilibrium point, while $\hat{\theta}=0$ is
1867: unstable. 
1868: If $\lambda V_1(\hat{I_1}) H^{''}_0(\hat{I_1})$ is negative, the stable and unstable
1869: points are interchanged.
1870: 
1871: There are two qualitatively different types of motion:
1872: \begin{itemize}
1873: \item Close to the stable equilibrium point of the pendulum, $\hat{\theta}$
1874: oscillates
1875: periodically, with an amplitude smaller than $\pi$.
1876: This is the ``librational motion'' of the pendulum inside the
1877: resonance
1878: island. Any trajectory started within 
1879: this region of phase space (the shaded area in
1880: fig.~\ref{pendulum})
1881: exhibits 
1882: librational motion. It should be realized
1883: that the resonance island confines the motion to finite intervals in $\hat{I}$
1884: and $\hat{\theta}$, and thereby strongly affects all trajectories with action
1885: close to the resonant action. According to eqs.~(\ref{stablep},\ref{resoact}),
1886: the resonance island is associated with the energy range
1887: $[H_0(\hat{I_1})-\omega\hat{I}_1- |\lambda V_1(\hat{I_1})|,
1888: H_0(\hat{I_1})-\omega\hat{I}_1+ |\lambda
1889: V_1(\hat{I_1})|].$
1890: \item For any initial energy outside that energy range
1891:  the pendulum has sufficient
1892: kinetic energy to rotate. This is the ``rotational
1893: motion'' of the pendulum outside the resonance island, where $\hat{\theta}$ is
1894: an unbounded and
1895: monotonous function of time. Far from the center of the island, the motion
1896: occurs at almost constant unperturbed action $\hat{I}$, with an almost
1897: constant
1898: angular velocity in $\hat{\theta},$ 
1899: tending to the unperturbed
1900: motion. This illustrates that the effect of the perturbation is 
1901: important
1902: for initial conditions close to the resonance island, but negligible for 
1903: non-resonant trajectories.
1904: \end{itemize}
1905: 
1906: The size of the resonance island can be simply estimated from
1907: eq.~(\ref{eqpend}) and
1908: fig.~\ref{pendulum}. The extension in $\hat{\theta}$ is $2\pi$, its width in
1909: $\hat{I}$ (which depends on $\hat{\theta})$ is:
1910: \begin{equation}
1911: \Delta \hat{I} = 4 \sqrt{\left|
1912: \frac{\lambda V_1(\hat{I_1})}{ H^{''}_0(\hat{I_1})}\right|},
1913: \label{width-island}
1914: \end{equation}
1915: and the total area \cite{lichtenberg83}:
1916: \begin{equation}
1917: A(\lambda) = 16 \sqrt{\left|
1918: \frac{\lambda V_1(\hat{I_1})}{ H^{''}_0(\hat{I_1})}\right|}.
1919: \label{area}
1920: \end{equation}
1921: 
1922: The dependence of $A(\lambda)$ on $\sqrt{|\lambda|}$ implies that even
1923: a small perturbation may
1924: induce
1925: significant changes in the phase space structure, provided the pertubation
1926: is resonant.
1927: 
1928: The above picture is valid in the rotating frame defined by
1929: eqs.~(\ref{rotframe_a}-\ref{rotframe_c}). If we go back to the original
1930: action-angle coordinates $(I,\theta),$ the stable (resp. unstable) fixed
1931: point of the secular Hamiltonian is mapped on a
1932: stable (resp. unstable) periodic orbit
1933: whose period is {\em exactly} equal to the period of the driving perturbation,
1934: as a consequence of eq.~(\ref{resonance_condition}).
1935: Any trajectory started in the vicinity of the stable periodic orbit will
1936: correspond to an initial point close to the fixed
1937: point in the rotating frame,
1938: and thus will remain trapped within the resonance island. In the original
1939: coordinate frame, it will appear
1940: as a trajectory evolving close to the stable periodic orbit forever.
1941: In particular, 
1942: the difference in $\theta$ between the stable periodic orbit and any orbit 
1943: trapped in the resonance island remains bounded within $(-\pi,+\pi)$, for arbitrarily 
1944: long times. This means that the phase of any trapped trajectory cannot drift with 
1945: respect to the phase of the periodic orbit.
1946:  As the latter evolves at the driving frequency, we
1947: reach the conclusion that the phase of any trajectory started in the
1948: resonance island
1949: will be {\em locked} on the phase of the driving field. This is the very origin
1950: of the phase locking phenomenon discussed in section~\ref{INTRO3}
1951: above. A crucial point herein
1952: is that
1953: the resonance island
1954: covers a significant part of phase space with finite volume:
1955: it is the
1956: whole structure, not few trajectories, which is phase locked. This is why, 
1957: further down, we will be able to build quantum wave-packets on this structure, 
1958: which will be phase locked to the classical orbit 
1959: and will not spread. The classical version of a non-spreading
1960: wave-packet
1961: thus consists of a family of trajectories, trapped within the resonance island, such
1962: that
1963: this family is invariant under the evolution generated by the pendulum
1964: Hamiltonian. The simplest
1965: possibility
1966: is to sample all trajectories within the ``energy" range
1967: $[H_0(\hat{I_1}) -\omega \hat{I}_1 - |\lambda V_1(\hat{I_1})|,
1968: H_0(\hat{I_1})-\omega \hat{I}_1+ |\lambda
1969: V_1(\hat{I_1})|]$ of the pendulum.
1970: In real space, this will appear as a localized probability density
1971: following the classical stable periodic orbit, reproducing 
1972: its shape exactly after each period of the drive.
1973: 
1974: So far, to derive the characteristics
1975: of the resonance island, we have consistently used
1976: first order
1977: perturbation theory, which is valid for small $\lambda.$ At higher values of
1978: $\lambda,$ higher order terms come into play and modify the shape and the
1979: precise location
1980: of the resonance island. However,
1981: it is crucial to note that the island itself considered as a structure is
1982: robust,
1983: and will survive up to rather high values of $\lambda$ (as a consequence of
1984: the KAM theorem \cite{lichtenberg83}).
1985: Since the size of the resonance island grows with
1986: $|\lambda|$, 
1987: eq.~(\ref{area}), the island may occupy
1988:  a significant
1989: area in phase space and eventually interact with
1990: islands associated
1991: with other resonances, for sufficiently large $|\lambda|$. The mechanism of this ``resonance
1992: overlap'' is rather well understood \cite{chirikov59,chirikov79}: in general,
1993: the motion close to the separatrix (where the period of the classical motion
1994: of the pendulum tends to infinity) is 
1995: most sensitive to higher order 
1996: corrections.
1997: The general scenario
1998: is thus the 
1999: nonintegrable perturbation of the separatrix and the 
2000: emergence of a
2001: ``stochastic" layer of
2002: chaotic motion in phase space, as 
2003: $|\lambda|$ is increased. At
2004: still larger
2005: values of $|\lambda|$, 
2006: chaos may invade large parts of phase space, 
2007: and the resonance island may shrink
2008: and finally
2009: disappear.
2010: While considering  
2011: realistic examples later-on, we shall 
2012: enter the
2013: non-perturbative regime. Let us, however, consider first the quantum
2014: perturbative picture.
2015: 
2016: \subsubsection{Quantum dynamics}
2017: \label{QD}
2018: 
2019: As shown in the previous section, the dynamics of
2020: a one-dimensional system exposed to a weak, resonant, periodic driving
2021: is essentially regular and analogous to the one of a pendulum,
2022: eq.~(\ref{second_derivative})
2023: (in the rotating frame, eqs.~(\ref{rotframe_a}-\ref{rotframe_c})).
2024: In the present section, we will
2025: show that the same
2026: physical picture can be employed in quantum mechanics, to construct
2027: non-dispersive wave-packets. They will follow the stable
2028: classical trajectory locked on the external drive,
2029: and 
2030: exactly reproduce their initial shape 
2031: after 
2032: each period.
2033: 
2034: Our starting point is the time-dependent Schr\"odinger equation
2035: associated with the Hamiltonian (\ref{ham}):\footnote{For simplicity,
2036: we use the same notation for
2037: classical and quantum quantities, the distinction between them
2038:  will
2039: become clear from the context.}
2040: \begin{equation}
2041: i\hbar \frac{d|\psi(t)\rangle}{dt} = (H_0 + \lambda V  \cos \omega t)
2042: |\psi(t)\rangle.
2043: \label{se}
2044: \end{equation}
2045: Since the Hamiltonian (\ref{ham}) is periodic in time, the Floquet
2046: theorem\footnote{The Floquet theorem \protect\cite{floquet1883} in the time
2047: domain is strictly equivalent to the Bloch theorem for potentials
2048: periodic in space \protect\cite{mermin76}.}
2049: guarantees that the general solution of eq.~(\ref{se}) is given by
2050: a linear combination
2051: of elementary, time-periodic states -- the so-called ``Floquet eigenstates'' of
2052: the system -- multiplied by oscillatory functions:
2053: \begin{equation}
2054: |\psi(t)\rangle = \sum_j c_j \exp\left( - i \frac{{\cal E}_jt}
2055: {\hbar}\right) |{\cal E}_j(t)\rangle,
2056: \label{solse}
2057: \end{equation}
2058: with
2059: \begin{equation}
2060: |{\cal E}_j(t+T)\rangle =|{\cal E}_j(t)\rangle .
2061: \label{flstate}
2062: \end{equation}
2063: The ${\cal E}_j$ are the 
2064: ``quasi-energies" of the system.
2065: Floquet states  and quasi-energies are eigenstates
2066: and eigenenergies
2067: of the Floquet Hamiltonian
2068: \begin{equation}
2069:  {\cal H} = H_0 + \lambda V  \cos \omega t - i \hbar \frac{\partial}{\partial
2070:  t}.
2071: \label{calhq}
2072: \end{equation}
2073: Note that, because
2074: of the time-periodicity with period $T=2\pi/\omega$, the quasi-energies are
2075: defined modulo $\hbar \omega$ \cite{shirley65}.
2076: 
2077: The Floquet Hamiltonian (\ref{calhq}) is nothing but the quantum analog of
2078: the classical Hamiltonian (\ref{hamext})
2079: in 
2080: extended phase space.
2081: Indeed, $- i \hbar \partial /\partial t$ is the quantum version of the
2082: canonical momentum $P_t$ conjugate to time $t.$
2083: In 
2084: strict analogy with the classical discussion of the previous section,
2085: it is the Floquet Hamiltonian in 
2086: extended phase space which will be the
2087: central object of our discussion. It contains all the
2088: relevant information on the system, encoded in its eigenstates.
2089: 
2090: In a quantum optics or atomic physics context -- 
2091: with the external perturbation 
2092: given by quantized modes of the electromagnetic field -- the
2093: concept of ``dressed atom" is widely used \cite{cct92}. There, 
2094: a given field mode 
2095: and the atom are treated on an equal footing, as a composite quantum system, 
2096: leading to a time-independent Hamiltonian (energy is conserved for the entire system
2097: comprising atom {\em and} field).
2098: This picture
2099: is 
2100: indeed very close to the Floquet picture. If the field mode is in a 
2101: coherent state \cite{cct92} with a large average number of
2102: photons, 
2103: the electromagnetic field can be treated (semi)classically -- i.e., replaced 
2104: by a $\cos$ time dependence and a fixed amplitude $F$ -- and the
2105: energy spectrum of the dressed atom exactly coincides with the
2106: spectrum of the Floquet Hamiltonian \cite{shirley65}.
2107: 
2108: By its mere definition, eq.~(\ref{flstate}),
2109: each Floquet eigenstate is
2110: associated with a strictly time-periodic probability density in configuration
2111: space.
2112: Due to this periodicity with the period of the driving field,
2113: the probability density of a Floquet eigenstate in general changes its shape
2114: as time evolves, but recovers its initial shape after 
2115: each period.
2116: Hence, the Floquet picture provides clearly the simplest approach to
2117: non-dispersive wave-packets. Given the ability to build a Floquet
2118: state which is well localized at a given phase of the driving field, it will
2119: {\em automatically} represent a non-dispersive wave-packet.
2120: In our opinion, this is a much simpler approach than the attempt to
2121: build an {\it a priori} localized wave-packet and try to minimize
2122: its spreading during the subsequent evolution
2123: \cite{farrelly95a,ibb95,ibb94,kalinski95a}.
2124: 
2125: Note that also
2126: the reverse property holds true. Any state with $T$-periodic probability
2127: density (and, in particular, any localized wave-packet propagating along a
2128: $T$-periodic classical orbit)
2129: has to be a single Floquet eigenstate:
2130: Such a state can be expanded into the Floquet eigenbasis, and during one
2131: period, the various components of the expansion accumulate phase factors
2132: $\exp(-i{\mathcal E}_iT/\hbar).$ Hence, 
2133: the only solution which allows for a $T$-periodic density is
2134: a one-component expansion, i.e., a single Floquet eigenstate
2135: \footnote{One
2136: might argue that Floquet states differing in energy by an
2137: integer multiple of $\hbar\omega$ could be used. However, 
2138: the Floquet spectrum is $\hbar\omega$-periodic by construction
2139: \cite{shirley65}, and two
2140: such states
2141: represent the same physical state.}.
2142: 
2143: To summarize, the construction of non-dispersive
2144: wave-packets in a time-periodic system is equivalent to finding
2145: localized Floquet eigenstates. The existence of such states is far
2146: from obvious, as the Floquet spectrum is usually very complex, composed of
2147: quasi-bound states, resonances and continua. This is why
2148: a semiclassical analysis can be very helpful in finding these objects.
2149: 
2150: \subsubsection{Semiclassical approximation}
2151: \label{sapp}
2152: 
2153: 
2154: Dealing with highly excited states, a semiclassical
2155: approximation can be used to determine quasi-energies and
2156: Floquet eigenstates~\cite{breuer91}. If the driving perturbation is sufficiently weak,
2157: we have shown in section~\ref{CD} that the classical dynamics close to a
2158: nonlinear resonance is
2159: essentially regular and accurately described by the pendulum
2160: Hamiltonian~(\ref{eqpend}). It describes a system with two
2161: degrees of freedom (along $t$ and $\hat\theta$, with their
2162: conjugate momenta $\hat{P}_t$ and $\hat{I}$, respectively)
2163: which is essentially regular. For semiclassical quantization, 
2164: we may then
2165: use the standard EBK rules, eq.~(\ref{EBK}), introduced in section~\ref{SQ}.
2166: 
2167: The momentum $\hat{P}_t$ is a constant of the motion,
2168: and the isovalue curves of $\hat{P}_t,\hat{I},\hat{\theta}$
2169: lying on the
2170: invariant tori
2171: can be used for the EBK quantization scheme. 
2172: Along such a curve, $t$
2173: evolves from $0$ to $2\pi/\omega$, with $\hat{\theta}=\theta-\omega t$
2174: kept constant. Thus, $\theta$ itself is changed by $2\pi$, what
2175: implies that  the Maslov index $\mu$ of the unperturbed $(I,\theta)$ motion
2176: has to be included,
2177: leading to the following quantization condition for $\hat{P}_t$,
2178: \begin{equation}
2179: \frac{1}{2\pi}\int_0^{T}\hat{P}_tdt=\frac{\hat{P}_tT}{2\pi}=
2180: \left(k+\frac{\mu}{4}\right)\hbar ,
2181: \label{ebkpt}
2182: \end{equation}
2183: with integer $k$.
2184: Since $T=2\pi/\omega$ is just the period of the resonant driving,
2185: we get the quantized values of $\hat{P}_t:$ 
2186: \begin{equation}
2187: \hat{P}_t = \left(k+\frac{\mu}{4}\right)\ \hbar\omega
2188: \label{ebkpt2}
2189: \end{equation}
2190: which are equally spaced by
2191: $\hbar \omega.$ Thus, we recover semiclassically 
2192: the $\omega$-periodicity of the Floquet spectrum.
2193: 
2194: For the motion in the $(\hat{I},\hat{\theta})$ plane, we can use
2195: the isocontour lines
2196: of the pendulum Hamiltonian $H_{\mathrm pend}$, eq.~(\ref{eqpend}), 
2197: as closed paths,
2198: keeping $\hat{P}_t$ and $t$ constant.
2199: Depending on the nature of the pendulum motion (librational or
2200: rotational), the topology of the closed paths is different,
2201: leading to distinct expressions:
2202: \begin{itemize}
2203: \item For trapped librational motion, inside the resonance island,
2204: the path is isomorphic to a circle in the $(\hat{I},\hat{\theta})$ plane,
2205: with a Maslov index equal to two. The quantization condition
2206: is:
2207: \begin{equation}
2208: \frac{1}{2\pi}\oint \hat{I}d\hat{\theta}=\left(N+\frac{1}{2}\right)\hbar,\
2209: ({\rm librational\ motion})
2210: \label{ebkires}
2211: \end{equation}
2212: with $N$ a non-negative integer.
2213: Of special interest is the ``fundamental" state, $N=0$, which exhibits
2214: maximum localization within the resonance island and
2215: is therefore
2216: expected to represent the optimal non-dispersive wave-packet.
2217: \item For unbounded rotational motion, outside the
2218: resonance island, the path includes a $2\pi$ phase change for $\theta$
2219: and acquires the Maslov index of the unperturbed motion:
2220: \begin{equation}
2221: \frac{1}{2\pi}\int_{0}^{2\pi} \hat{I}d\hat{\theta}=\left(
2222: N+\frac{\mu}{4}\right)\hbar,\ ({\rm rotational \ motion}).
2223: \label{ebkinres}
2224: \end{equation}
2225: \end{itemize}
2226: 
2227: This semiclassical quantization scheme is expected to work provided
2228: the classical phase space velocity is sufficiently large, see section~\ref{SQ}.
2229: This may fail close to the stable and unstable fixed points where
2230: the velocity vanishes. Near the stable equilibrium
2231: point, the expansion of the Hamiltonian at second order leads to an
2232: approximate harmonic Hamiltonian with frequency:
2233: \begin{equation}
2234: \omega_{\mathrm harm} = \sqrt{|\lambda V_1(\hat{I_1})H^{''}_0(\hat{I_1})|}.
2235: \label{omega_harmonic}
2236: \end{equation}
2237: 
2238: In this harmonic approximation, the semiclassical quantization
2239: is known to be exact~\cite{landau2}.
2240:  Thus, close to the stable
2241: equilibrium, the quasi-energy
2242: levels, labeled by the non-negative integer $N$, are
2243: given by the harmonic approximation. There are various cases which
2244: depend on the signs
2245: of $H^{''}_0(\hat{I_1})$, $V_1(\hat{I_1})$, and $\lambda$,
2246:  with the general
2247: result
2248: given by
2249: \begin{equation}
2250: {\mathcal E}_{N,k} = H_0(\hat{I_1}) - \omega \hat{I_1} +
2251: \left(k+\frac{\mu}{4}\right)\hbar\omega
2252: -\ {\mathrm sign}(H^{''}_0(\hat{I_1}))
2253: \left[ |\lambda V_1(\hat{I_1})| -
2254: \left(N+\frac{1}{2}\right) \hbar \omega_{\mathrm harm} \right].
2255: \label{spectrum_harmonic}
2256: \end{equation}
2257: 
2258: For $N=0,k=0,$ this gives
2259:  a fairly accurate estimate of the
2260: energy of the non-dispersive wave-packet with optimum localization.
2261: The EBK semiclassical scheme provides us also with some
2262: interesting information on the eigenstate. Indeed, the invariant
2263: tori considered here are tubes surrounding the resonant
2264: stable periodic orbit. They cover the $[0,2\pi]$ range
2265: of the $t$ variable but are well localized in the
2266: transverse $(\hat{I}=I,\hat{\theta}=\theta-\omega t)$ plane,
2267: with an approximately Gaussian phase space distribution.
2268: Hence, at any fixed time $t$, the Floquet
2269: eigenstate will appear as a Gaussian distribution
2270: localized around the point $(I=\hat{I_1},\theta=\omega t).$ 
2271: As this point precisely defines the resonant, stable
2272: periodic orbit, one expects the $N=0$ state to be
2273: a Gaussian wave-packet following the classical orbit.
2274: In the original $(p,z)$ coordinates, the width of the
2275: wave-packet will depend on the system under consideration
2276: through the change of variables $(p,z)\to (I,\theta),$
2277: but the Gaussian character is expected to be approximately valid
2278: for both the phase space density and the configuration space
2279: wave-function, as long as the change of variables is smooth.
2280: 
2281: Let us note that low-$N$ states may be considered as excitations
2282: of the $N=0$ ``ground'' state. Such states has been termed
2283: ``flotons'' in \cite{holthaus95} where
2284: their wave-packet character was, however, not considered.
2285: 
2286: The number of eigenstates trapped within the resonance island -- i.e.
2287: the number of non-dispersive wave-packets -- is easily evaluated
2288: in the semiclassical limit, as it is the maximum $N$ with librational
2289: motion.  It is roughly the area of the
2290: resonance island, eq.~(\ref{area}), divided by $2\pi\hbar$:
2291: \begin{equation}
2292: {\mathrm Number\ of\ trapped\ states} \simeq \frac{8}{\pi \hbar}
2293: \sqrt{\left|\frac{\lambda V_1(\hat{I_1})}{ H^{''}_0(\hat{I_1})}\right|}.
2294: \label{number_of_trapped_states}
2295: \end{equation}
2296: 
2297: Near the unstable fixed point -- that is at the energy which separates
2298: librational and rotational motion -- the semiclassical quantization fails 
2299: because of the critical slowing down in its  vicinity~\cite{marion}.
2300: The corresponding
2301: quantum states -- known as separatrix states~\cite{leopold94} -- are
2302: expected to be dominantly
2303: localized near the unstable fixed point, simply because the classical
2304: motion there slows down, and the pendulum spends more time close to its
2305: upright position. This localization is once again of purely classical origin,
2306: but not perfect: some part of the wave-function
2307: must be also localized 
2308: along the separatrix, which autointersects
2309: at the hyperbolic fixed point. Hence, the Floquet eigenstates associated with
2310: the unstable fixed points
2311: are not expected to form non-dispersive wave-packets with optimum
2312: localization (see also sec.~\ref{scars}).
2313: 
2314: \subsubsection{The Mathieu approach}
2315: \label{section_mathieu}
2316: The pendulum approximation, eq.~(\ref{eqpend}), for a resonantly driven system can also be found
2317: by a pure quantum description~\cite{berman77,holthaus95}. Let us consider
2318: a Floquet state of the system.
2319: Its spatial part can
2320: be expanded in the eigenbasis of the unperturbed Hamiltonian $H_0$,
2321: \begin{equation}
2322: H_0 |\phi_n\rangle = E_n|\phi_n\rangle ,
2323: \end{equation}
2324: while the time-periodic wave-function can be
2325: expanded in a Fourier series. One obtains:
2326: \begin{equation}
2327: |\psi(t)\rangle =\ \sum_{n,k} c_{n,k}
2328: \ \exp (-ik\omega t)\ \ |\phi_n\rangle,
2329: \label{wfmat}
2330: \end{equation}
2331: where the coefficients $c_{n,k}$ are to be determined.
2332: The Schr\"odinger equation for the Floquet states (eigenstates of ${\cal H}$),
2333: eq.~(\ref{calhq}), with quasi-energies ${\cal E}$ and 
2334: time dependence (\ref{solse}),
2335: reads:
2336: \begin{equation}
2337: c_{n,k}({\cal E}+k\hbar \omega - E_n) = \frac{\lambda}{2} \ \sum_{p}
2338: {\langle \phi_n|V|\phi_{n+p}\rangle\ (c_{n+p,k+1} + c_{n+p,k-1})}.
2339: \label{coupled_equations}
2340: \end{equation}
2341: 
2342: For $\lambda=0,$ the solutions of eq.~(\ref{coupled_equations})
2343: are trivial: ${\cal E}=E_n-k\hbar\omega$, which is nothing
2344: but the unperturbed energy spectrum modulo $\hbar\omega.$
2345: \footnote{Note that, as a consequence of the negative sign of the argument 
2346: of the exponential factor in eq.~(\ref{wfmat}), the energy shift $k\hbar\omega$ 
2347: appears here with a negative sign in the expression for $\cal E$ -- in contrast 
2348: to semiclassical expressions alike eq.~(\ref{spectrum_harmonic}), where we chose
2349: the more suggestive positive sign. Since $k=-\infty\ldots +\infty$, both conventions 
2350: are strictly equivalent.}
2351: In the presence of a small perturbation, only quasi-degenerate states
2352: with
2353: values close to
2354: $E_n-k\hbar\omega$ will be efficiently coupled. In the semiclassical limit,
2355: see section \ref{SQ}, eq.~(\ref{cprinc}),
2356: the unperturbed eigenenergies $E_n,$ labeled by a non-negative
2357: integer,
2358: are locally approximately spaced by $\hbar \Omega,$
2359: where $\Omega$ is the frequency of the unperturbed classical motion.
2360: Close to 
2361: resonance, $\Omega \simeq \omega$ (eq.~(\ref{resonance_condition})), and 
2362: thus: 
2363: \begin{equation}
2364: E_{n}-k \hbar \omega \simeq E_{n+1} - (k+1) \hbar \omega \simeq  E_{n+2} -
2365: (k+2)\hbar \omega\simeq  ...,
2366: \label{selrule}
2367: \end{equation}
2368: so that only states with the same value of $n-k$ will be efficiently coupled.
2369: 
2370: The first approximation is thus to neglect the couplings which do not
2371: preserve $n-k.$ This is just the quantum version of the
2372: secular approximation for the classical dynamics.
2373: Then, the set of equations~(\ref{coupled_equations})
2374: can be rearranged in independent blocks, each subset being characterized by $n-k$.
2375: The various subsets are in fact identical, except for a shift in energy
2376: by an integer multiple of $\hbar\omega.$ This is nothing but the
2377: $\hbar\omega$-periodicity  
2378: of the Floquet spectrum already encountered in secs.~\ref{QD} and \ref{sapp}.
2379: As a consequence, we can
2380: consider
2381: the
2382: $n-k=0$ block alone.
2383: 
2384: Consistently, since eq.~(\ref{selrule}) is valid close to the center
2385: of the resonance only, one can expand the quantities of interest in the
2386: vicinity of the center of the resonance, and use semiclassical 
2387: approximations for 
2388: matrix elements of $V.$ Let $n_0$ denote the effective, resonant quantum number
2389: such that, with eqs.~(\ref{cprinc}) and (\ref{resonance_condition}),
2390: \begin{equation}
2391: \left. \frac{dE_n}{dn} \right|_{n=n_0}= \hbar \omega.
2392: \label{space}
2393: \end{equation}
2394: Note that, by this definition, $n_0$ is {\em not} necessarily an integer.
2395:  In the semiclassical
2396: limit, where $n_0$ is large, the WKB approximation connects $n_0$ to the center
2397: of the classical resonance island, see eq.~(\ref{semen}),
2398: \begin{eqnarray}
2399: &&n_0 + \frac{\mu}{4} = \frac{\hat{I_1}}{\hbar},\\
2400: &&E_{n_0} = H_0(\hat{I_1}),
2401: \end{eqnarray}
2402: with $\mu$ being the Maslov index along the resonant trajectory.
2403: Furthermore, for $n$ close to $n_0,$ we can expand
2404: the unperturbed energy at second order in $(n-n_0)$,
2405: \begin{equation}
2406: E_n \simeq E_{n_0} + (n-n_0) \hbar \omega + \frac{1}{2}
2407: \ \left.\frac{d^2E_n}{dn^2}\right|_
2408: {n_0} \ (n-n_0)^2,
2409: \end{equation}
2410: where
2411: the second derivative $d^2E_n/dn^2$
2412: is directly related to the classical quantity $H^{''}_0$, see
2413: eq.~(\ref{second_derivative}), within the semiclassical WKB
2414: approximation, eq.~(\ref{semen}). Similarly, the matrix elements of $V$
2415: are related to the classical Fourier components
2416: of the potential~\cite{landau2}, eq.~(\ref{fourier_components}),
2417: \begin{equation}
2418: \langle \phi_n|V|\phi_{n+1} \rangle \simeq 
2419: \langle \phi_{n+1}|V|\phi_{n+2} \rangle \simeq V_1(\hat{I_1}),
2420: \label{matel_mat}
2421: \end{equation}
2422: evaluated at the center $\hat{I}=\hat{I}_1$ -- 
2423: see eq.~(\ref{resonance_condition}) --
2424: of the resonance zone.
2425: 
2426: With these ingredients and $r=n-n_0$, eq.~(\ref{coupled_equations}) 
2427: is transformed
2428: in the following set of approximate equations:
2429: \begin{equation}
2430: \left[ {\cal E} - H_0(\hat{I_1}) + 
2431: \omega\left(\hat{I}_1 -\frac{\mu\hbar}{4}\right) - 
2432: \frac{\hbar^2}{2} H^{''}_0(\hat{I_1}) r^2
2433: \right] d_r =
2434: \lambda V_1 (d_{r+1} + d_{r-1}),
2435: \label{tridiag}
2436: \end{equation}
2437: where
2438: \begin{equation}
2439: d_r \equiv c_{n_0+r,n_0+r}.
2440: \label{diag_mat}
2441: \end{equation}
2442: 
2443: Note that, because of eq.~(\ref{space}),
2444: the $r$ values are not necessarily integers, but all have the same
2445: fractional part.
2446: 
2447: The tridiagonal set of coupled equations (\ref{tridiag})
2448: can be rewritten as a differential
2449: equation. Indeed, if one introduces the following function associated with
2450: the Fourier
2451: components $d_r$,
2452: \begin{equation}
2453: f(\phi) = \sum_r{\exp (ir\phi) d_r},
2454: \label{dual}
2455: \end{equation}
2456: eq.~(\ref{tridiag}) can be written as
2457: \begin{equation}
2458: \left[- \frac{\hbar^2}{2} H^{''}_0(\hat{I_1}) \frac{d^2 }{d\phi^2} +
2459: H_0(\hat{I_1}) -
2460: \omega\left(\hat{I}_1 -\frac{\mu\hbar}{4}\right)
2461: + \lambda V_1 \cos \phi \right] f(\phi) = {\cal E} f(\phi),
2462: \label{mathieu}
2463: \end{equation}
2464: which is nothing but the quantum version of the pendulum
2465: Hamiltonian,
2466: eq.~(\ref{eqpend}).
2467: Thus, the present calculation is just the purely quantum 
2468: description of the non-linear resonance phenomenon.
2469: The dummy variable
2470: $\phi$
2471: introduced for convenience coincides with the classical angle
2472: variable  $\hat{\theta}$. 
2473: In general, $r$ is not an integer, so that the various
2474: $\exp (ir\phi)$ in eq.~(\ref{dual}) are not periodic functions
2475: of $\phi.$ However, as all $r$ values have the same fractional part,
2476: if follows that $f(\phi)$ must satisfy ``modified'' 
2477: periodic boundary conditions of the form
2478: \begin{equation}
2479: f(\phi + 2\pi) = \exp (-2i\pi n_0) f(\phi).
2480: \label{boundary}
2481: \end{equation}
2482: The reason for this surprising boundary condition is clear:
2483: $r=n-n_0$ is the quantum analog of $\hat{I}-\hat{I}_1.$
2484: In general, the resonant action $\hat{I}_1$ 
2485: is {\em not} an integer or half-integer multiple
2486: of $\hbar$ -- exactly as $n_0$ is not an integer. The semiclassical
2487: quantization, eq.~(\ref{ebkinres}), which expresses
2488: the $\hat{\theta}$ periodicity of the eigenstate,
2489: applies for the $\hat{I}$ variable. When expressed 
2490: in terms of the variable $\hat{I}-\hat{I}_1$,
2491: it contains the additional phase shift 
2492: present in eq.~(\ref{boundary}).
2493:   
2494: Few words of caution are in order: the equivalence
2495: of the semiclassical quantization with the pure
2496: quantum approach holds in the semiclassical limit only,
2497: when the quantum problem
2498: can be mapped on a pendulum problem. In the general case, it is not possible to
2499: define a quantum angle variable \cite{loudon}.
2500: Hence, the quantum treatment presented here
2501: is no
2502: more general or more powerful than the semiclassical treatment. They both rely
2503: on the same approximations and have the same limitations: perturbative
2504: regime (no overlap of resonances) and semiclassical approximation.
2505: 
2506: Finally, eq.~(\ref{mathieu}) can be written in its standard form, known as
2507: the ``Mathieu equation" \cite{abramowitz72}:
2508: \begin{equation}
2509: \frac{d^2y}{dv^2} + (a-2q\cos 2v) y = 0.
2510: \label{mat_eq}
2511: \end{equation}
2512: The correspondence with eq.~(\ref{mathieu}) is established via:
2513: \begin{eqnarray}
2514: &&\phi = 2v,
2515: \label{map_mathieu1}\\
2516: &&a=\frac{8\left[{\cal E}-H_0(\hat{I_1})+
2517: \omega\left(\hat{I}_1 -\frac{\mu\hbar}{4}\right)
2518: \right]}{\hbar^2 H^{''}_0(\hat{I_1})},
2519: \label{map_mathieu2}\\
2520: &&q=\frac{4\lambda
2521: V_1}{\hbar^2 H^{''}_0(\hat{I_1})}.
2522: \label{map_mathieu3}
2523: \end{eqnarray}
2524: 
2525: The boundary condition, eq.~(\ref{boundary}),
2526: is fixed by the so-called ``characteristic exponent"
2527: in the Mathieu equation,
2528: \begin{equation}
2529: \nu = -2 n_0\ \ \ ({\mathrm mod}\ 2).
2530: \label{cexp}
2531: \end{equation}
2532: 
2533: The Mathieu equation has solutions (for a given characteristic exponent)
2534: for a discrete set of values of $a$ only. That implies quantization of
2535: the quasi-energy levels, according to eq.~(\ref{map_mathieu2}).
2536: The quantized values $a_{\kappa}(\nu,q)$
2537: depend on $q$ and $\nu$, and are labeled\footnote{In 
2538: the standard text books as \protect\cite{abramowitz72},
2539: the various solutions of the Mathieu equation are divided in 
2540: odd and even solutions, and furthermore in $\pi$- and $2\pi$-periodic 
2541: functions. In our case, only the
2542: ``$a_{2p}$" and ``$b_{2p}$" (in the language of \protect\cite{abramowitz72})
2543: are to be considered.} 
2544: by a non-negative
2545: integer $\kappa.$
2546: They are well known
2547: -- especially asymptotic expansions are available both in the
2548: small and in the large $q$ regime -- and
2549: can be found in standard handbooks
2550: \cite{abramowitz72}. For example, fig.~\ref{fig_mathieu}(a) shows the
2551: first $a_{\kappa}(q)\equiv a_{\kappa}(\nu=0,q)$ curves for the case of 
2552: ``optimal" resonance (see below), where
2553: $n_0$ is an integer and thus
2554: the characteristic exponent $\nu$ vanishes.
2555: Equivalently, the figure can be interpreted as the evolution of the
2556: energy levels of a pendulum
2557: with the gravitational field.
2558: 
2559: 
2560: \psfull
2561: \begin{figure}
2562: \centerline{\psfig{figure=bdzf09a.ps,width=7cm,angle=-90}
2563: \psfig{figure=bdzf09b.eps,width=7cm,angle=-90}}
2564: \caption{Eigenvalues $a_{\kappa}(q)$ of the Mathieu equation, 
2565: for a characteristic
2566: exponent $\nu=0.$ These represent the energy levels of a pendulum as a function
2567: of the gravitational field, see
2568: eqs.~(\protect\ref{eqpend},\protect\ref{mathieu}). 
2569: (a): Eigenvalues in the range $[-2q,2q]$ are
2570: associated with the librational bounded motion of the pendulum, while
2571: eigenvalues above $2q$ are associated with rotational modes.
2572: The dotted lines represent the energies of the stable equilibrium
2573: point (lower line) and of the unstable equilibrium point (separatrix, upper
2574: line). Near the separatrix, the classical motion slows down and the
2575: quantal energy levels get closer. (b): Details for the first excited states
2576: together with the semiclassical WKB prediction for the energy levels
2577: (dashed lines, eqs.~(\protect\ref{semen})). 
2578: The semiclassical prediction is very accurate, except
2579: in the vicinity of the separatrix.}
2580: \label{fig_mathieu}
2581: \end{figure}
2582: 
2583: The quasi-energy levels of the driven
2584: system can now be expressed as a function of $a_{\kappa}(\nu,q):$
2585: \begin{equation}
2586: {\cal E}_{\kappa} = H_0(\hat{I_1}) - 
2587: \omega\left(\hat{I}_1 -\frac{\mu\hbar}{4}\right)
2588: + \frac{\hbar^2}{8} H^{''}_0(\hat{I_1})\ a_{\kappa}(\nu,q).
2589: \label{prediction_mathieu}
2590: \end{equation}
2591: Together with eqs.~(\ref{map_mathieu1})-(\ref{map_mathieu3}),
2592: this equation gives the quasienergy
2593: levels of a periodically driven system in the vicinity of the resonance zone.
2594: The full, quasi-resonant part of the Floquet spectrum of the driven system is built
2595: from these quantized values through shifts $k\hbar\omega,$ with
2596: arbitrary integer values of $k$.
2597: 
2598: A visual inspection of fig.~\ref{fig_mathieu} immediately shows the existence
2599: of two regions in the energy diagram: within the ``inner region",
2600: $|a_{\kappa}(\nu,q)| \leq 2q,$ the energy levels
2601: form a regular fan of curves 
2602: and tend to decrease with $q.$ On the contrary,
2603: for $a_{\kappa}(\nu,q) > 2q,$ the energy levels increase with $q$. 
2604: Around $a_{\kappa}(\nu,q)=2q,$
2605: a transition region is visible with a series of apparent avoided crossings
2606: between the levels. This has a simple semiclassical explanation.
2607: The stable fixed point of the pendulum described by the Mathieu
2608: equation (\ref{mat_eq}) lies at $v=0$ with energy $-2q$,
2609: what explains why
2610: the $a_{\kappa}(\nu,q)$ values are always larger. The unstable
2611: fixed point has an energy $+2q.$ Thus, in the range 
2612: $a_{\kappa}(\nu,q) \in [-2q,2q],$
2613: the pendulum
2614: is trapped in a region of
2615: librational motion. The energy levels can be approximated
2616: using the standard WKB quantization in the $(\hat{I},\hat{\theta)}$
2617: plane, as described in section~\ref{CD}. 
2618: The number of such states is given by eq.~(\ref{number_of_trapped_states})
2619: which can be rewritten, using eq.~(\ref{map_mathieu3}), as :
2620: \begin{equation}
2621: {\mathrm Number\ of\ trapped\ states} \simeq \frac{4\sqrt{|q|}}{\pi}.
2622: \label{number_of_trapped_states_2}
2623: \end{equation}
2624: At the center of the island (states with small $\kappa$
2625: and/or large $|q|$), an
2626: approximate expression for $a_{\kappa}(\nu,q)$ reads \cite{abramowitz72}:
2627: \begin{equation}
2628: a_{\kappa}(\nu,q) \approx -2|q|+4\left(\kappa+\frac{1}{2}\right) \sqrt{|q|}.
2629: \label{math_asy}
2630: \end{equation}
2631: It does not depend on $\nu$ (what physically means in our case 
2632: that the states deeply inside the resonance island are insensitive
2633: to the boundary condition).
2634: When inserted in eq.~(\ref{prediction_mathieu}), it
2635: yields exactly the energy levels of eq.~(\ref{ebkires}), with
2636: $\kappa=N.$ Hence, the
2637: Mathieu approach agrees with the harmonic approximation within the
2638: resonance island, for sufficiently large islands.
2639:  
2640: For $a_{\kappa}(\nu,q) > 2q,$ the
2641: pendulum undergoes
2642: a rotational, unbounded motion, which again can be quantized using WKB.
2643: For small $|q|$, 
2644: $a_{\kappa}(\nu,q)\approx 4([(\kappa+1)/2]-\nu/2)^2$ (where
2645: $[\ ]$ stands for the integer part), see \cite{abramowitz72}. 
2646: With this  in
2647: eq.~(\ref{prediction_mathieu}), 
2648: one recovers the known Floquet spectrum, in the limit of a
2649: vanishingly  weak perturbation. 
2650: Note, however, that in this rotational mode, the eigenstates
2651: {\em are} sensitive to the boundary conditions (and the $a_{\kappa}(\nu,q)$
2652: values depend on $\nu$). This is essential for
2653: the correct $\lambda \rightarrow 0$ limit.
2654: 
2655:  
2656: Around $a_{\kappa}(\nu,q) = 2q,$ 
2657: the pendulum is close to the separatrix between
2658: librational and rotational motion: the period of the classical motion
2659: tends to infinity (critical slowing down). That explains the locally enhanced
2660: density of states apparent in fig.~\ref{fig_mathieu}(a).
2661: 
2662: In fig.~\ref{fig_mathieu}(b), we also plot the semiclassical WKB
2663: prediction for the
2664: quantized $a_{\kappa}(q)$ values. 
2665: Obviously, the  agreement with the exact ``quantum"
2666: Mathieu result is very good, even for weakly excited states, except
2667: in the vicinity of the separatrix. This is not unexpected because the
2668: semiclassical
2669: approximation is known
2670: to break down near the unstable fixed point, see
2671: sections~\ref{SQ} and \ref{sapp} above.
2672: 
2673: The Mathieu equation
2674: yields accurate predictions for
2675: properties of
2676: non-dispersive wave-packets in  periodically driven
2677: quantum systems. Indeed, in the range  $a_{\kappa}(\nu,q) \in [-2q,2q],$ 
2678: the classical
2679: motion is trapped inside the resonance island, and the corresponding
2680: quantum eigenstates are expected to be non-dispersive wave-packets.
2681: In particular,
2682: the lowest state in the resonance island, associated
2683: with the ground state of the pendulum $\kappa=0$, 
2684: corresponds to the semiclassical eigenstate $N=0$, see eq.~(\ref{ebkires}), 
2685: and represents 
2686: the non-dispersive wave-packet
2687: with the best localization properties.
2688: As can be seen in fig.~\ref{fig_mathieu}(b), the
2689: semiclassical quantization for this state is in excellent agreement with the
2690: exact Mathieu result. This signifies that eq.~(\ref{spectrum_harmonic})
2691: can be used for {\em quantitative} predictions of the quasi-energy of this
2692: eigenstate.
2693: 
2694: As already mentioned,
2695: the characteristic exponent $\nu$ does not play a major role inside the
2696: resonance island, as the eigenvalues  $a_{\kappa}(\nu,q)$ there depends
2697: very little on $\nu.$ However, it is an important parameter outside the
2698: resonance, close to the separatrix especially at small $q.$ Indeed,
2699: at $q=0$, the minimum eigenvalue is obtained
2700: for $\nu=0: a_{0}(\nu=0,q=0)=0.$ 
2701: This implies that, even for very small $q,$ the ground state
2702: enters most rapidly the resonance island. On the opposite, the
2703: worst case is $\nu=1$ where the lowest eigenvalue is doubly degenerate:
2704: $a_{0}(\nu=1,q=0)=a_{1}(\nu=1,q=0)=1.$ 
2705: As we are interested in the ground state of the
2706: pendulum (the one with maximum localization), 
2707: the situation for $\nu=0$ is preferable: not only the state enters rapidly the
2708: resonance island, but it is also separated from the other states by an
2709: energy gap and is thus more robust versus any perturbation.
2710: We will call this situation ``optimal resonance". From eq.~(\ref{cexp}),
2711: it is associated with an integer value of $n_0.$ On the contrary, a
2712:  half-integer value of $n_0$ corresponds to $\nu=1$ and the
2713:  least optimal case. 
2714: 
2715: 
2716: \subsection{Rydberg states in external fields}
2717: \label{RSEF}
2718: 
2719: \subsubsection{Rydberg atoms}
2720: 
2721: In order to construct non-dispersive wave-packets, a quantum
2722: system subject to periodic driving with classically non-linear dynamics is
2723: needed. The latter requirement rules out the harmonic oscillator,
2724: and all its variants. The simplest periodic driving is certainly
2725: provided by
2726: an externally applied, monochromatic electromagnetic field.
2727: Extremely stable, tunable and well controlled sources
2728: exist over a wide range of frequencies.
2729: 
2730: Furthermore, incoherent
2731: processes which destroy the phase coherence of the quantum wave-function have
2732: to be minimized. Otherwise, they will spoil
2733: the localization properties
2734: of the non-dispersive wave-packets and -- in the worst case -- destroy
2735: them completely. Therefore, the characteristic time scales of the
2736: incoherent processes should be at least much longer than the period of the
2737: driving field.
2738: In this respect, atomic electrons appear as
2739: very good candidates,
2740: since -- given suitable experimental conditions -- atoms can be
2741: considered as practically isolated from the external world, with
2742: spontaneous emission of photons as the only incoherent process.
2743: Spontaneous emission is usually a very slow 
2744: mechanism, especially for highly
2745: excited states: the spontaneous life-time of typical atomic states
2746: is at least four or five orders
2747: of magnitude longer than the classical Kepler period (typically
2748: nanoseconds vs. femtoseconds for weakly excited states~\cite{bethe77}).
2749: 
2750: The Coulomb interaction between  
2751: the nucleus and the
2752: electrons is highly
2753: non-linear, which is very favourable. The efficiency
2754: of the coupling with an external electromagnetic
2755: field is known to increase rapidly with the degree of excitation of the atom
2756: \cite{bethe77}.
2757: As we have seen in the preceeding sections~\ref{CD} and \ref{QD},
2758: non-dispersive wave-packets are the quantum mechanical counterparts
2759: of nonlinear resonances in periodically driven Hamiltonian systems, where
2760: the period of the drive matches some intrinsic time scale of the unperturbed
2761: Hamiltonian dynamics.
2762: Due to the immediate correspondence between the classical Kepler problem
2763: and the hydrogen atom, the relevant time scale in this simplest atomic system
2764: is the
2765: unperturbed classical Kepler period, which 
2766: -- compare eq.~(\ref{cprinc}) --
2767: coincides with the inverse level spacing between neighbouring
2768: eigenstates of the unperturbed atom, for large quantum numbers.
2769: Hence, the driving field frequency has to be chosen resonant
2770: with an atomic transition
2771: in the Rydberg regime, typically around the
2772: principal quantum number $n_0=60$. This is the microwave domain,
2773:  where excellent sources exist.
2774: Thus, we believe that atomic Rydberg states are very well suited
2775: for the experimental preparation of non-dispersive wave-packets\footnote{Note,
2776: however, that this is a specific choice. Any driven quantum system 
2777: with a sufficiently high denisty of states and mixed regular-chaotic 
2778: classical dynamics will exhibit nondispersive wave-packets. Since a mixed 
2779: phase space structure is the generic scenario for dynamical systems,
2780: nondispersive wave-packets are expected to be a completely general and 
2781: ubiquitous phenomenon.}.
2782: 
2783: In most cases, the energy scale involved in the dynamics of Rydberg
2784: electrons is so small that the inner electrons of the ionic core
2785: can be considered as frozen and ignored. Thus, we will consider
2786: mainly the hydrogen atom as the simplest prototype. Multi-electron
2787: effects are discussed in section~\ref{HE}. Alternative systems
2788: for observing non-dispersive wave-packets are considered in section
2789: \ref{OGB}.
2790: 
2791: Compared to our simple one-dimensional model introduced in section~\ref{GM} 
2792: above,
2793: an {\em atom}
2794: displays a couple of additional features:
2795: \begin{itemize}
2796: \item A real hydrogen atom is a three-dimensional (3D) system. 
2797: However, it is a degenerate system, 
2798: because the energy depends on the principal quantum number
2799: $n$ only, but not on the angular or magnetic quantum numbers $L$ and $M$,
2800: respectively.
2801: Thus, the structure of the energy levels, which is crucial for the
2802: properties of the non-dispersive wave-packet, see section \ref{QD},
2803: is identical in the 1D and 3D cases. Before discussing the properties
2804: of 3D wave-packets in sections~\ref{LIN3D} to \ref{MA}, we will
2805: consider a simplified 1D model of the hydrogen atom in section~\ref{LIN1D}.
2806: \item Although spontaneous emission is a weak incoherent process,
2807: it 
2808: nontheless limits the life time of the non-dispersive wave-packets which may
2809: decay to lower lying states, losing the phase coherence
2810: of the electronic wave-function. Non-dispersive wave-packets
2811: exhibit specific spontaneous decay properties which are studied in
2812: section~\ref{SPO};
2813: \item The electron in a hydrogen atom is not necessarily bound.
2814: It may ionize, especially when the atom is exposed to
2815: a microwave field. This is a {\em coherent} decay process 
2816: where the ionized electron keeps its phase coherence.
2817: There are no exact bound states in the system, but rather resonances.
2818: From a quantum point of view, the Floquet spectrum is no longer
2819: discrete but continuous and we actually deal with an {\em open system}. 
2820: In a more elementary language,
2821: the atom can successively absorb several photons so that its energy
2822: exceeds the ionization threshold. 
2823: If initially prepared in a wave-packet eigenstate, this is a pure quantum
2824: phenomenon, since the classical dynamics remain trapped within the
2825: resonance island forever. The multi-photon ionization 
2826: may then be 
2827: considered as a tunneling process from 
2828: inside the resonance island
2829: to the non-resonant part of phase space, 
2830: where the Rydberg electron eventually escapes
2831: to infinity. 
2832: This picture is elaborated in  section~\ref{ION}.
2833: \end{itemize}
2834: 
2835: \subsubsection{Hamiltonian, basis sets and selection rules}
2836: 
2837: In the presence of a microwave field, the dipole 
2838: approximation~\cite{cct92,loudon}
2839: can be used to describe  the atom-field interaction. Different gauges
2840: can be used, the physics being of course independent of the choice
2841: of gauge. The most common choices are the length and the velocity gauges.
2842: For simplicity, in our discussion,
2843: we shall use the length gauge, although actual
2844: quantum calculations are usually a bit easier in the velocity gauge
2845: \cite{shakeshaft88,abu95c,cormier96}.
2846: The Hamiltonian reads:
2847: \begin{equation}
2848: H = \frac{\vec{p}^2}{2m} - \frac{q^2}{4\pi \epsilon_0 r}
2849: -q \vec{r}\cdot\vec{F}(t)
2850: \label{ham_3d_si}
2851: \end{equation}
2852: where $q$ is the (negative) charge of the electron, $m$ its mass, and
2853: $\vec{F}(t)$ the microwave electric field acting on the atom.
2854: We neglect here all relativistic, spin, QED effects, etc, and assume
2855: an infinitely massive nucleus.
2856: Later, unless specified otherwise, we shall use atomic units, where
2857: $|q|\equiv 4\pi\epsilon_0\equiv m\equiv \hbar\equiv 1$, and the
2858: Hamiltonian reduces to
2859: \begin{equation}
2860: H = \frac{\vec{p}^2}{2} - \frac{1}{r}
2861: +\vec{r}\cdot\vec{F}(t).
2862: \label{h}
2863: \end{equation}
2864: Here, 
2865: \begin{equation}
2866: H_0= \frac{\vec{p}^2}{2} - \frac{1}{r}
2867: \label{h0}
2868: \end{equation}
2869: describes the unperturbed atomic part, with 
2870: bound energy spectrum: 
2871: \begin{equation}
2872: E_n= - \frac{1}{2n^2},\ n\geq 1.
2873: \label{spectrum}
2874: \end{equation}
2875: The external driving
2876: \begin{equation}
2877: V = \vec{r}\cdot \vec{F}(t)
2878: \label{v}
2879: \end{equation}
2880: is characterized by
2881: the amplitude $F$ of the microwave field, and by its  
2882: frequency
2883: $\omega.$ 
2884: We shall consider in detail the case of linear polarization, 
2885: \begin{equation}
2886: V=F z \cos \omega t,
2887: \label{vlinear} \end{equation}
2888: in section \ref{LIN}, and that of circular polarization
2889: \begin{equation}
2890: V=F( x \cos \omega t +  y \sin \omega t),
2891: \label{vcirc} 
2892: \end{equation}
2893: in section~\ref{CP}. The general case of elliptic polarization
2894: will be 
2895: studied in section~\ref{EP}.
2896: 
2897: In the treatment of the perturbed Coulomb problem, eq.~(\ref{h}),
2898: there is an apparent difficulty with the singularity of the Coulomb
2899: potential at the origin, especially for the restricted one-dimensional
2900: model of the atom.
2901: This led several authors,
2902: see, e.g.,
2903: \cite{javanainen88,su91,reed91,burnett92,grobe93,grobe94}, to consider
2904: unphysical potentials without singularity, for example
2905: of the type $(r^2+a^2)^{-1/2}.$ This is completely unnecessary
2906: and potentially dangerous. Indeed, such a potential breaks
2907: the Coulomb degeneracy which is responsible for the closed character
2908: of elliptical Kepler orbits (in the classical world), and
2909: for the degeneracy of the energy levels (in the quantum world).
2910: Such an unphysical symmetry breaking strongly modifies the structure
2911: of the non-linear resonance island, and affects the existence
2912: and properties of non-dispersive wave-packets outlined in section~\ref{LIN3D}.
2913: 
2914: 
2915: The Coulomb singularity can be rigorously regularized (in any
2916: dimension), both in classical and in quantum mechanics. In classical
2917: mechanics, this is made possible through the well-known
2918: Kustaanheimo-Stiefel transformation \cite{kusta65}, used by
2919: various authors for perturbed Coulomb problems, see, e.g.,
2920: \cite{delande84,rath88,griffiths92,gebarowski95,schlagheck99}.
2921:  In quantum mechanics, one may  use a basis set
2922: of non-orthogonal functions, known as the Sturmian functions.
2923: Ultimately, the whole analysis relies on the dynamical symmetry
2924: properties of the Coulomb interaction and the associated SO(4,2) group
2925: \cite{barut79,chen80,chen81,delandeth}.
2926: It not only allows to treat the Coulomb singularity properly, but also
2927: to define a basis set of Sturmian functions 
2928:  extremely efficient for
2929: numerical calculations. The most common set 
2930: are ``spherical" Sturmian functions characterized by three quantum numbers,
2931: $L$ and $M$ for the angular structure (the associated wave-functions are
2932: the usual spherical harmonics), and the positive integer $n$, for the radial
2933: part. 
2934: 
2935: As discussed in section~\ref{QD}, the quantum properties
2936: of the non-dispersive wave-packets are 
2937: encoded in the
2938: spectrum of the Floquet Hamiltonian (\ref{calhq}), which acts
2939: in configuration space extended by the time axis. The temporal
2940: properties are completely independent of the spatial dimension,
2941: and any Floquet eigenstate can be expanded in a Fourier series
2942: indexed by the integer $k$, 
2943: as in eq.~(\ref{wfmat}). The Schr\"odinger equation
2944: for the Floquet Hamiltonian is tridiagonal in $k,$ as in
2945: eq.~(\ref{coupled_equations}).
2946: When the spatial part of the wave-function is 
2947: expressed in a Sturmian basis,
2948: one finally obtains
2949: a generalized eigenvalue problem $(A-{\cal E} B)|\psi> = 0$, where both,
2950: $A$ and $B$, are sparse matrices, the elements of which are known analytically
2951: and obey the following selection rules:
2952: \begin{equation}
2953: |\Delta M|\leq 1,\ \ |\Delta L| \leq 1,\ \ |\Delta n| \leq 2,\ \ |\Delta k| \leq 1.
2954: \end{equation}
2955: When a static electric or magnetic field is added, see section~\ref{MA},
2956: some additional non-zero matrix elements exist, but sparsity is preserved.
2957: The eigenvalues can then be calculated using an efficient diagonalization
2958: routine such as the Lanczos 
2959: algorithm \cite{lanczos,delande91,ericsson80,grimes94}.
2960: 
2961: Because -- in the presence of a microwave field -- the system is unbounded,
2962: there are in general no exact bound states but rather resonances.
2963: Using Sturmian functions, the
2964: properties of the resonances can be calculated {\em directly} 
2965: using the complex
2966: rotation technique 
2967: \cite{balslev71,graffi85,yajima82,nicolaides78,reinhardt83,ho83,moiseyev98}. 
2968: The price
2969: to pay is to diagonalize complex symmetric matrices, instead of
2970: real symmetric ones. The advantage is that the resonances are obtained
2971: as complex eigenvalues $E_n-i\Gamma_n/2$ of the complex rotated Hamiltonian,
2972: $E_n$ being the position of the resonance, and $\Gamma_n$ its width.
2973: All
2974: essential properties of resonances can be obtained from complex
2975: eigenstates \cite{abu94}.
2976: 
2977: \subsubsection{Simplified 1D and 2D models}
2978: 
2979: Because explicit calculations for the real 3D
2980:  hydrogen atom may be
2981: rather complicated, it is fruitful to study also simplified 
2982: 1D and 2D approximations of the real world.
2983: Let us first consider the simplified
2984: restriction of the atomic motion to one single dimension 
2985: of configuration space,
2986: \begin{equation}
2987: H_0=\frac{p^2}{2} - \frac{1}{z}, \ \ \ \ {\mathrm with}\ z>0,
2988: \label{h0_1d}
2989: \end{equation}
2990: with the external driving along $z$,
2991: \begin{equation}
2992: V = F z \cos \omega t.
2993: \label{v_1d}
2994: \end{equation}
2995: The energy spectrum of $H_0$ is identical to the spectrum
2996: in 3D~\cite{englefield72}:
2997: \begin{equation}
2998: E_n^{(1D)} = - \frac{1}{2n^2},\ \ \ \ {\mathrm with}\ n\geq 1.
2999: \label{spectrum_1d}
3000: \end{equation}
3001: Such  a one-dimensional model allows to grasp
3002: essential features of the driven atomic dynamics, and provides the simplest
3003: example for the creation of non-dispersive wave-packets by a near-resonant
3004: microwave field. The classical dynamics live on a three-dimensional
3005: phase space, spanned by the single dimension of configuration space, the canonically conjugate momentum, and
3006: by time. This is the lowest dimensionality for a Hamiltonian system to display
3007: mixed regular-chaotic character \cite{lichtenberg83}.
3008: 
3009: For a circularly (or elliptically) polarized microwave, a 1D model
3010: is of course inadequate. One can use a two-dimensional model
3011: where the motion of the electron is restricted to the polarization
3012: plane. The energy spectrum in two dimensions is:
3013: \begin{equation}
3014: E_n^{(2D)}= - \frac{1}{2\left(n+\half\right)^2},\ \ \ \ {\mathrm with}\ n\geq 0.
3015: \label{spectrum_2d}
3016: \end{equation}
3017: It differs from the 3D (and 1D) energy spectrum by the additional
3018: 1/2 in the denominator, due to the specific
3019: Maslov indices induced by  the Coulomb 
3020: singularity.
3021: 
3022: \subsubsection{Action-angle coordinates}
3023: 
3024: \label{action-angle_hydrogen}
3025: 
3026: In order to apply the general theory of nonlinear resonances and
3027: non-dispersive wave-packets derived in section~\ref{GM}, we need
3028: the action-angle coordinates for the hydrogen atom.
3029: For the simplified 1D model, the result is simple:
3030:  the principal action $I$ and the canonically conjugate
3031: angle $\theta$ are defined by \cite{goldstein80,jensen84}
3032: \begin{eqnarray}
3033: I & = & \sqrt{\frac{a}{2}}, \nonumber \\
3034: \theta & = & \left\{
3035: \begin{array}{lc}
3036: \eta-\sin\eta, & \ \ p\geq 0, \\ 2\pi-\eta+\sin\eta, & \ \ p<0,
3037: \end{array} \right. \nonumber \\
3038: \eta & = & 2\sin^{-1}\sqrt{\frac{z}{a}},\ \ \  a=-E^{-1},
3039: \label{aa_lin1d}
3040: \end{eqnarray}
3041: where $a$
3042: is the maximum distance.
3043: In celestial mechanics, $\theta$ and $\eta$ are known as
3044: the {\em mean} and the {\em eccentric anomaly}, respectively \cite{marion}.
3045: The Hamilton function depends on the action through:
3046: \begin{equation}
3047: H_0=-\frac{1}{2I^2},
3048: \label{h0_I}
3049: \end{equation}
3050: and the classical Kepler frequency reads:
3051: \begin{equation}
3052: \Omega = \frac{dH_0}{dI} = \frac{1}{I^3}.
3053: \label{omk}
3054: \end{equation}
3055: Due to the  Coulomb singularity at $z=0$,
3056:  the Maslov index
3057: of this system is $\mu=0$ instead of $\mu=2$, and the semiclassical 
3058: energy spectrum,
3059:  eq.~(\ref{semen})\footnote{As we are using atomic units, $\hbar$ is 
3060: unity, and
3061: the principal quantum number just coincides with the action.},
3062: matches the exact quantum spectrum, eq.~(\ref{spectrum_1d}).
3063: 
3064: The classical equations of motion can be solved exactly and it is easy
3065: to obtain the Fourier components of the dipole operator \cite{meerson82}:
3066: \begin{equation}
3067: z(\theta) = I^2 \left(\frac{3}{2} - 2\sum_{m=1}^{\infty}{ \frac{J'_m(m)}{m}
3068: \cos (m\theta)} \right),
3069: \label{v_1dfou}
3070: \end{equation}
3071: where $J'_n(x)$ denotes the  derivative of the usual Bessel function.
3072: The strongly non-linear character of the Coulomb interaction is responsible
3073: for the slow decrease of the Fourier components at high $m$.
3074: 
3075: For the 2D and 3D hydrogen atom, the action-angle variables
3076: are similar, but more complicated because
3077: of the existence of angular degrees of freedom.
3078: The classical trajectories are ellipses with focus at the nucleus.
3079: The fact that all bounded trajectories are periodic manifests the
3080: degeneracy of the classical dynamics. As a consequence, although
3081: phase space is six-dimensional with three angle and three action 
3082: variables in 3D
3083: -- four-dimensional with two angle and two action variables in 2D
3084: -- the Hamilton function depends only on the total action $I$, precisely
3085: like the 1D hydrogen atom, i.e. through eq.~(\ref{h0_I}).
3086: In 3D, the Maslov index is zero, so that the energy spectrum is
3087: again given by eq.~(\ref{spectrum}), and the semiclassical
3088: approximation is exact. However, a different
3089: result holds for the 2D hydrogen atom, where the
3090: Maslov index is $\mu=2$ (still yielding exact agreement between the
3091: semiclassical and   the quantum spectrum, cf.
3092: eqs.~(\ref{ebkgen},\ref{spectrum_2d})).
3093: 
3094: The action-angle variables which 
3095: parametrize a general Kepler ellipse  are well known
3096: \cite{goldstein80}.
3097: In addition to the action-angle variables $(I,\theta)$ which determine
3098: the total action and the angular position
3099: of the electron along the Kepler ellipse,
3100: respectively, the orientation of the ellipse in space is defined by 
3101: two angles: $\psi$, canonically conjugate to the total angular
3102: momentum $L$, and the polar angle $\phi$, canonically conjugate to $M$, the
3103: $z$-component of the angular momentum.
3104: The angle $\psi$ conjugate to $L$ has a
3105: direct physical meaning for $M=0$: it represents
3106: the angle between the Runge-Lenz vector $\vec{A}$ (oriented along the major
3107: axis) of the Kepler ellipse, and the
3108: $z$-axis.
3109: For the 2D hydrogen atom (in the $(x,y)$ plane), 
3110: the orientation 
3111: of the ellipse is defined by the angle $\psi$
3112: (canonically conjugate to the total angular
3113: momentum $L$)
3114:  between the Runge-Lenz vector
3115: and the $x$-axis.
3116: 
3117: Also the Fourier components of the unperturbed
3118: classical position operator $\vec{r}(t)$
3119: are well known \cite{casati88}. In the local coordinate
3120: system of the Kepler ellipse (motion in the
3121: $(x',y')$ plane, with major axis along $x'$), one gets:
3122: \begin{eqnarray}
3123: x'&=&-\frac{3e}{2}I^2+2I^2\sum_{m=1}^\infty
3124: \frac{J_{m}^{'}(me)}{m}\cos m\theta \label{v_3d_x} \\
3125: y'&=&2I^2\frac{\sqrt{1-e^2}}{e}\sum_{m=1}^\infty
3126: \frac{J_{m}(me)}{m}\sin m\theta \label{v_3d_y} \\
3127: z'&=&0
3128: \label{v_3d_z}
3129: \end{eqnarray}
3130: where 
3131: \begin{equation}
3132: e = \sqrt{1-\frac{L^2}{I^2}}
3133: \label{eccentricity}
3134: \end{equation}
3135: denotes the eccentricity of the ellipse.
3136: $J_{m}(x)$ and $J^{'}_{m}(x)$ are the ordinary
3137: Bessel function and
3138: its derivative, respectively.
3139: 
3140: In the laboratory frame, the various components can be found
3141: by combining these expressions with the usual Euler
3142: rotations~\cite{marion,goldstein80}. The set of three Euler angles
3143: describes the successive rotations required for the
3144: transformation between the
3145: laboratory frame and the frame $(x',y',z')$ linked to the classical
3146: Kepler ellipse. We choose to rotate successively by an angle $\phi$ 
3147: around the $z$ laboratory axis, an angle $\beta$ around the 
3148: $y$-axis\footnote{Some authors define the second Euler rotation with
3149: respect to the $x$-axis. The existence of the two definitions 
3150: makes a cautious physicist's life much harder,
3151:  but the physics does not -- or at least
3152: should not -- depend
3153: on such ugly details.}, and an angle $\psi$ around the $z'$ axis.
3154: The physical interpretation of $\phi$ and $\psi$ is simple: 
3155: $\phi$ corresponds to a rotation around the $z$-axis, and is thus 
3156: canonically conjugate
3157: to the $z$-component of the angular momentum, noted $M$. Similarly,
3158: $\psi$ corresponds to a rotation around the axis of the total angular
3159: momentum $\vec{L}$, and is thus canonically conjugate
3160: to $L.$ 
3161: By construction, the third angle $\beta$ is 
3162: precisely the angle
3163: between the angular momentum $\vec{L}$ and the $z$-axis. Thus:
3164: \begin{equation}
3165: \cos \beta = \frac{M}{L},
3166: \label{beta}
3167: \end{equation}
3168: and $\beta =\pi/2$ for $M=0$.
3169: Altogether, the coordinates in the laboratory frame are related
3170: to the local coordinates through
3171: \begin{eqnarray}
3172: x&=&(\cos \psi \cos \beta \cos \phi - \sin \psi \sin \phi)\ x' +
3173: (-\sin\psi \cos\beta \cos\phi -\cos\psi \sin\phi)\ y' + \sin \beta \cos\phi\ z',
3174: \nonumber \\
3175: y&=&(\cos \psi \cos \beta \sin \phi + \sin \psi \cos \phi)\ x' +
3176: (-\sin\psi \cos\beta \sin\phi +\cos\psi \cos\phi)\ y' + \sin \beta \sin\phi \ z',
3177: \nonumber \\
3178: z&=& -\cos\psi \sin\beta \ x'  +  \sin\psi \sin\beta \ y' +  \cos\beta \ z',
3179: \label{rot_euler}
3180: \end{eqnarray}
3181: which, combined with eqs.~(\ref{v_3d_x}-\ref{v_3d_z}), allows for a complete
3182: expansion of the classical trajectories in terms of action-angle\
3183: coordinates.
3184: 
3185: The situation is somewhat simpler for the 2D model of the hydrogen atom.
3186: There, the angular momentum $\vec{L}$ is aligned along the $z$-axis,
3187: which means that $L=M$ and $\beta=0.$ Also, the rotation around
3188: $\vec{L}$ by an angle $\psi$ 
3189: can be absorbed in a rotation by an angle $\phi$ around the $z$-axis.
3190: Therefore, one is left with
3191: two pairs $(I,\theta)$ and $(M,\phi)$ of action-angle variables.
3192: The relation between the laboratory and the local coordinates
3193: reads:
3194: \begin{eqnarray}
3195: x&=& \cos \phi \ x' - \sin\phi \ y' 
3196: \nonumber \\
3197: y&=&\sin \phi \ x' + \cos\phi\ y',
3198: \label{twodrotfr} 
3199: \end{eqnarray} 
3200: which is nothing but a rotation of angle $\phi$ in the plane of the
3201: trajectory.
3202: Formally, the 2D result, eq.~(\ref{twodrotfr}), can be obtained from the 3D one,
3203: eq.~(\ref{rot_euler}), by
3204: specializing to $\beta=\psi=0$.
3205: The eccentricity, eq.~(\ref{eccentricity}), of
3206: the trajectory now reads:
3207: \begin{equation}
3208: e = \sqrt{1-\frac{M^2}{I^2}}.
3209: \end{equation}
3210: 
3211: \subsubsection{Scaling laws}
3212: 
3213: \label{scaling_laws}
3214: 
3215: It is well known that the Coulomb interaction 
3216: exhibits particular
3217: scaling properties: for example, all 
3218: bounded trajectories are similar
3219: (ellipses), whatever the (negative) energy. Also, the
3220: classical period 
3221: scales in a well-defined way with the
3222: size of the orbit (third Kepler law). This originates from
3223: the fact that the Coulomb potential is a homogeneous function
3224: -- of degree $-1$ -- of the radial distance $r$. Similarly,
3225: the dipole operator responsible for the coupling between the Kepler 
3226: electron and the 
3227: external driving field is a homogenous function
3228: -- of degree $1$ -- of $r.$ It follows that the
3229: classical equations of motion of the hydrogen
3230: atom exposed to an electromagnetic field
3231: are invariant under the following
3232: scaling transformation:
3233: \begin{eqnarray}
3234: \left \{
3235: \begin{array}{l}
3236: {\vec r}\rightarrow\alpha^{-1} {\vec r}, \\
3237: {\vec p}\rightarrow\alpha^{1/2} {\vec p}, \\
3238: H_0\rightarrow \alpha H_0, \\
3239: t\rightarrow\alpha^{-3/2} t, \\
3240: F\rightarrow\alpha^2 F,  \\
3241: \omega\rightarrow\alpha^{3/2}\omega , \\
3242: V \rightarrow \alpha V.
3243: \end{array}
3244: \right.
3245: \label{scaling_law}
3246: \end{eqnarray}
3247: where $\alpha $ is an arbitrary, positive real number.
3248: Accordingly, the action-angle variables
3249: transform as
3250: \begin{eqnarray}
3251: \left \{
3252: \begin{array}{l}
3253: I \rightarrow \alpha^{-1/2} I,\\
3254: L \rightarrow \alpha^{-1/2} L,\\
3255: M \rightarrow \alpha^{-1/2} M,\\
3256: \theta \rightarrow \theta,\\
3257: \psi \rightarrow \psi,\\
3258: \phi \rightarrow \phi.
3259: \end{array}
3260: \right.
3261: \label{scaling_aa}
3262: \end{eqnarray}
3263: It is therefore useful to introduce the ``scaled" total angular momentum
3264: and its component 
3265: along the $z$-axis, by chosing 
3266: \begin{equation}
3267: \alpha =I^2,
3268: \label{choosescale}
3269: \end{equation}
3270: what leads to
3271: \begin{eqnarray}
3272: \left \{
3273: \begin{array}{l}
3274: L_0 = \frac{L}{I},\\
3275: M_0 = \frac{M}{I}.
3276: \label{scaling_lm}
3277: \end{array}
3278: \right.
3279: \end{eqnarray}
3280: The eccentricity of the classical ellipse then reads:
3281: \begin{equation}
3282: e=\sqrt{1-L_0^2},
3283: \end{equation}
3284: and only depends -- as it should -- on scaled quantities. Similarly,
3285: the Euler angles describing the orientation of the ellipse are
3286: scaled quantities, by virtue of eq.~(\ref{scaling_aa}).
3287: 
3288: When dealing with non-dispersive wave-packets, it will be 
3289: useful
3290: to scale the amplitude and the frequency of the external field
3291: with respect of the action $\hat{I}_1$ of the resonant orbit.
3292: With the above choice of $\alpha$, eq.~(\ref{choosescale}), the scaling 
3293: relation (\ref{scaling_law}) for $\omega$ defines the scaled frequency 
3294: \begin{equation}
3295: \omega_0=\omega I^3\,
3296: \label{omega0}
3297: \end{equation}
3298: which turns into $\omega_0=\Omega \hat{I}_1^3$ with the resonance condition,
3299: eq.~(\ref{resonance_condition}), and enforces
3300: \begin{equation}
3301: \hat{I}_1 = \omega^{-1/3},
3302: \label{scaling_action}
3303: \end{equation}
3304: by virtue of eq.~(\ref{omk}).
3305: Correspondingly, the scaled external field is defined as 
3306: \begin{equation}
3307: F_0=FI^4,
3308: \label{scaling_F}
3309: \end{equation}
3310: which, with eq.~(\ref{scaling_action}), turns into $F_0=F\omega^{-4/3}$ at
3311: resonance.
3312: Hence, except for a global multiplicative factor $I^{-2},$
3313: the Hamiltonian of a hydrogen atom in an external field
3314: depends only on scaled quantities.
3315: 
3316: Finally, note that the {\em quantum} dynamics is not invariant
3317: with respect to the above scaling transformations. Indeed, the Planck constant
3318: $\hbar$ fixes an absolute scale for the various action variables. 
3319: Thus, the spectrum
3320: of the Floquet Hamiltonian will not be scale invariant,
3321: while the underlying classical phase space
3322: structure is. This latter feature 
3323: will be used to identify in the quantum spectrum 
3324: the remarkable features we are interested in.
3325:  
3326: 
3327: \subsection{Rydberg states in linearly polarized microwave fields}
3328: \label{LIN}
3329: 
3330: We are now ready to consider specific examples of non-dispersive
3331: wave-packets. We consider first the simplest, one-dimensional,
3332: driven hydrogen atom, as defined by eqs.~(\ref{h0_1d}),(\ref{v_1d}).
3333: 
3334: For a real 3D atom, this corresponds to driving the electron
3335:  initially prepared in a one-dimensional 
3336: eccentricity one orbit along the polarization axis of the field. 
3337: In fact,
3338: it turns out that the one-dimensionality of the dynamics is not
3339: stable under the external driving: the Kepler ellipse (with
3340: orientation fixed in configuration
3341: space by the Runge-Lenz vector) slowly precesses off the
3342: field polarization axis (see sections~\ref{LIN3D} and \ref{staticmw}).
3343: Thus, the 1D presentation which follows has mostly
3344: pedagogical value -- being
3345: closest to the 
3346: general case discussed later.
3347: However, a one-dimensional model allows to grasp
3348: essential features of the driven atomic dynamics and provides the simplest
3349: example for the creation of non-dispersing wave-packets by a near-resonant
3350: microwave field. 
3351: A subsequent section will describe the dynamics of the real 3D atom under
3352: linearly polarized driving, and amend on the flaws and
3353: drawbacks of the one-dimensional model.
3354: 
3355: \subsubsection{One-dimensional model}
3356: \label{LIN1D}
3357: 
3358: From eqs.~(\ref{h0_1d}),(\ref{v_1d}), the Hamiltonian of the driven 1D atom
3359: reads
3360: \begin{equation}
3361: H=\frac{p^2}{2}-\frac{1}{z}+Fz\cos (\omega t),\ z>0.
3362: \label{ham_lin1d}
3363: \end{equation}
3364: This has precisely the general form, eq.~(\ref{ham}), and we can therefore 
3365: easily
3366: derive explicit expressions for the secular Hamiltonian subject to the
3367: semiclassical quantization conditions, eqs.~(\ref{ebkpt}),
3368: (\ref{ebkires}),(\ref{ebkinres}), as well
3369: as for the quantum mechanical eigenenergies, eq.~(\ref{prediction_mathieu}), in the
3370: pendulum approximation.
3371: With the Fourier expansion, eq.~(\ref{v_1dfou}), and
3372: identifying $\lambda$ and $V(p,z)$ in eq.~(\ref{ham}) with $F$ and $z$ in
3373: eq.~(\ref{ham_lin1d}), respectively, the Fourier coefficients in
3374: eq.~(\ref{fourier_components}) take the explicit form
3375: \begin{equation}
3376: V_0=\frac{3}{2}I^2;\ \ \ \ V_m=-I^2\frac{J_m'(m)}{m},\ m\neq 0.
3377: \label{coeff_lin1d}
3378: \end{equation}
3379: The resonant action -- which defines the position of the resonance island in 
3380: fig.~\ref{pendulum} -- 
3381: is given by eq.~(\ref{scaling_action}).
3382: In a quantum 
3383: description, 
3384: the resonant action coincides with the 
3385: resonant principal quantum number:
3386: \begin{equation}
3387: n_0=\hat{I}_1 = \omega^{-1/3},
3388: \label{acteqqn}
3389: \end{equation}
3390: since the Maslov index vanishes in 1D, see section~\ref{RSEF}, and 
3391: $\hbar\equiv 1$ in atomic units.
3392: The resonant coupling is then given by:
3393: \begin{equation}
3394:  V_1=-\hat{I}^2J_1'(1),
3395: \label{v1_LIN1D}
3396: \end{equation}
3397: and the secular Hamiltonian, eq.~(\ref{hsec_ap}), reads:
3398: \begin{equation}
3399: {\cal H}_{\mathrm sec} = \hat{P}_t - \frac{1}{2\hat{I}^2} 
3400: -\omega \hat{I} -  J_1'(1)\hat{I}^2 F \cos\hat{\theta}.
3401: \label{hsec_h1d}
3402: \end{equation}
3403: This Hamiltonian has the standard form of a secular
3404: Hamiltonian with a resonance island centered around
3405: \begin{equation}
3406: \hat{I}=\hat{I}_1 = \omega^{-1/3}=n_0,\ \ \ \ \hat{\theta} = \pi,
3407: \label{posres}
3408: \end{equation}
3409: sustaining librational motion  within its boundary.
3410: 
3411: Those energy values of ${\cal H}_{\mathrm sec}$ which
3412: define contour lines (see fig.~\ref{pendulum}) 
3413: such that the contour integrals, eqs.~(\ref{ebkires}),(\ref{ebkinres}),
3414: lead to 
3415: non-negative integer values of $N$, are the semiclassical quasienergies
3416: of the 1D hydrogen atom under external driving. The non-dispersive wave-packet
3417: eigenstate of this model atom in the electromagnetic field is represented by
3418: the ground state $N=0$ of ${\cal H}_{\mathrm sec}$,
3419: localized (in phase space) near the center
3420: of the resonance island. A detailed comparison
3421: of the semiclassical energies to the exact quantum solution of our problem
3422: will be provided in the next subsection,
3423: where we treat the
3424: three-dimensional atom in the field. There it will turn out that the
3425: spectrum of the 1D model is actually neatly 
3426: embedded  in the 
3427: spectrum of
3428: the real 3D atom.
3429: 
3430: In the immediate vicinity of the resonance island, 
3431: the secular Hamiltonian can be further simplified, leading to the
3432: pendulum approximation, see section~\ref{CD} and eq.~(\ref{eqpend}).
3433: The second derivative of the unperturbed Hamiltonian with respect
3434: to the action is:
3435: \begin{equation}
3436: H^{''}_0=-\frac{3}{n_0^4},
3437: \label{h0second}
3438: \end{equation}
3439: and the pendulum Hamiltonian reads:
3440: \begin{equation}
3441: {\cal H}_{\mathrm pend} = \hat{P}_t - \frac{3}{2n_0^2}
3442: - J_1'(1)n_0^2 F \cos\hat{\theta} -\frac{3}{2n_0^4}(\hat{I}-n_0)^2.
3443: \label{hpend_lin1d}
3444: \end{equation}
3445: Remember that $n_0$ is the resonant action, not necessarily an integer.
3446: As we are interested in states deeply inside the resonance island,
3447: we can employ the harmonic approximation around the
3448: stable fixed point $(\hat{I}=n_0,\hat{\theta}=\pi)$, and finally
3449: obtain the semiclassical energies of the non-dispersive
3450: wave-packets:
3451: \begin{equation}
3452: {\cal E}_{N,k} = k\omega -\frac{3}{2n_0^2} + J_1'(1)n_0^2 F -
3453: \left(N+\frac{1}{2}\right) \omega_{\mathrm harm},
3454: \label{semiharm1dval}
3455: \end{equation}
3456: where, in agreement with eq.~(\ref{omega_harmonic})
3457: \begin{equation}
3458: \omega_{\mathrm harm}= 
3459: \sqrt{3J_1'(1)} \frac{\sqrt{F}}{n_0} = \omega \sqrt{3J_1'(1)F_0}
3460: \label{omega_h1d}
3461: \end{equation}
3462: is the classical librational frequency in the resonance island.
3463: The quantum number $k$ reflects the global $\omega$ periodicity
3464: of the Floquet spectrum, as a consequence of eq.~(\ref{ebkpt2}).
3465: 
3466: As already noted in section~\ref{scaling_laws}, the semiclassical quantization
3467: breaks the scaling of the classical dynamics. Nontheless, 
3468: the semiclassical
3469: energy levels can be written in terms of the
3470: scaled parameters introduced above, by virtue of
3471: eqs.~(\ref{scaling_F},\ref{posres}):
3472: \begin{equation}
3473: {\cal E}_{N,k=0} = \frac{1}{n_0^2}
3474: \left[ -\frac{3}{2} + J_1'(1) F_0 - \frac{N+1/2}{n_0} \sqrt{3J_1'(1)F_0}
3475: \right].
3476: \label{effective}
3477: \end{equation}
3478: Note that the term $(N+1/2)/n_0$ highlights 
3479: the role of $1/n_0$ as an effective Planck constant.
3480: 
3481: As discussed in section~\ref{section_mathieu},
3482: the fully ``quantum" quasienergies of the
3483: resonantly driven atom can be obtained using the
3484: very same pendulum approximation of the system, 
3485: together with the solutions
3486: of the Mathieu equation. In our case, the  characteristic exponent
3487: in the Mathieu equation is given by eq.~(\ref{cexp})
3488: and the Mathieu parameter is, according to 
3489: eqs.~(\ref{map_mathieu3}),(\ref{v1_LIN1D}),(\ref{h0second}):
3490: \begin{equation}
3491: q=\frac{4}{3}Fn_0^6J'_1(1)=\frac{4}{3}F_0n_0^2J'_1(1).
3492: \label{mathind_LIN1D}
3493: \end{equation}
3494: The quantum quasienergy levels are then given by eq.~(\ref{prediction_mathieu})
3495: which reads:
3496: \begin{equation}
3497: {\cal E}_{\kappa,k=0} = - \frac{3}{2n_0^2} 
3498: - \frac{3}{8n_0^4} \ a_{\kappa}(\nu,q).
3499: \label{prediction_mathieu_lin1d}
3500: \end{equation}
3501: 
3502: As discussed in section~\ref{QD}, the non-dispersive wave-packet
3503: with maximum localization is associated with $\kappa=0$ and is well
3504: localized inside the non-linear resonance island between the internal
3505: coulombic motion and the external driving, provided the parameter $q$
3506: is of the order of unity (below this value, the resonance island
3507: is too small to support a localized state). The minimum scaled 
3508: microwave amplitude is thus of the order of:
3509: \begin{equation}
3510: F_{0,\mathrm{trapping}}= F_{\mathrm{trapping}}n_0^4 \simeq \frac{1}{n_0^2}
3511: \label{ftrapping}
3512: \end{equation}
3513: which is thus {\em much smaller} -- by a factor $n_0^2$, i.e.
3514: three orders of magnitude in typical experiments --  
3515: than the electric field created
3516: by the nucleus. This illustrates that a well chosen weak perturbation
3517: may strongly influence the dynamics of a non-linear system. From the
3518: experimental point of view, this is good news, a limited
3519: microwave power is sufficient to create non-dispersive wave-packets.
3520: 
3521: We have
3522: so far given a complete description of the dynamics of the
3523: resonantly driven, one-dimensional
3524: Rydberg electron, from a semiclassical as well as from a quantum mechanical
3525: point of view, in the resonant approximation. These approximate treatments
3526: are now complemented by a numerical solution of the exact quantum mechanical
3527: eigenvalue problem described by the Floquet equation (\ref{calhq}),
3528: with $H_0$ and $V$ from eqs.~(\ref{h0_1d}),(\ref{v_1d}),
3529: as well as by the numerical integration of the classical equations of motion
3530: derived from eq.~(\ref{ham_lin1d}). Using this machinery, we 
3531: illustrate some of the essential properties of nondispersive wave-packets
3532: associated
3533: with the principal resonance in this system, whereas we 
3534: postpone the discussion of other primary resonances
3535: to section \ref{ORH}.
3536: 
3537: Fig.~\ref{lin1d_00} compares the phase space structure of the exact classical
3538: dynamics generated by the Hamilton function (\ref{ham_lin1d}), and the isovalue
3539: curves of the pendulum dynamics, eq.~(\ref{hpend_lin1d}),
3540: for the case $n_0=60,$
3541: at scaled field strength
3542: $F_0=Fn_0^4=0.01$. The Poincar\'e surface of section is taken
3543: at 
3544: phases $\omega t=0$ (mod $2\pi$) and plotted in
3545: $(I,\theta)$ variables which, for such times, coincide
3546: with the $(\hat{I},\hat{\theta})$ variables, 
3547: see eqs.~(\ref{rotframe_a}),(\ref{rotframe_b}).
3548: Clearly, the pendulum approximation predicts the
3549: structure of the invariant curves very well, 
3550: with the resonance island surrounding
3551: the stable periodic orbit at $(\hat{I}\approx 60,\hat{\theta}=\pi),$
3552: the unstable fixed point at $(\hat{I}\approx 60,\hat{\theta}=0),$
3553: the separatrix, and the rotational motion outside the resonance island.
3554: Apparently, only tiny regions of stochastic motion invade the
3555: classical phase space, which hardly affects the quality of the pendulum
3556: approximation. It should be emphasized that -- because of the scaling laws,
3557: see section~\ref{scaling_laws} -- the figure depends on
3558: the scaled field strength $F_0$ only. Choosing a different microwave
3559: frequency with the same scaled field leads, via eq.~(\ref{acteqqn}), 
3560: to a change
3561: of $n_0$ and, hence, of the scale of $I.$
3562: \begin{figure}
3563: \centerline{\psfig{figure=bdzf10.ps,width=15cm,angle=-90}}
3564: \caption{Poincar\'e surface of section for the dynamics of a 1D hydrogen
3565: atom driven by an external oscillatory electric field, see eq.~(\protect\ref{ham_lin1d}).
3566: The driving frequency is chosen as $\omega=1/60^3,$ such that the nonlinear
3567: resonance island is centered at principal action (or effective
3568: principal quantum number) $n_0=60.$ The scaled external field amplitude
3569: is set to $F_0=Fn_0^4=0.01$. Although the driving field is much weaker
3570: than the Coulomb field between the electron and the nucleus, 
3571: it suffices to create a relatively
3572: large resonance island which supports several non-dispersive
3573: wave-packets.}
3574: \label{lin1d_00}
3575: \end{figure}
3576: 
3577: Fig.~\ref{lin1d_0} compares the prediction of the Mathieu approach,
3578: eq.~(\ref{prediction_mathieu_lin1d}), for the quasienergy levels of the
3579: Floquet Hamiltonian to
3580: the exact numerical result obtained by diagonalization of the full Hamiltonian,
3581: see section~\ref{RSEF}. Because of the $\hbar\omega$ periodicity
3582: of the Floquet spectrum, the sets of energy levels of the pendulum,
3583: see fig.~\ref{fig_mathieu},
3584: are folded in one single Floquet zone.
3585: For  states located inside or in the vicinity
3586: of the resonance island -- the only ones plotted in 
3587: fig.~\ref{fig_mathieu}a --
3588: the agreement is very good for low and
3589: moderate
3590: field strengths.
3591: Stronger
3592: electromagnetic fields lead to deviations between the Mathieu and the exact
3593: result. This indicates 
3594: higer order corrections to the pendulum approximation.
3595: \begin{figure}
3596: \centerline{\psfig{figure=bdzf11.ps,width=15cm,angle=-90}}
3597: \caption{Comparison of the exact quasienergy spectrum of the 1D hydrogen atom
3598: driven by a linearly polarized microwave field (right),
3599: eq.~(\protect\ref{ham_lin1d}),
3600: with the prediction of the pendulum approximation (left),
3601: eq.~(\protect\ref{prediction_mathieu_lin1d}). Because of the $\hbar\omega$
3602: periodicity of the Floquet spectrum, the energy levels
3603: described by the pendulum approximation are folded inside one
3604: Floquet zone. The agreement between the exact quantum result and the
3605: pendulum approximation is very good. 
3606: The filled circle shows the 
3607: most localized non-dispersive wave-packet $N=0$ shown in
3608: figs.~\protect\ref{lin1d_2}-\protect\ref{lin1d_4}, while
3609: the filled square represents the hyperbolic non-dispersive wave-packet
3610: partly localized in the vicinity of the unstable
3611: equilibrium point of the pendulum, shown in
3612: figs.~ \protect\ref{lin1d_5} and \protect\ref{lin1d_7}. 
3613: The open circle and square compare the exact location of the
3614: respective quasienergy values with the Mathieu prediction, which is
3615: considerably better for the ground state as compared to the separatrix 
3616: state.}
3617: \label{lin1d_0}
3618: \end{figure}
3619: 
3620: Fig.~\ref{lin1d_1} shows a typical Poincar\'e surface of section of the
3621: classical dynamics of the driven Rydberg electron, at different values
3622: of the phase of the driving field. The field amplitude is chosen
3623: sufficiently high to
3624: induce largely chaotic dynamics,
3625: with the
3626: principal resonance as the only remnant of regular motion occupying
3627: an appreciable volume of phase space.
3628: The figure clearly illustrates
3629: the temporal evolution of the elliptic island with the phase of the driving
3630: field, i.e. the locking of the electronic motion on the external driving. The
3631: classical stability island follows the dynamics of the unperturbed electron
3632: along the 
3633: resonant trajectory.
3634: The distance from the nucleus is parametrized by the variable $\theta,$ 
3635: see eq.~(\ref{aa_lin1d}). At $\omega t = \pi,$ the 
3636: classical electron hits 
3637: the nucleus (at $\theta=0$),
3638: its velocity
3639: diverges and changes sign discontinuously.
3640: This explains the
3641: distortion of the resonance island as it approaches $\theta=0,\pi$.
3642: \begin{figure}
3643: \centerline{\psfig{figure=bdzf12.eps,width=15cm,angle=-90}}
3644: \caption{Surface of section of the classical phase space of a 1D hydrogen atom,
3645: driven by a linearly polarized microwave field of amplitude $F_0=0.053$, 
3646: for different values of
3647: the phase: $\omega t=0$ (top left), $\omega t=\pi/2$ (top right), $\omega t=\pi$ (bottom left),
3648: $\omega t=3\pi/2$ (bottom right). At this driving strength,
3649: the principal
3650: resonance remains as the
3651: only region of regular motion of appreciable size,
3652: in a globally chaotic phase space. The action-angle variables 
3653: $I,\theta$ are defined by
3654: eq.~(\protect\ref{aa_lin1d}), according to which a 
3655: collision with
3656: the nucleus occurs 
3657: at $\theta=0$.}
3658: \label{lin1d_1}
3659: \end{figure}
3660: 
3661: Quantum mechanically, we expect a non-dispersive wave-packet eigenstate 
3662: to be localized within the resonance island. 
3663: The semiclassical prediction 
3664: of its quasienergy, eq.~(\ref{prediction_mathieu_lin1d}),
3665: facilitates to identify the nondispersive wave-packet within  the 
3666: exact Floquet spectrum, after numerical diagonalization of the Floquet 
3667: Hamiltonian (\ref{calhq}).
3668: The wave-packet's configuration space
3669: representation is shown 
3670: in fig.~\ref{lin1d_2}, for the same phases of the field as
3671: in the plots of the 
3672: classical dynamics in fig.~\ref{lin1d_1}. Clearly, the wave-packet
3673: is very well localized at the outer turning point of the Kepler electron
3674: at phase $\omega t=0$
3675: of the driving field, and is reflected off the nucleus half a period of the
3676: driving field later. 
3677: On reflection, the electronic density exhibits some
3678: interference structure, as well as some transient spreading. This is a
3679: signature of the quantum mechanical uncertainty in the angle $\theta$:
3680: part of the wave-function, which still
3681: approaches
3682: the Coulomb singularity, interferes with 
3683: the other part already reflected off the nucleus.  
3684: The transient spreading is equally 
3685: manifest in the temporal
3686: evolution of the uncertainty product $\Delta z\Delta p$ itself, which is
3687: plotted in fig.~\ref{lin1d_3}.
3688: Apart from this singularity at $\omega t=\pi,$ the wave-packet
3689: is approximately Gaussian at any time, with a time-dependent 
3690: width
3691: (compare $\omega t=0$ and
3692: $\omega t=\pi/2$).
3693: \begin{figure}
3694: \centerline{\psfig{figure=bdzf13.ps,width=14cm,angle=-90}}
3695: \caption{Configuration space representation of the electronic 
3696: density of the non-dispersive wave-packet eigenstate
3697: of a 1D hydrogen atom,
3698: driven by a linearly polarized microwave field, for the same field amplitude
3699: and phases as in fig.~\protect\ref{lin1d_1}. 
3700: The eigenstate is 
3701: centered on the principal resonance of classical phase space,
3702: at action (principal quantum number) $n_0=60$.
3703: In configuration space,
3704: the wave-packet is localized at the outer
3705: turning point (with zero average velocity) at time $t=0$, then propagates
3706: towards the nucleus which it hits at $t/T=0.5$ (where $T=2\pi/\omega$ is the
3707: microwave period). 
3708: Afterwards, it propagates outward to the apocenter
3709: which is reached at time $t/T=1$. After one period, the
3710: wave-packet recovers {\em exactly} its initial shape, and
3711: will therefore propagate along the classical trajectory 
3712: forever, without spreading.
3713: The wave-packet has approximately Gaussian shape(with time-dependent width) 
3714: except at $t/T\simeq 0.5$.
3715: At this instant, the head of the wave-packet, which 
3716: already has been reflected off the nucleus, interferes with its tail,
3717: producing interference fringes. 
3718: }
3719: \label{lin1d_2}
3720: \end{figure}
3721: \begin{figure}
3722: \centerline{
3723: \psfig{figure=bdzf14a.eps,width=7cm,angle=-90}
3724: \psfig{figure=bdzf14b.eps,width=7cm,angle=-90}}
3725: \caption{(a) Dashed line: 
3726: Classical temporal evolution of the position of the Rydberg electron 
3727: resonantly driven by a linearly polarized microwave field, in the 
3728: one dimensional model, eq.~(\protect\ref{ham_lin1d}), of the hydrogen atom.
3729: ($n_0=60$; scaled field amplitude $F_0=Fn_0^4=0.053)$. 
3730: Thick line: Expectation value $\langle z\rangle$ for the non-dispersive
3731: wave-packet shown in fig.~\protect\ref{lin1d_2}, as a function of time.
3732: It follows the classical trajectory remarkably well (except for collisions
3733: with the nucleus).
3734: Dotted line: The position uncertainty $\Delta z = \sqrt{\langle z^2\rangle -
3735: \langle z \rangle ^2}$ 
3736: of the wave-packet. 
3737: $\Delta z$ being much smaller than
3738: $\langle z \rangle$ (except near collisions with the nucleus)
3739: highlights 
3740: the efficient localization of the wave-packet.
3741: (b) Uncertainty product $\Delta z\Delta p$ of the 
3742: wave-packet. The periodically repeating maxima 
3743: of this quantity indicate the collision of the electron 
3744: with the atomic nucleus.
3745: Note that the minimum uncertainty at
3746: the outer turning point of the wave-packet is very close to the Heisenberg
3747: limit $\hbar/2$. Although the wave-packet is {\em never}
3748: a minimal one, it is nevertheless well localized and
3749: an excellent 
3750: approximation of a classical particle.}
3751: \label{lin1d_3}
3752: \end{figure}
3753: 
3754: To complete the analogy between classical and quantum motion, we finally
3755: calculate the Husimi distribution -- the
3756: phase space representation of the wave-packet eigenstate
3757: defined in section~\ref{coherent_states}, eq.~(\ref{husimi_def})  --
3758: in order to obtain a direct comparison between classical and
3759: quantum dynamics in phase space. Fig.~\ref{lin1d_4} shows the resulting 
3760: phase
3761: space 
3762: picture, 
3763: again
3764: for different phases of the driving field. The association of the quantum
3765: mechanical time evolution with the classical resonance island (see fig.~\ref{lin1d_1}) 
3766: is
3767: unambiguous. The transient spreading at the collision with 
3768: the nucleus ($\omega t=\pi$) is due to the divergence
3769: of the classical velocity 
3770: upon reflection.
3771: \begin{figure}
3772: \centerline{\psfig{figure=bdzf15an.ps,width=7cm}
3773: \psfig{figure=bdzf15bn.ps,width=7cm}}
3774: \smallskip
3775: \centerline{\psfig{figure=bdzf15cn.ps,width=7cm}
3776: \psfig{figure=bdzf15dn.ps,width=7cm}}
3777: \caption{Husimi representation of the wave-packet eigenstate of
3778: fig.~\protect\ref{lin1d_2} in classical phase space, for the same 
3779: phases $\omega t$ and 
3780: scales $(0\leq \theta \leq 2\pi, 30\leq I \leq 90)$ as employed for the
3781: classical surface of section in fig.~\protect\ref{lin1d_1}. Clearly, the
3782: quantum mechanical eigenstate of the atom in the field follows the classical
3783: evolution without dispersion, except for its transient spreading when 
3784: reflected off 
3785: the nucleus (at $\omega t=\pi$, bottom left), due to the divergence
3786: of the classical velocity at that position.
3787: }
3788: \label{lin1d_4}
3789: \end{figure}
3790: 
3791: As discussed in section~\ref{CD} and visible in fig.~\ref{lin1d_00},
3792: there is an hyperbolic fixed point (i.e. an unstable equilibrium point)
3793: at $(\hat{I}=n_0,\hat{\theta}=0)$: it corresponds to the
3794: unstable equilibrium position of the pendulum 
3795:  when it points ``upwards". For the driven system,
3796: it corresponds to an unstable periodic orbit resonant with the
3797: driving frequency: it is somewhat similar to the stable
3798: orbit supporting the non-dispersive wave-packets, except that is is
3799: shifted in time by half a period. 
3800: 
3801: As discussed in section~\ref{scars}, the classical motion slows down 
3802: at the hyperbolic fixed point
3803: (the time to reach the unstable equilibrium point with zero velocity
3804: diverges~\cite{marion}), and the eigenfunction must 
3805: exhibit a maximum of the electronic density at this
3806: position. 
3807: In addition, due to the periodicity
3808: of the drive, the corresponding (``hyperbolic'') wave-packet eigenstate
3809: necessarily follows the dynamics of a classical particle 
3810: which evolves along the unstable periodic orbit.
3811: However, because the orbit is unstable, the quantum 
3812: eigenstate cannot remain 
3813: fully localized -- some probability
3814: has to flow away along 
3815: the unstable manifold of the classical flow in the vicinity of the 
3816: hyperbolic fixed point.
3817: Consequently,
3818: such an eigenstate is partially localized along 
3819: the separatrix between librational
3820: and rotational motion.
3821: For an illustration, first consider fig.~\ref{lin1d_4a}, which 
3822: shows classical surfaces
3823: of section of the driven (1D) hydrogen atom, at $F_0=0.034$, 
3824: again for different phases $\omega t$.
3825: Comparison with fig.~\ref{lin1d_1} shows a larger 
3826: elliptic island at this slightly lower field amplitude, 
3827: as well as remnants of the $s=2$ resonance island 
3828: at slightly larger
3829: actions $I/n_0\simeq 1.2\ldots 1.3$. 
3830: The time evolution of the electronic density of the
3831: eigenstate localized near the
3832: hyperbolic fixed point is displayed in fig.~\ref{lin1d_5},
3833: for different phases of the driving field.
3834: Clearly, as compared to fig.~\ref{lin1d_2}, the wave-packet moves in phase
3835: opposition to the driving field, and displays slightly irregular localization
3836: properties.  Accordingly, the Husimi representation in
3837: fig.~\ref{lin1d_7}
3838: exhibits reasonnably good localization on top of the hyperbolic point at phase
3839: $\omega t=\pi$, but the electronic probability spreads significantly
3840: along the
3841: separatrix layer at phase, as visible at $\omega t=0$.
3842: \begin{figure}
3843: \centerline{\psfig{figure=bdzf16.eps,width=14cm,angle=-90}}
3844: \caption{Surface of section of the classical phase space of a 1D hydrogen atom
3845: driven by a linearly polarized microwave field, for different values of
3846: the phase: $\omega t=0$ (top left), $\omega t=\pi/2$ (top right), $\omega t=\pi$ (bottom left),
3847: $\omega t=3\pi/2$ (bottom right). The action angle variables $I,\theta$ are defined by
3848: eq.~(\protect\ref{aa_lin1d}). At this value of the field amplitude,
3849: $F_0=0.034$, the principal resonance island 
3850: (and a small remnant of the $s=2$ resonance island) remain as the
3851: only regions of regular motion, in a globally chaotic phase space.}
3852: \label{lin1d_4a}
3853: \end{figure}
3854: \begin{figure}
3855: \centerline{\psfig{figure=bdzf17n.ps,width=14cm,angle=-90}}
3856: \caption{Wave-packet eigenstate anchored to the hyperbolic fixed point of the
3857: principal resonance of the 1D hydrogen atom driven by a linearly polarized 
3858: microwave field,
3859: in configuration space,
3860: for the same phases
3861: of the driving field as in fig.~\protect\ref{lin1d_4a}. 
3862: The wave-function is partly localized, especially close to
3863: the outer turning point at $t=0.5\times 2\pi/\omega$, but the localization
3864: is far from being perfect.
3865: Comparison to figs.~\protect\ref{lin1d_1}, \protect\ref{lin1d_2} and
3866: \protect\ref{lin1d_4a} shows
3867: that the state evolves in phase opposition with the stable, non-dispersive
3868: wave-packet, with significantly worse localization properties.}
3869: \label{lin1d_5}
3870: \end{figure}
3871: \begin{figure}
3872: \centerline{\psfig{figure=bdzf18a.ps,width=7cm}
3873: \psfig{figure=bdzf18b.ps,width=7cm}}
3874: \smallskip
3875: \centerline{\psfig{figure=bdzf18c.ps,width=7cm}
3876: \psfig{figure=bdzf18d.ps,width=7cm}}
3877: \caption{Husimi
3878: representation of the ``hyperbolic" wave-packet eigenstate of
3879: fig.~\protect\ref{lin1d_5}, for the same phases 
3880: (scales as in fig.~\protect\ref{lin1d_4a}).
3881: Clearly, the
3882: quantum mechanical eigenstate of the atom in the field follows the classical
3883: evolution. It is partly
3884: localized on top of the hyperbolic fixed point, but also
3885: spreads along the separatrix confining the principal
3886: resonance.
3887: The localization is more visible at $t=0$ (top left),
3888: the spreading more visible  at
3889: $t=0.5 \times 2\pi/\omega$ (bottom left).}
3890: \label{lin1d_7}
3891: \end{figure}
3892: 
3893: In the above discussion of the localization properties of the wave-packet 
3894: eigenstate we represented the wave-function in the $I$-$\theta$ phase space
3895: of {\em classically bounded} motion (i.e., classical motion with negative 
3896: energy). However, as we shall see in more detail in section~\ref{ION}, the 
3897: microwave driving actually induces a nonvanishing overlap of {\em all} 
3898: Floquet eigenstates \cite{graffi85,yajima82}, and, hence, of the wave-packet
3899: eigenstates, with the atomic continuum. It suffices to say here that 
3900: the associated finite decay rates induce finite life times of approx. 
3901: $10^6$ unperturbed Kepler orbits for the quantum objects considered in this 
3902: chapter, and are therefore irrelevant on the present level of our discussion.
3903: In figs.~\ref{lin1d_2}, \ref{lin1d_4}, \ref{lin1d_5}, and \ref{lin1d_7}, a
3904: finite decay rate would manifest as a slow reduction of the electronic 
3905: density, without affecting its shape or localization properties, after 
3906: $10^6$ classical Kepler periods.
3907: 
3908: \subsubsection{Realistic three-dimensional atom}
3909: \label{LIN3D}
3910: 
3911: Extending our previous analysis to the three-dimensional hydrogen atom driven
3912: by a linearly polarized microwave field, we 
3913: essentially expand the accessible 
3914: phase space. Since the Hamiltonian 
3915: \begin{equation}
3916: H_{\rm LP}=\frac{{\vec p}^2}{2}-\frac{1}{r}+Fz\cos(\omega t)
3917: \label{ham_lin_3d}
3918: \end{equation}
3919: is invariant under rotations around the field polarization axis,  
3920: the projection of the angular momentum is a conserved quantity and gives
3921: rise to a good quantum number $M$. 
3922: Hence, only two dimensions of configuration space are left, which, together
3923: with the explicit, periodic time dependence, span a five-dimensional 
3924: phase space. 
3925: 
3926: In the 1D situation described previously, the key ingredient for the existence
3927: of non-dispersive wave-packets was the phase locking of the internal
3928: degree of freedom on the external drive. In the 3D situation, there remains
3929: one single drive, but there are 
3930: several internal degrees of freedom. In the generic
3931: case, not all internal degrees of freedom can be simultaneously locked
3932: on the external drive, and one can expect only partial
3933: phase locking, i.e. only partially localized wave-packets. The non-trivial
3934: task is to understand how the phase locking of one degree of freedom
3935: modifies the dynamics along the other degrees of freedom. In atomic systems,
3936: the Coulomb degeneracy makes it possible to gain a full understanding
3937: of this phenomenon. 
3938: 
3939: The starting point is similar to the 1D analysis in
3940: sec.~\ref{LIN1D}, that is the expression of the Floquet
3941: Hamiltonian  -- whose eigenstates are of interest --
3942: as a function of action-angle coordinates
3943: $(I,\theta)$, $(L,\psi)$,
3944: $(M,\phi)$ introduced in section \ref{RSEF}.
3945: Using eqs.~(\ref{beta},\ref{rot_euler})
3946: and the Fourier expansion, eq.~(\ref{v_3d_x}-\ref{v_3d_z}),
3947: of the position operator, one obtains:
3948: \begin{equation}
3949: {\mathcal H} = P_t -\frac{1}{2I^2}+ F \sqrt{1-\frac{M^2}{L^2}} 
3950: \sum_{m=-\infty}^{+\infty}{\left[-X_m\cos\psi  \cos(m\theta-\omega t)
3951: +  Y_m \sin\psi\sin(m\theta-\omega t)\right]},
3952: \label{hamaa}
3953: \end{equation}
3954: with
3955: \begin{eqnarray}
3956: X_m(I)&=& I^2\frac{J_m'(me)}{m},\ \ \ \ m\neq 0, \label{xm} \\
3957: Y_m(I)&=& I^2 \frac{\sqrt{1-e^2}J_m(me)}{me},\ \ \ \ m\neq 0,  \label{ym}\\ 
3958: X_0(I)&=& -\frac{3}{2}eI^2, \label{x0}\\
3959: Y_0(I)&=&0.
3960: \label{y0}
3961: \end{eqnarray}
3962: where $e=\sqrt{1-L^2/I^2}$ is, as before, the eccentricity of the Kepler orbit
3963: (see 
3964: eq.~(\ref{eccentricity})).
3965: The absence of $\varphi$ in the 
3966: Hamiltonian reflects the azimuthal symmetry around the field axis and 
3967: ensures the conservation of $M$.
3968: 
3969: Precisely as in the treatment of the 
3970: one-dimensional 
3971: problem, we now transform 
3972: to slowly varying variables, given by 
3973: eqs.~(\ref{rotframe_a})-(\ref{rotframe_c}): 
3974: \begin{eqnarray}
3975: \hat{\mathcal H} & = & \hat{P}_t -\frac{1}{2 \hat I^2}- \omega\hat I +
3976: F \sqrt{1-\frac{M^2}{L^2}} \nonumber \\
3977: \times & & \sum_{m=-\infty}^{+\infty} {\left[-X_m \cos\psi  
3978: \cos(m\hat{\theta} + (m-1)\omega t)
3979: + Y_m \sin\psi \sin(m\hat{\theta} + (m-1)\omega t)\right]}.
3980: \label{hamab}
3981: \end{eqnarray}
3982: Averaging over the fast variable $t$ (over the driving field period 
3983: $T$) gives
3984: the secular Hamiltonian of the three-dimensional problem
3985: \begin{equation}
3986: {\mathcal H}_{\rm sec}  = \hat{P}_t -\frac{1}{2\hat I^2}- 
3987: \omega\hat{I}
3988:  +F \sqrt{1-\frac{M^2}{L^2}}(-X_1(\hat{I})\cos\psi \cos\hat\theta+
3989: Y_1(\hat{I})\sin\psi \sin\hat\theta).
3990: \label{hamsc2}
3991: \end{equation}
3992: Its physical interpretation is rather simple: the $X_1$ term represents
3993: the oscillating dipole (resonant with the frequency of the drive) 
3994: along the major
3995: axis of the classical Kepler ellipse, while the $Y_1$ term represents
3996: the oscillating dipole along the minor axis. As these 
3997: two components of the oscillating dipole are in quadrature, they interact
3998: with two orthogonal components of the external drive, hence
3999: the $\cos\hat\theta$ and $\sin\hat\theta$ terms.
4000: Finally, both components can be combined to produce the compact form
4001: \begin{equation}
4002: {\mathcal H}_{\rm sec}=\hat{P}_t-\frac{1}{2\hat I^2}-\omega \hat{I}
4003: +F \chi_1 \cos(\hat \theta + \delta_1),
4004: \label{hamscfin}
4005: \end{equation}
4006: with
4007: \begin{eqnarray}
4008: \chi_1(\hat I,L,\psi) & := & \sqrt{1-\frac{M^2}{L^2}}\ \sqrt{X_1^2\cos^2\psi+Y_1^2\sin^2\psi} ,
4009: \label{substc} \\
4010: \tan\delta_1 (L,\psi)& := & \frac{Y_1}{X_1}\tan\psi = 
4011: \frac{J_1(e)\sqrt{1-e^2}}{J'_1(e)e} \tan\psi.
4012: \label{substb}
4013: \end{eqnarray}
4014: In this form, the secular Hamiltonian has the same structure as
4015: the general 1D expression, eq.~(\ref{hsec_ap}), and its specialized
4016: version for the 1D hydrogen atom, eq.~(\ref{hsec_h1d}).
4017: The difference is that the additional action angle-variables
4018: $(L,\psi)$, $(M,\phi)$ {\em only} enter in the amplitude and phase of 
4019: the coupling defining the resonance island. 
4020: This allows to separate various time scales 
4021: in the system:
4022: \begin{itemize}
4023: \item The shortest time scale is associated with the Kepler
4024: motion, which is also the period of the external drive.
4025: In the resonant approximation discussed in detail
4026: in section~\ref{CD}, this time scale is eliminated by passing to the
4027: rotating frame. 
4028: \item The time scale of the secular (or pendulum) motion
4029: in the $(\hat{I},\hat{\theta})$ plane is significantly longer.
4030: It is the inverse of the classical pendulum frequency,
4031: eq.~(\ref{omega_harmonic}), of the order of
4032: $1/\sqrt{F_0}$ Kepler periods. In the regime of weak external
4033: driving we are interested in, $F_0 \ll 1$, it is thus much longer
4034: than the preceding time scale.
4035: \item The time scale of the ``transverse" (or angular) motion along the
4036: $(L,\psi)$, $(M,\phi)$ variables. Because these 
4037: are constant for the unperturbed Coulomb system, the time
4038: derivatives like $dL/dt$ and 
4039: $d\psi/dt$ generated by eqs.~(\ref{hamscfin})-(\ref{substb}) 
4040: are proportional 
4041: to $F$, and the resulting time scale is proportional
4042: to $1/F$. More precisely, it is of the order
4043: of $1/F_0$ Kepler periods, i.e., once again, significantly longer
4044: than the preceding time scale.
4045: \end{itemize}
4046: 
4047: From this separation of time scales, it 
4048: follows that we can use the following, additional secular approximation:
4049: for the motion in the $(\hat{I},\hat{\theta})$ plane, 
4050: $\chi_1$ and $\delta_1$ are adiabatic invariants,
4051: which can be considered as constant quantities. We then exactly
4052: recover the Hamiltonian discussed for the 1D model of the atom,
4053: with 
4054: a resonance island confining trajectories
4055: with librational motion in the $(\hat{I},\hat{\theta})$ plane, and
4056: rotational motion outside the resonance island.
4057: The center of the 
4058: island is located at:
4059: \begin{equation}
4060: \hat{I}=\hat{I}_1 = \omega^{-1/3}=n_0,\ \ \ \ \ \ 
4061: \hat{\theta} = -\delta_1.
4062: \label{Kures}
4063: \end{equation}
4064: As already pointed out in section \ref{CD}, 
4065: the 
4066: size of the resonance island in phase space is determined
4067: by the strength of the resonant coupling $\chi_1(\hat I,L,\psi).$
4068: In the pendulum approximation, its extension in $\hat{I}$ 
4069: scales as $\sqrt{\chi_1(\hat{I}_1,L,\psi)}$, i.e. with
4070: $\chi_1$ evaluated at the center, eq.~(\ref{Kures}), of the island.
4071: 
4072: The last step is to consider the slow motion in the $(L,\psi)$ plane.
4073: As usual, when a secular
4074: approximation is employed, the slow motion
4075: is due to an effective Hamiltonian which is 
4076: obtained by averaging of the secular Hamiltonian over the fast motion.
4077: Because the coupling $\chi_1(\hat{I},L,\psi)$ exhibits a simple scaling 
4078: with $\hat{I}$ (apart from a global $\hat{I}^2$ dependence, it depends
4079: on the scaled angular variables $L_0$ and $\psi$ only),
4080: the averaging over the fast motion results in an effective
4081: Hamiltonian for the ($L_0$,$\psi$) motion which depends
4082: on $\chi_1(\hat{I},L,\psi)$ only.  
4083: We deduce
4084: that the slow motion 
4085: follows curves of constant
4086:  $\chi_1(\hat{I}_1,L,\psi),$ at a velocity which
4087: depends on the average over the fast variables.
4088: $\chi_1(\hat{I}_1,L,\psi)$ is thus a constant of motion, both for the fast
4089: $(\hat{I},\hat{\theta})$  and the slow $(L,\psi)$ motion.  
4090: This also implies that the order of the 
4091: quantizations in the fast and slow variables
4092: can be interchanged: using the dependence of $\chi_1(\hat{I}_1,L,\psi)$ 
4093: on $(L,\psi),$ we obtain quantized values of $\chi_1$
4094: which in turn can be used as constant values to quantize the
4095: $(\hat{I},\hat{\theta})$ fast motion. 
4096: Note that the separation of time scales is here essential
4097: \footnote{If one considers non-hydrogenic atoms -- with
4098: a core potential in addition to the Coulomb potential -- the classical
4099: unperturbed ellipse precesses, adding an additional time scale,
4100: and the separation of time scales is much less obvious. See also
4101: section~\protect{\ref{SPEC}}.}.
4102: Finally,
4103: the dynamics in $(M,\phi)$ is trivial, since $M$ is a constant
4104: of motion. In the following, we will consider 
4105: the case $M=0$ for simplity. 
4106: Note that, when the eccentricity of the classical
4107: ellipse tends to $1$ -- i.e. $L\to 0$ -- and when
4108: $\psi \to 0,$ the Hamiltonian (\ref{hamsc2}) coincides
4109: exactly with the Hamiltonian of the 1D atom, eq.~(\ref{hsec_h1d}).
4110: This is to be expected, as it corresponds to 
4111: a degenerate classical Kepler ellipse along the $z$ axis.
4112: 
4113: The adiabatic separation of the 
4114: radial and of the angular motion allows the separate WKB quantization of
4115: the various degrees of freedom. In addition to the quantization conditions
4116: in $(\hat{P}_t,t)$ and $(\hat{I},\hat{\theta})$, 
4117: eqs.~(\ref{ebkpt}-\ref{ebkinres}),
4118: already formulated in our general 
4119: description of the semiclassical approach in 
4120: section~\ref{sapp},
4121: we
4122: additionally need to quantize the angular motion, according to:
4123: \begin{equation}
4124: \frac{1}{2\pi}\oint_\gamma L d\psi =  \left(p+\frac{1}{2}\right)\hbar,
4125: \label{ebkp}
4126: \end{equation}
4127: along a loop $\gamma$ of constant
4128: $\chi_1$ in the $(L,\psi)$ plane.
4129: 
4130: Importantly, the loops of constant $\chi_1$ are independent
4131: of the microwave amplitude $F$ and scale simply with $\hat{I}.$
4132: Thus, the whole quantization in the 
4133: $(L,\psi)$ plane has to be done only once.
4134: With
4135: this prescription we can unravel the semiclassical 
4136: structure of the quasienergy spectrum induced by the additional 
4137: degree of freedom spanned by $(L,\psi)$, as an amendment to the spectral 
4138: structure of the one-dimensional model discussed in section~\ref{LIN1D}.
4139: Fig.~\ref{lin3d_1} shows the equipotential curves of $\chi_1$ in the
4140: $(L,\psi)$ plane.
4141: For a comparison with quantal data, the
4142: equipotential lines plotted correspond to the quantized
4143: values of $\chi_1$ for $n_0=21$.
4144: Using the well-known properties of the Bessel functions~\cite{abramowitz72},
4145: it is easy to show that $\chi_1(L,\psi)$ has the following
4146: fixed points:
4147: \begin{itemize}
4148: \item ($L=\hat{I}_1$, arbitrary $\psi$). This corresponds to a Kepler
4149: ellipse with maximum angular momentum, i.e. a circular orbit 
4150: in a plane containing the microwave polarization axis along $\hat z$. 
4151: As such a circle corresponds to a degenerate family of elliptical
4152: orbits with arbitrary orientation of the major axis, $\psi$ is a
4153: dummy angle. This fixed point corresponds to a global
4154: maximum of $\chi_1(L=\hat{I}_1)= \hat{I}_1^2/2$,
4155: and is surrounded by ``rotational"
4156: trajectories in the $(L,\psi)$ plane.
4157: An alternative representation of the ($L,\psi$) motion on the unit sphere,
4158: spanned by $L$ and the $z$ and $\rho$-components of the Runge-Lenz vector,
4159: contracts the line representing this orbit in 
4160: fig.~\ref{lin3d_1} 
4161: to an elliptic fixed point \cite{abu97}. 
4162: \item ($L=0, \psi=\pi/2,3\pi/2).$ This corresponds to a 
4163: degenerate straight line trajectory perpendicular to the
4164: microwave field. Because of the azimuthal
4165: symmetry around the electric field axis, the two points
4166: actually correspond to the same physics.
4167: The oscillating dipole clearly vanishes there,
4168: resulting in a global minimum of $\chi_1(L=0,\psi=\pi/2,3\pi/2)=0.$
4169: This stable fixed point is surrounded by ``librational" trajectories
4170: in the $(L,\psi)$ plane.
4171: \item ($L=0, \psi=0,\pi).$ This corresponds to a 
4172: degenerate, straight line trajectory along the
4173: microwave field, i.e. the situation already considered in the
4174: 1D model of the atom. $\psi=0$ and $\pi$ correspond
4175: to the two orbits pointing up and down, which are of course
4176: equivalent. This is a saddle point of
4177: $\chi_1(L=0,\psi=0,\pi)=J'_1(1)\hat{I}_1^2.$
4178: Hence, it is an unstable equilibrium point. As an implication,
4179: in the real 3D world, the motion along the microwave axis, with the
4180: phase of the radial motion locked on the external drive,
4181: is angularly unstable (see also section \ref{staticmw}). This leads 
4182: to a slow precession of the initially degenerate Kepler 
4183: ellipse off the axis, 
4184: and will manifest itself in the localization properties
4185: of the 3D analog of the nondispersive wave-packet displayed in 
4186: fig.~\ref{lin1d_2}. This motion takes place along
4187: the separatrix between librational and rotational
4188: motion.
4189: \end{itemize}
4190: \begin{figure}
4191: \centerline{\psfig{figure=bdzf19.ps,width=10cm,angle=-90}}
4192: \caption{Isovalue curves of the angular part $\chi_1$, eq.~(\ref{substc}),
4193: of the secular
4194: Hamiltonian ${\mathcal H}_{\rm sec}$ represented in the plane
4195: of the $L_0=L/\hat{I}_1$ and $\psi$ coordinates.
4196: The slow evolution of the
4197: Kepler ellipse of a Rydberg electron driven 
4198: by a resonant, linearly polarized microwave field, takes place
4199: along such isovalue curves.
4200: $L_0=L/\hat{I}$ represents the total angular
4201: momentum (a circular trajectory in a plane containing 
4202: the field polarization axis has $L_0=1$), and
4203: $\psi$ the [canonically conjugate] angle between
4204: the field polarization axis and the major axis of the Kepler ellipse. 
4205: The separatrix emanating from the unstable fixed point
4206: ($L_0=0,\psi=0$) separates rotational and librational motion, both 
4207: ``centered" around their respective stable
4208: fixed points ($L_0=1$, $\psi$ arbitrary) and ($L_0=0$, $\psi=\pi/2$). 
4209: The former corresponds to a circular orbit
4210: centered around the nucleus. The latter
4211: represents a straight linear orbit perpendicular to the field axis. The
4212: unstable fixed point corresponds to linear motion along the
4213: polarization axis. However, this initially degenerate Kepler ellipse will
4214: slowly precess in the azimuthal plane. The equipotential curves shown here
4215: satisfy the quantization condition (\protect\ref{ebkp}), 
4216: for $n_0=21$
4217: and $p=0\ldots 20$. At lowest order, the motion in the $(L_0,\psi)$ 
4218: plane is independent of 
4219: the microwave field strength and of the resonant principal quantum
4220: number $n_0$.
4221: }
4222: \label{lin3d_1} 
4223: \end{figure}
4224: Once the quantized values of $\chi_1$ (represented by 
4225: the trajectories in fig.~\ref{lin3d_1})
4226: have been determined,
4227: we can quantize the (${\hat I},{\hat \theta}$) 
4228: motion with 
4229: these values fixed. 
4230: Fig.~\ref{lin3d_2} shows the
4231: equipotential lines of ${\mathcal H}_{\rm sec}$, 
4232: for the three values of $\chi_1$
4233: corresponding to the $p=0$, $10$ and $20$ states, see eq.~(\ref{ebkp}), 
4234: of the 
4235: $n_0=21$ manifold. In each case, the contour for the lowest state
4236: $N=0$ has been drawn, together with the separatrix between the
4237: librational and rotational (${\hat I},\hat{\theta}$) modes. 
4238: The separatrix
4239: determines the size of the principal resonance island
4240: for the different substates of the transverse
4241: motion. Note that the principal resonance is largest for the 
4242: $p=20$ state, localized closest to the stable circular
4243: orbit (hence associated with the
4244: maximum value of $\chi_1$), whereas the smallest resonance
4245: island is obtained for the $p=0$ state, localized
4246: in the vicinity  
4247: of (though not precisely at) 
4248: the straight line orbit perpendicular to the field axis
4249: (minimum value of $\chi_1$).
4250: For the latter orbit itself, the first order coupling 
4251: vanishes identically ($\chi_1=0$), which 
4252: shows that the semiclassical results obtained 
4253: from our first order approximation (in $F$) for the Hamiltonian may be quite
4254: inaccurate in the vicinity of this orbit. Higher order corrections 
4255: may become important.
4256: \begin{figure}
4257: \centerline{\psfig{figure=bdzf20.eps,width=14cm,angle=-90}}
4258: \caption{
4259: Isovalue curves of the secular Hamiltonian ${\mathcal H}_{\rm sec}$,
4260: eq.~(\ref{hamscfin}),
4261: generating the ($\hat I,\hat \theta$)
4262: motion of a Rydberg electron in a resonant microwave field. 
4263: $\hat I$ and $\hat \theta$ correspond
4264: to the atomic principal quantum number, and to the polar angle of the electron on
4265: the Kepler ellipse, respectively. The scaled microwave amplitude is fixed at
4266: $F_0=0.03$. Since the isovalues of 
4267: ${\mathcal H}_{\rm sec}$ depend on the 
4268: transverse motion in $(L,\psi)$ via the constant value
4269: of $\chi_1$, eq.~(\ref{substc}), 
4270: contours (solid lines) are shown for three characteristic values of 
4271: $\chi_1$, corresponding to fixed quantum numbers $p=0,10,20$, 
4272: eq.~(\ref{ebkp}), 
4273: of the angular
4274: motion for the $n_0=21$ resonant manifold.
4275: Only the ``ground state'' orbit satisfying eq.~(\protect\ref{ebkires}) 
4276: with $N=0$ is 
4277: shown, 
4278: together with the separatrix (dashed lines) 
4279: between librational 
4280: and rotational motion in
4281: the ($\hat I,\hat \theta$) plane. The separatrix encloses the  
4282: principal
4283: resonance island in phase space,
4284: see also 
4285: eqs.~(\ref{eqpend},\ref{area}). Panel (a) corresponds to the orbit 
4286: with
4287: $L_0=L/n_0\simeq 1$ (rotational orbit, $p=20$), panel (b) to the orbit
4288: close to the separatrix of the angular motion ($p=10$), panel (c)
4289: to the librational orbit close to the stable fixed point $L_0=0$,
4290: $\psi=\pi/2$. Note that the resonance island  is smallest for librational, 
4291: largest
4292: for rotational, and of intermediate size for separatrix modes of the
4293: angular motion.}
4294: \label{lin3d_2}
4295: \end{figure}
4296: 
4297: As discussed above, the classical motion in the $(L,\psi)$ plane is 
4298: slower than
4299: in the (${\hat I},\hat{\theta}$) plane. In the semiclassical approximation,
4300: the spacing between consecutive states corresponds to the
4301: frequency of the classical motion (see also eq.~(\ref{cprinc})). 
4302: Hence, it is to be expected that states
4303: with the same quantum number $N$, but with successive quantum numbers $p$,
4304: will lie at neighboring energies, building well-separated manifolds associated
4305: with a single value of $N$. 
4306: The energy spacing 
4307: between
4308: states in the same manifold should scale as $F_0$, while the spacing between
4309: manifolds should scale as $\sqrt{F_0}$ (remember that $F_0\ll 1$ in the case
4310: considered here). Accurate quantum calculations fully
4311: confirm this prediction, with manifolds originating from
4312: the degenerate hydrogenic energy levels at $F_0=0$, as we shall demonstrate
4313: now. 
4314: We first concentrate on the $N=0$ manifold, originating
4315: from $n_0=21$. Fig.~\ref{lin3d_3} shows 
4316: the comparison between the
4317: semiclassical and the quantum energies, for different values of the 
4318: scaled driving
4319: field amplitude $F_0=Fn_0^4$. The agreement is excellent, except for 
4320: the lowest lying states 
4321: in the manifold for $F_0=0.02$. The lowest energy level ($p=0$)
4322: corresponds to motion close to the stable fixed point $L=0$, $\psi =\pi/2$ 
4323: in fig.~\ref{lin3d_1}; the highest energy level ($p=20$)
4324: corresponds to rotational motion $L/n_0\simeq 1$. The levels with
4325: the smallest energy difference ($p=10,11$) correspond to the librational and
4326: the rotational trajectories closest to the separatrix, respectively. 
4327: The narrowing
4328: of the level spacing in their vicinity is just a consequence of the slowing 
4329: down of the
4330: classical motion~\cite{delande97}. 
4331: In the same figure, we also plot (as a dashed line) the corresponding 
4332: exact
4333: quasienergy level for the 1D model of the atom (see section~\ref{LIN1D}).
4334: As expected, it closely follows the separatrix state $p=10$. 
4335: Such good agreement is a direct proof of the validity
4336: of the adiabatic separation between the
4337: slow motion in $(L,\psi)$, and the fast motion
4338: in (${\hat I},\hat{\theta}$).
4339: \begin{figure}
4340: \centerline{\psfig{figure=bdzf21.eps,width=15cm,angle=-90}}
4341: \caption{Comparison of the semiclassical quasienergies (circles; with
4342: $N=0$, see eq.~(\protect{\ref{ebkires}})),
4343: originating from the unperturbed $n_0=21$ manifold, 
4344: to the exact quantum
4345: result (crosses), at different values of the (scaled) driving field amplitude
4346: $F_0=Fn_0^4=0.02$ (a), $0.03$ (b), $0.04$ (c). The agreement is excellent. The
4347: quantum number $p=0\ldots 20$ labels the quantized classical
4348: trajectories plotted in fig.~\protect{\ref{lin3d_1}}, starting from the
4349: librational state $\mid p=0\rangle$ at lowest energy, rising through the
4350: separatrix states $\mid p=10\rangle$ and $\mid p=11\rangle$, up 
4351: to the rotational
4352: state $\mid p=20\rangle$. The dashed line indicates the exact quasienergy of
4353: the corresponding wave-packet eigenstate of the 1D model discussed in section
4354: \ref{LIN1D}. The 1D dynamics is neatly embedded in the spectrum of
4355: the real, driven 3D atom.}
4356: \label{lin3d_3}
4357: \end{figure}
4358: 
4359: Fig.~\ref{lin3d_4} shows a global comparison of the 
4360: semiclassical prediction with
4361: the exact level dynamics (energy levels vs. $F_0$), 
4362: in a range from $F_0=0$ to $F_0=0.06$, which exceeds the
4363: typical ionization threshold ($F_0\simeq 0.05$) observed in current 
4364: experiments
4365: \cite{bayfield89,koch95b,bellermann96}. We observe that the semiclassical 
4366: prediction 
4367: tracks the exact
4368: quasienergies quite accurately, even
4369: for large $F_0$-values, 
4370: where the resonant 
4371: $n_0=21$ manifold 
4372: overlaps with other Rydberg manifolds, or with side bands of 
4373: lower or higher lying
4374: Rydberg states.
4375: The agreement becomes unsatisfactory only 
4376: in the region of very small $F_0$, where the size
4377: of 
4378: the resonance island in $(\hat I, \hat \theta)$ is very small.
4379: This is not unexpected, as semiclassics should fail when the
4380: area of the resonance island is comparable to $\hbar$, compare 
4381: eq.~(\ref{number_of_trapped_states}).
4382: In this weak driving regime, the pendulum approximation can be used
4383: to produce more accurate estimates of the energy levels.
4384: The fast $(\hat{I},\hat{\theta})$ motion is essentially identical to the one
4385: of the 1D driven hydrogen atom: thus, the Mathieu approach used
4386: in section~\ref{LIN1D} can be trivially extended to the 3D case.
4387: The only amendment is to replace the factor $J_1'(1)n_0^2$ in
4388: the expression of the Mathieu parameter $q$ by the various
4389: quantized values of $\chi_1$ for $0\leq p \leq n_0-1,$ and to use
4390: the same equation (\ref{prediction_mathieu_lin1d}) for the energy levels.
4391: 
4392:  
4393: \begin{figure}
4394: \centerline{\psfig{figure=bdzf22.ps,width=10cm,angle=-90}}
4395: \caption{Level dynamics of the numerically 
4396: exact quasienergies (dotted lines)
4397: in the vicinity of the resonant manifold emerging from $n_0=21$ ($N=0$),
4398: compared to the semiclassical prediction (full lines), for 
4399: $F_0=Fn_0^4=0\ldots 0.06$.
4400: Note that the maximum field amplitude exceeds the typical ionization
4401: thresholds measured in current experiments for $\omega n_0^3 \simeq 1$
4402: \protect\cite{bayfield89,koch95b,bellermann96}. 
4403: Nontheless, the semiclassical prediction accurately 
4404: tracks the exact
4405: solution across a large number of avoided crossings
4406: with other Rydberg manifolds.
4407: }
4408: \label{lin3d_4}
4409: \end{figure}
4410: 
4411: The semiclassical construction of the energy levels from classical
4412: orbits is -- necessarily -- reflected in the localization properties of the
4413: associated eigenstates, as demonstrated by the electronic densities of the
4414: states $\mid p=0\rangle$, $\mid p=10\rangle$, and $\mid p=20\rangle$ in
4415: fig.~\ref{lin3d_5}, for the same field amplitudes as in 
4416: fig.~\ref{lin3d_3}. Note that, in this plot, the electronic densities 
4417: are averaged over one field cycle,
4418: hence display only the angular localization properties of the eigenstates.
4419: Their localization along the classical orbits defined by the
4420: stable or unstable fixed points of the
4421: $(L,\psi)$ dynamics is obvious \cite{abu96,abu98a,abu97}.
4422: Note in particular the nodal
4423: structure of the state $\mid p=10\rangle$, associated with the unstable 
4424: fixed point:
4425: there are sharp nodal lines perpendicular to the $z$-axis, reflecting the
4426: dominant motion along the $z$-axis, but also nodal lines of low visibility in
4427: the angular direction. They are a manifestation of the slow classical
4428: precession of the Kepler ellipse, i.e. the slow secular evolution in
4429: the ($L,\psi$) plane. 
4430: The quantum state, however, dominantly
4431: exhibits the motion along the $z$-axis, as a signature of the effective
4432: separation of time scales of radial and angular motion.
4433: Finally, it should be realized from a comparison of the top to the middle and
4434: bottom row of fig.~\ref{lin3d_5}
4435: that the quasiclassical localization properties of the eigenstates are
4436: essentially unaffected as $F$ rises, despite various avoided crossings which
4437: occur at intermediate field values, see fig.~\ref{lin3d_4}. Especially,
4438: the angular structure does not depend at all on $F$, as predicted by the
4439: secular approximation.
4440: \begin{figure}
4441: \centerline{\psfig{figure=bdzf23.ps,width=15cm}}
4442: \caption{Electronic densities of the extremal librational ($p=0$, left), 
4443: separatrix ($p=10$, center), and extremal rotational ($p=20$, right) 
4444: quasienergy states of the $n_0=21$  manifold of a 3D hydrogen
4445: atom exposed to a resonant microwave field
4446: in cylindrical coordinates $(\rho,z)$, at different values of the
4447: driving field amplitude $F_0=Fn_0^4=0.02$ (top), $0.03$ (middle), $0.04$ (bottom), 
4448: averaged
4449: over one period of the driving field. Note the clear localization along the
4450: classical orbits corresponding to the respective contours in
4451: fig.~\protect\ref{lin3d_1}, for {\em all} field amplitudes. The nodal lines
4452: of the electronic densities clearly exhibit the direction of the underlying
4453: classical motion. The field-induced finite decay rate of the eigenstates 
4454: (see section~\protect\ref{ION}) is negligible on time scales shorter than 
4455: approx. $10^6$ Kepler periods. Each box extends over 
4456: $\pm 1000$ Bohr radii, in both $\rho$ (horizontal) and
4457: $z$ (vertical) directions, with the nucleus at the center of the plot. The microwave polarization axis is oriented  
4458: vertically along $z.$}
4459: \label{lin3d_5}
4460: \end{figure}
4461: 
4462: The eigenstates displayed here are localized
4463: along classical trajectories which are resonantly driven by the external
4464: field. Hence, we should expect them to exhibit wave-packet like motion along
4465: these trajectories, as the phase of the driving field is changed. 
4466: This is indeed the case as illustrated  
4467: in fig.~\ref{lin3d_8} for the state $p=20$ 
4468: with maximal angular momentum 
4469: $L/n_0\simeq 1$ \cite{kalinski95b,abu96,abu98a}. 
4470: Due to the azimuthal symmetry of the problem, the actual 3D electronic
4471: density is obtained by rotating the figure around the vertical axis.
4472: Thus, the wave-packet is  actually a doughnut moving periodically
4473: from the north to the south
4474: pole (and back)
4475: of a sphere, slightly deformed along the field direction. The interference
4476: resulting from the contraction of this doughnut to a compact wave-packet at
4477: the poles is clearly visible at phases $\omega t=0$ and $\omega t=\pi$ in the
4478: plot. Note that the creation of unidirectional wave-packet
4479: eigenstates moving along a circle in the plane containing the field
4480: polarization axis is not possible for the real 3D atom \cite{abu98a},
4481:  as opposed to the
4482: reduced 2D problem studied in \cite{kalinski95b}, due to the 
4483: abovementioned azimuthal
4484: symmetry (see also section \ref{EP}).
4485: \begin{figure}
4486: \centerline{\psfig{figure=bdzf24.ps,width=14cm}}
4487: \caption{Temporal evolution of the electronic density
4488: of the extremal rotational quasienergy state
4489: $\mid p=20\rangle$ of the $n_0=21$ resonant manifold, 
4490: for different phases $\omega t=0$ (left),
4491: $\omega t=\pi/2$ (center), $\omega t=\pi$ (right) of the driving field, at
4492: amplitude $F_0=0.03$, in cylindrical coordinates. 
4493: Each box extends over $\pm 700$ Bohr radii, in both directions, $\rho$
4494: (horizontal) and
4495: $z$ (vertical). The microwave polarization axis is oriented 
4496: along $z$. Because of the azymuthal symmetry of the problem, 
4497: the actual 3D electronic
4498: density is obtained by rotating the figure around the vertical axis. 
4499: The state represents
4500: a non-dispersive wave-packet shaped like a doughnut, moving periodically from
4501: the north to the south pole (and back) of a sphere. For higher $n_0$, the
4502: angular localization on the circular orbit should improve.}
4503: \label{lin3d_8}
4504: \end{figure}
4505: For other states in the $n_0=21$ resonant manifold, the longitudinal
4506: localization along the periodic orbit is less visible. The reason is that
4507: $\chi_1$ is smaller than for the $p=20$ state, leading to a smaller
4508: resonance island in $(\hat{I},\hat{\theta})$ (see fig.~\ref{lin3d_2}) 
4509: and, consequently, to less
4510: efficient localization. Proceeding to higher $n_0$-values should improve
4511: the situation.
4512: 
4513: Let us briefly discuss ``excited'' states in the resonance island,
4514: i.e. manifolds corresponding to $N>0$ in eq.~(\ref{ebkires}). 
4515: Fig.~\ref{lin3d_9}
4516: shows the exact level dynamics, with the semiclassical prediction for
4517: $N=1$ superimposed \cite{abu98a}.
4518:  The states in this manifold originate from $n_0=22$. We
4519: observe quite good agreement between the quantum and semiclassical results for
4520: {\it high} lying states in the manifold (for which the principal action
4521: island is large, see fig.~\ref{lin3d_2}). 
4522: For lower lying states the agreement is improved for higher
4523: values of $F_0$. If $F_0$ is too low, the states are not fully localized 
4524: inside 
4525: the resonance island and, consequently,
4526: are badly reproduced by the resonant semiclassical 
4527: approximation.  This is further
4528: exemplified in fig.~\ref{lin3d_10}, for $N=2$. 
4529: Here, the
4530: agreement is worse than for smaller values of $N$, and is observed only 
4531: for
4532: large $F_0$ and large $p$. This confirms the picture that the validity
4533: of the semiclassical approach outlined here 
4534: is directly related to the size,
4535: eqs.~(\ref{area},\ref{number_of_trapped_states}), of the
4536: resonance island in $(\hat I, \hat\theta)$ space (see also the discussion
4537: in section~\ref{PTP}).
4538: \begin{figure}
4539: \centerline{\psfig{figure=bdzf25.ps,width=10cm,angle=-90}}
4540: \caption{Comparison of the numerically 
4541: exact level dynamics (dotted lines)
4542: with the
4543: semiclassical prediction (solid lines),
4544: for the $n_0=22$ ($N=1$) manifold of a 3D hydrogen atom exposed to
4545: a microwave field with frequency $\omega=1/(21)^3,$ resonant
4546: with the $n_0=21$ manifold.
4547: For sufficiently high $F_0,$ the quantum 
4548: states originating at $F_0=0$ from the unperturbed $n_0=22$ level 
4549: are captured by the principal resonance island and then represent 
4550: the first excited state of the motion in the ($\hat I, \hat \theta$)
4551: plane (i.e., $N=1$ in eq.~(\protect{\ref{ebkires}})). Since the island's 
4552: size  
4553: depends on the angular ($L,\psi$) motion (value of $p$ in 
4554: eq.~(\protect{\ref{ebkp}}), see also fig.~\protect\ref{lin3d_2}),
4555: states with large $p$ enter the resonance zone first. For these,
4556: the agreement between quantum and semiclassical quasienergies 
4557: starts to be satisfactory at lower $F_0$ values than for low-$p$ states.}
4558: \label{lin3d_9}
4559: \end{figure}
4560: \begin{figure}
4561: \centerline{\psfig{figure=bdzf26.ps,width=10cm,angle=-90}}
4562: \caption{Same as fig.~\protect{\ref{lin3d_9}}, but for $N=2$. The quantum 
4563: states originate
4564: from the manifold $n_0=20$. The 
4565: resonance island in the ($\hat I,\hat \theta$) 
4566: coordinates
4567: is too small to support any $N=2$ states for $F_0<0.03$, as seen from the
4568: negative slope of the quasienergy levels. The quantum states 
4569: ``cross'' the separatrix as the amplitude is further increased,
4570: and successively enter the resonance zone, starting from the 
4571: largest value of $p$ in eq.~(\protect{\ref{ebkp}}) 
4572: (the resonance island size increases with $p$).
4573: Even for $F_0>0.04,$ only a minority of substates of 
4574: the $n_0=20$ manifold is well represented by the resonant semiclassical 
4575: dynamics (indicated by the solid lines).}
4576: \label{lin3d_10}
4577: \end{figure}
4578: 
4579: Finally note that, as already mentioned at the end of section~\ref{LIN1D}, all
4580: wave-packet eigenstates have a finite decay rate which induces a slow, global
4581: reduction of the electronic density localized on the resonantly driven
4582: classical periodic orbit. However, the time scale of this decay is of the
4583: order of thousands to millions of Kepler cycles, and therefore leaves our
4584: above conclusions unaffected. However, some very intriguing consequences of
4585: the nonvanishing continuum coupling will be discussed in section~\ref{ION}.
4586: 
4587: \subsection{Rydberg states in circularly polarized microwave fields}
4588: \label{CP}
4589: 
4590: As shown in the preceding section, the use of a linearly polarized
4591: microwave field is not sufficient to produce a non-dispersive
4592: wave-packet fully localized in all three dimensions,
4593: due to the azymuthal symmetry around the
4594: microwave polarization axis. To get more flexibility, one may consider
4595: the case of arbitrary polarization. It turns out that the results
4596: are especially simple in circular polarization. They are the subject of this
4597: section.
4598: 
4599: In most experiments on 
4600: microwave driven Rydberg atoms,
4601: linearly polarized (LP) microwaves have 
4602: been used \cite{bayfield89,koch95b,bayfield74,bayfield96,galvez88}. 
4603: For circular polarization (CP),
4604: first experiments were performed for alkali
4605: atoms in the late eighties \cite{fu90,cheng96}, 
4606: with hydrogen atoms 
4607: following only recently \cite{bellermann96}. The latter experiments 
4608: also studied 
4609: the general case of elliptic polarization (EP). While, 
4610: at least theoretically, different
4611: frequency regimes 
4612: were considered for CP microwaves
4613: (for a review see \cite{delande97b}) --
4614: we shall restrict 
4615: our discussion here to 
4616: resonant driving.
4617: Given a different 
4618: microwave polarization, and thus a different form 
4619: of the interaction Hamiltonian, eq.~(\ref{vcirc}),
4620: Kepler trajectories which are distinct from those considered in the LP case
4621: will be most efficiently locked on the external driving. Hence, in the sequal,
4622: we shall launch nondispersive wave-packets along periodic orbits which are
4623: distinct from those encountered above.
4624: 
4625: Historically, the creation of non-dispersive wave-packets in 
4626: CP and LP microwave fields,
4627: respectively, 
4628: has been considered quite independently.
4629: In particular, in the CP case, the notion of nondispersive 
4630: wave-packets has been introduced \cite{ibb94} 
4631: along quite 
4632: different lines 
4633: than the one adopted in this
4634: review. The original work, as well as subsequent studies of the CP situation
4635: \cite{farrelly95a,ibb95,kalinski95a,delande95,kuba95a,kalinski96b,kuba97a,kuba97b}
4636: used 
4637: the fact that, in this specific case, 
4638: the time-dependence of the
4639: Hamiltonian may be removed by a unitary transformation to the rotating frame
4640: (see below). Thus, the stable periodic orbit 
4641: at the center of the island 
4642: turns into a stable equilibrium point in the rotating frame. This allows the
4643: expansion of the Hamiltonian into a Taylor series in the vicinity of 
4644: the fixed point,
4645: and in particular a standard harmonic treatment using normal modes. 
4646: We shall review 
4647: this line of reasoning in detail below. It is, however, instructive 
4648: to first 
4649: discuss
4650: the very same system using the general resonance approach
4651: exposed in section~\ref{GM}. 
4652: 
4653: 
4654: \subsubsection{Hamiltonian}
4655: 
4656: With $\hat z$ the propagation direction of the microwave,
4657: the electric field rotates
4658: in the $x-y$ plane, and the Hamiltonian (\ref{h}) takes
4659: the
4660: following explicit form:
4661: \begin{equation}
4662: H_{\rm CP}=\frac{\vec{p}^2}{2}-\frac{1}{r}+
4663: F\left\{x\cos(\omega t)+y\sin(\omega t)\right\}.
4664: \label{ham_cp_3d}
4665: \end{equation}
4666: 
4667: In contrast to the LP 
4668: situation, there is no simplified one-dimensional
4669: model in the CP case. However, a simplified two-dimensional 
4670: model 
4671: exists, where
4672: the motion is restricted to the $x-y$ plane. As long as one is interested only
4673: in the dynamics of non-dispersive wave-packets, this motion is 
4674: stable (see below),
4675: which means that a small deviation from the $z=0$ plane does not affect the
4676: qualitative behavior. Hence, much physical insight can be obtained
4677: from the simplified 2D model. It 
4678: will be discussed in section \ref{2d_model}.
4679: 
4680: \subsubsection{Resonance analysis}
4681: \label{reso_an}
4682: We follow the general treatment exposed in sections~\ref{CD} and \ref{LIN1D}
4683: for the 1D case,
4684: and 
4685: express the external perturbation as a function of the action-angle
4686: variables 
4687: $(I,\theta)$, $(L,\psi)$ and $(M,\phi)$ introduced in section~\ref{RSEF}. 
4688: By inserting equations~(\ref{v_3d_x},\ref{v_3d_y}) in eq.(\ref{ham_cp_3d}), 
4689: after appropriate account for the projection of the body-fixed frame 
4690: $(x',y',z')$ onto the laboratory frame $(x,y,z)$, eq.~(\ref{rot_euler}),
4691: we obtain for the Floquet Hamiltonian:
4692: \begin{equation}
4693: {\mathcal H}=
4694: P_t
4695: -\frac{1}{2I^{2}}+F   \sum_{m=-\infty}^{\infty}\left[
4696:     V_m \cos(m\theta+\phi-\omega t)
4697:     -U_m
4698:     \sin(m\theta+\phi-\omega t)\right],
4699: \label{h_circ_aa}
4700: \end{equation}
4701: where the Fourier coefficients are given by 
4702: (see also eqs.~(\ref{xm}-\ref{y0})):
4703: \begin{eqnarray}
4704:  V_{m}(I,L,M,\psi) &=& \cos\psi \left(Y_m + \frac{M}{L} X_m\right), \label{Vb}\\
4705:  U_{m}(I,L,M,\psi) &=& \sin\psi \left(X_m + \frac{M}{L} Y_m\right).
4706: \label{Ub}
4707: \end{eqnarray}
4708: Once again, transformation 
4709: to the ``rotating frame", eqs.~(\ref{rotframe_a}-\ref{rotframe_c}), and
4710: averaging over one field period $T=2\pi/\omega$ (thereby neglecting all 
4711: rapidly varying terms) leaves us with the explicit form of the
4712: secular Hamiltonian:
4713: \begin{equation}
4714: {\mathcal H}_{\rm sec}=\hat{P}_t-\frac{1}{2\hat I^2}-\omega \hat I
4715: +F\left[V_1(\hat{I},L,M,\psi)\cos (\hat{\theta}+\phi) - 
4716: U_1(\hat{I},L,M,\psi)\sin (\hat{\theta}+\phi) \right],
4717: \label{hamsccir}
4718: \end{equation}
4719: similar to eq.~(\ref{hamsc2}).
4720: This can be rewritten as:
4721: \begin{equation}
4722: {\mathcal H}_{\rm sec}=\hat{P}_t-\frac{1}{2\hat I^2}-\omega \hat I
4723: +F \chi_1 \cos (\hat{\theta}+\phi+\delta_1)  ,
4724: \label{hamsccir2}
4725: \end{equation}
4726: with the effective perturbation
4727: \begin{equation}
4728: \chi_1(\hat{I},L,M,\psi)=\sqrt{V_1^2+U_1^2},
4729: \label{widthcir}
4730: \end{equation}
4731: and
4732: \begin{equation}
4733: \tan \delta_1(\hat{I},L,M,\psi) = \frac{U_1}{V_1}.
4734: \label{phasecirc} 
4735: \end{equation}
4736: 
4737: This secular Hamiltonian, which has, once again, 
4738: the same structure as the 1D secular Hamiltonian~(\ref{hsec_ap}),
4739: governs the ``slow" dynamics of the system
4740: in the vicinity of the resonance. Similarly to the LP case, the
4741: various degrees of freedom evolve on different time scales:
4742: \begin{itemize}
4743: \item In the $(\hat{I},\hat{\theta})$ plane, the situation is exactly like
4744: for a one-dimensional system. There is a resonance island around
4745: the resonant action,
4746: with a pendulum-like structure. Non-dispersive wave-packets
4747: are associated with eigenstates localized at the center of this island,
4748: at the point (see eqs.~(\ref{scaling_action},\ref{hamsccir2})):
4749: \begin{equation}
4750: \hat{I}_1 = \omega^{-1/3} = n_0,\ \ \ \ \hat{\theta} = -(\phi + \delta_1).
4751: \label{center_island_CP}
4752: \end{equation}
4753: The period of the secular 
4754: classical motion close to the resonance center scales as 
4755: $1/\sqrt{F}$. 
4756: It 
4757: defines an
4758: intermediate time scale, slower than
4759: the Kepler frequency, but faster than the transverse motion in the
4760: other coordinates $(L,\psi)$ and $(M,\phi)$. 
4761: \item In the subspace spanned by $(L,M,\psi,\phi)$, the motion is much 
4762: slower, with
4763: a time scale proportional to $1/F$. The effective Hamiltonian
4764: describing this motion is obtained by averaging the fast
4765: motion in the perturbation which describes
4766: this motion $(\hat{I},\hat{\theta})$ plane, which in turns
4767: implies that $\chi_1$ itself is constant for both the motion in 
4768: $(\hat{I},\hat{\theta})$ and $(L,M,\psi,\phi)$ space.
4769: \end{itemize}
4770: Note that
4771: $\chi_1$ does not depend on the angle $\phi$. This,
4772: in turn, implies that $M$ is a constant of the slow motion. 
4773: This is because the circular polarization
4774: does not define any preferred direction in the  $x-y$ polarization plane.
4775: 
4776: Once again, much alike our discussion in section~\ref{LIN3D},
4777: the well-known properties of Bessel functions
4778: \cite{abramowitz72}, together with eqs.~(\ref{Vb},\ref{widthcir}), imply
4779: that, for given $M$, the maximum of $\chi_1$ occurs at 
4780: $L=M$,
4781: corresponding to the situation when the electronic motion is
4782: restricted to the polarization plane. In this plane, 
4783: the maximum $\chi_1=\hat{I}^2$
4784: is reached for the circular orbit defined by $M=L=\hat{I}$ (i.e. 
4785: $L_0=1$). 
4786: This defines
4787: a resonant periodic orbit locked on the 
4788: external microwave driving, which maximizes the
4789: effective Hamiltonian in each coordinate and is, therefore,
4790: fully 
4791: stable in all phase space directions. 
4792: The orbit is a circular Kepler orbit in the polarization
4793: plane, where the electron rotates around the nucleus with exactly
4794: the angular velocity of the microwave. It is not really surprising
4795: that this orbit maximizes the interaction energy with the external
4796: field: indeed, along this orbit, the atomic dipole rotates 
4797: exactly in phase with the polarization vector of 
4798: the circularly polarized microwave field. 
4799: As in the case of linear polarization discussed in 
4800: section~\ref{LIN3D}, the angular motion in the
4801: $(L,\psi,M,\phi)$ variables (which is trivial
4802: in $(M,\phi)$, since $M$ is constant) could be studied in detail.
4803: For the sake of brevity, we will not repeat such an analysis here. We rather
4804: concentrate on the  wave-packets 
4805: which are best localized in the resonance island
4806: near the circular orbit.
4807: The simplest approximation 
4808: to describe these states is to replace
4809: the largest quantized value of $\chi_1$ by its maximum value $\hat{I}^2$
4810: estimated at the center of the resonance island, eq.~(\ref{center_island_CP}).
4811: Then, the situation is similar to the 1D model of the atom, 
4812: eq.~(\ref{hsec_h1d}),
4813: except that the strength of the coupling is $\hat{I}^2$ instead of 
4814: $-J_1'(1)\hat{I}^2.$ In complete analogy to the steps leading from 
4815: eq.~(\ref{hsec_h1d}) to eqs.~(\ref{semiharm1dval},\ref{omega_h1d}) we employ 
4816: the pendulum approximation with a subsequent harmonic expansion
4817: around the pendulum's stable equilibrium point, 
4818: deeply inside the resonance island.
4819: The harmonic frequency of the motion in the $(\hat I,\hat\theta)$ plane is:
4820: \begin{equation}
4821: \omega_{\mathrm harm}= 
4822: \frac{\sqrt{3F}}{n_0} = \omega \sqrt{3F_0},
4823: \label{omega_harmonic_cp}
4824: \end{equation}
4825: and the quasi-energy levels are:
4826: \begin{equation}
4827: {\cal E}_{N,k} = k\omega -\frac{3}{2n_0^2} + n_0^2F - \left(N+\frac{1}{2}\right) \omega_{\mathrm harm}.
4828: \label{estharm}
4829: \end{equation}
4830: Note that, by construction, 
4831: $-3/2n_0^2 + F n_0^2$ is nothing but the energy at the center
4832: of the resonance island, i.e. the energy of the resonant circular orbit.
4833: For very small $F$, the resonance island shrinks and may
4834: support only a small number of states, or even no state at all. 
4835: In this regime, the harmonic
4836: approximation, eq.~(\ref{spectrum_harmonic}), breaks down.
4837: Alternatively, one can apply a quantum treatment
4838: of the pendulum motion in the $(\hat{I},\hat{\theta})$ plane, as explained
4839: in section~\ref{section_mathieu} and discussed in section~\ref{LIN1D} 
4840: for the
4841: 1D model of the atom exposed to a linearly polarized
4842: microwave. The analysis -- essentially identical to the one in
4843: section~\ref{LIN1D} --  yields the following expression for the
4844: energy levels:
4845: \begin{equation}
4846: {\cal E}_{k,N}= k\omega- \frac{3}{2n_0^2}-\frac{3a_N(\nu,q)}{8n_0^4},
4847: \end{equation}
4848: where $a_N(\nu,q)$ are the Mathieu eigenvalues (compare with 
4849: eq.~(\ref{map_mathieu2})
4850: for the general case), with
4851: \begin{equation}
4852: q=\frac{4}{3}Fn_0^6,
4853: \end{equation}
4854: and 
4855: \begin{equation}
4856: \nu=-2n_0 \ \ ({\mathrm mod}\ 2)
4857: \end{equation}
4858: the characteristic exponent.
4859: 
4860: These expressions are valid for the states localized close to the resonant
4861: circular orbit.
4862: For the other states, the calculation is essentially identical, the
4863: only amendment being the use of the values 
4864: of $\chi_1$
4865: following from the quantization of the secular motion, instead of the maximum
4866: value $n_0^2.$ 
4867: 
4868: Finally, as the center of the resonance island corresponds to a circular
4869: trajectory in the $(x,y)$ plane, the Floquet states associated with 
4870: the non-dispersive 
4871: wave-packets
4872: will be essentially composed of combinations of circular states 
4873: $|n,\ L=M=n-1\rangle$,
4874: with coefficients described by the solutions
4875: of the Mathieu equation, as explained in section~\ref{section_mathieu}.
4876: This Mathieu formalism has been rediscovered in this particular CP situation
4877: via complicated approximations 
4878: on the exact Schr\"odinger equation in \cite{kalinski96b}. 
4879: We believe that the standard resonance analysis using the
4880: pendulum approximation leads, at the same time, to simpler calculations, 
4881: and to
4882: a much more transparent
4883: physical picture.
4884:  
4885: \subsubsection{The two-dimensional model}
4886: \label{2d_model}
4887: 
4888: We shall now discuss the simplified 2D model of the CP problem, which 
4889: amounts to restricting the motion to the
4890: $(x,y)$ plane, but retains almost all the features of the full 3D problem. 
4891: Instead of the six-dimensional phase space 
4892: spanned by the 
4893: action-angle variables $(I,\theta)$, $(L,\psi)$, $(M,\phi)$, one is left
4894: with a four-dimensional submanifold 
4895: with coordinates $(I,\theta)$, $(M,\phi),$ see section~\ref{action-angle_hydrogen}. 
4896: The secular Hamiltonian  
4897: then reads 
4898: (compare eqs.~(\ref{hamsccir2},\ref{widthcir})):
4899: \begin{equation}
4900: {\mathcal H}_{\rm sec}=\hat{P}_t-\frac{1}{2\hat I^2}-\omega \hat I
4901: +FV_1(\hat{I},M)\cos (\hat{\theta}+\phi)  ,
4902: \label{hamsccir_2d}
4903: \end{equation}
4904: with (see eq.~(\ref{Vb}))
4905: \begin{equation}
4906: V_1(\hat{I},M)= \hat{I}^2 \left[ J_{1}^{'}(e) + {\mathrm sign}(M)
4907:  \frac{\sqrt{1-e^2}}{e}J_{1}(e) \right] ,
4908: \end{equation}
4909: where (as in eq.~(\ref{eccentricity}))
4910: \begin{equation}
4911: e=\sqrt{1-\frac{M^2}{\hat{I}^2}}.
4912: \end{equation}
4913: 
4914: However, the Maslov index for the $(\hat{I},\hat\theta)$ motion is different.
4915: Indeed, the energy spectrum of 
4916: the 2D atom is given by eq.~(\ref{spectrum_2d}).
4917: Thus, quantized values of the action
4918: are half-integer multiples of $\hbar$. The relation
4919: between the resonant action $\hat{I}_1=\omega^{-1/3}$ and the
4920: corresponding principal quantum number now reads (with $\hbar=1$):
4921: \begin{equation}
4922: n_0=\hat{I}_1-\frac{1}{2}.
4923: \end{equation}
4924: As explained in section~\ref{section_mathieu}, the optimal
4925: case for the preparation of non-dispersive wave-packets --
4926: where the states are the most deeply bound
4927: inside the resonance island --  is for integer
4928: values of $n_0,$ i.e. frequencies
4929: (compare with eqs.~(\ref{resonance_condition},\ref{omk}))
4930: \begin{equation}
4931: \omega = \frac{1}{\left(n_0+\frac{1}{2}\right)^3}.
4932: \end{equation}
4933: For the energy levels of the non-dispersive wave-packets,
4934: this also implies that the characteristic exponents in the Mathieu
4935: equation -- see section~\ref{section_mathieu} -- are shifted
4936: by one unit:
4937: \begin{equation}
4938: \nu=-2n_0 \ \ ({\mathrm mod}\ 2)\ \ = -2\hat{I}_1+1\ \ ({\mathrm mod}\ 2).
4939: \end{equation}
4940: 
4941: 
4942: \subsubsection{Transformation to the rotating frame}
4943: \label{rotating_frame}
4944: 
4945: The resonance analysis developed above is restricted to first
4946: order in the amplitude $F$ of the external drive. Extensions to higher
4947: orders are possible, but tedious. For CP,
4948: an alternative approach is possible,
4949: which allows higher orders to be included quite easily.
4950: It is 
4951: applicable to CP only and thus
4952: lacks the generality of the resonance approach we used so far. 
4953: Still, it is rather
4954: simple and deserves an analysis.
4955: 
4956: In CP, one may remove the time dependence of the
4957: Hamiltonian (\ref{ham_cp_3d}) by a transformation 
4958: to the noninertial frame rotating
4959: with the external frequency $\omega$. The unitary fransformation
4960: $U=\exp(i\omega L_z t)$ leads to \cite{bunkin64,grozdanov92}
4961: \begin{equation}
4962: H_{\rm rot}= U \ H_{\rm CP}\ U^{\dagger} + 
4963: i U \frac{\partial U^{\dagger}}{\partial t}
4964: =
4965: \frac{\vec{p}^2}{2} -\frac{1}{r} +Fx -\omega L_z.
4966: \label{hrot}
4967: \end{equation}
4968: Classically, such an operation corresponds to a time dependent rotation of the
4969: coordinate frame spanned by $\overline{x} = x\cos \omega t +y\sin \omega t$, 
4970: $\overline{y} = y\cos \omega t-x \sin \omega t$ 
4971: (and dropping the bar 
4972: hereafter)\footnote{Passing to the rotating frame implies  
4973: a change of 
4974: $\phi$ to $\overline{\phi}=\phi-\omega t$ in eq.~(\ref{h_circ_aa}). 
4975: That is definitely
4976: different from the change $\theta \to \hat{\theta}=\theta-\omega t$, 
4977: eq.~(\ref{rotframe_a}), used in the resonance analysis. 
4978: Both transformations are unfortunately
4979: known under the same name of ``passing to the rotating frame". This is
4980: quite confusing, but one has to live with it. Along the resonantly driven
4981: circular orbit we are considering here, 
4982: it happens that the azimuthal angle $\phi$ and the polar angle 
4983: $\theta$ actually coincide.
4984: It follows that the two approaches are equivalent {\em in the vicinity}
4985: of this orbit.}.
4986: The Hamiltonian~(\ref{hrot}), as a time-independent operator, has some energy
4987: levels and corresponding eigenstates. Its spectrum is not 
4988: $\omega$-periodic, although the unitary transformation assures that there is
4989: a one-to-one correspondence between its spectrum (eigenstates) and the
4990: Floquet spectrum 
4991: of eq.~(\ref{ham_cp_3d})\footnote{In
4992: fact, if $|\phi_i\rangle$ is an eigenstate of $H_{\rm rot}$ with energy $E_i$,
4993:  then $U^{\dagger}|\phi_i \rangle$ is
4994: a Floquet eigenstate with quasi-energy $E_i$, while states shifted in
4995: energy by $k\omega$ are of the form $\exp (ik\omega t)U^{\dagger}|\phi_i
4996:  \rangle.$ For a more detailed discussion of this point, 
4997: see \cite{delande98}.}.
4998: It was observed \cite{klar89} that the Hamiltonian~(\ref{hrot})
4999: allows for
5000: the existence of a stable fixed (equilibrium) 
5001: point in a certain range of the microwave 
5002: amplitude $F$. Later on, it was realized \cite{ibb94}
5003: that wave-packets initially
5004: localized in the vicinity of this fixed point will not disperse
5005: (being bound by the fact that the fixed point is stable) for 
5006: at least several Kepler periods. 
5007: In the laboratory frame, these wave-packets (also
5008: called ``Trojan states" \cite{farrelly95a,ibb95,ibb94,kalinski95a})
5009: appear as wave-packets moving around the
5010: nucleus along the circular trajectory, which is nothing
5011: but the periodic orbit at the center of the resonance island discussed in 
5012: section~\ref{reso_an}.
5013: In the original formulation \cite{ibb94} and the discussion which followed
5014: \cite{farrelly95a,ibb95,kalinski95a,kalinski96b},
5015: great attention was paid to the accuracy of the harmonic
5016: approximation (see below). This was of 
5017: utmost importance for 
5018: the non-spreading
5019: character of {\it Gaussian-shaped} Trojan wave-packets considered in
5020: \cite{farrelly95a,ibb95,ibb94,kalinski95a,kalinski96b}. 
5021: As soon pointed out in \cite{delande95}, however, 
5022: the accuracy of
5023: this approximation
5024: is immaterial for the very existence of
5025: the wave-packets, which are to be identified, as shown above, with
5026: well-defined Floquet states.
5027: 
5028: Let us 
5029: recapitulate the fixed point analysis of \cite{ibb94,klar89}
5030: in the rotating frame.
5031: Inspection of the classical version of the  Hamiltonian $H_{\rm rot}$,
5032: eq.~(\ref{hrot}), 
5033: shows that, due to symmetry, 
5034: one may seek the fixed point
5035: at $z=y=0$. The condition for an equilibrium (fixed) point,
5036:  i.e., $d\vec{r}/dt=0,\ d\vec{p}/dt=0$,
5037: yields immediately that $p_{z,\rm eq}=p_{x, \rm eq}=0$, 
5038: $p_{y,\rm eq}=\omega x_{\rm eq}$,
5039: with the subscript ``$\rm eq$'' for 
5040: ``equilibrium''.
5041: The remaining equation
5042: for $d p_x/dt$ gives the condition
5043: \begin{equation}
5044:   -F+\omega^2x_{\rm eq}-\frac{|x_{\rm eq}|}{x_{\rm eq}^3}=0,
5045: \label{wp0}
5046: \end{equation}
5047: that defines the position of the fixed point as a function of $F$.  
5048: Following \cite{ibb94} let us introduce the 
5049: dimensionless parameter
5050: \begin{equation}
5051: q=\frac{1}{\omega^{2}|x_{\rm eq}|^{3}}.
5052: \label{q}
5053: \end{equation}
5054: One may easily express the fixed point position, the microwave field
5055: amplitude, as well as the corresponding energy in terms of $q$ and $\omega$.
5056: Explicitly:
5057: \begin{equation}
5058: x_{{\rm eq}1}=\frac{1}{q^{1/3}\omega^{2/3}}, 
5059: \qquad F=\frac{1-q}{q^{1/3}}\omega^{4/3},
5060:  \qquad  E_{\rm eq}=\frac{1-4q}{2}\left(\frac{\omega}{q}\right)^{2/3},
5061: \label{wps}
5062: \end{equation}
5063: and
5064: \begin{equation}
5065: x_{{\rm eq}2}=-\frac{1}{q^{1/3}\omega^{2/3}}, 
5066: \qquad F=\frac{q-1}{q^{1/3}}\omega^{4/3},
5067:  \qquad  E_{\rm eq}=\frac{1-4q}{2}\left(\frac{\omega}{q}\right)^{2/3}.
5068: \label{wpu}
5069: \end{equation}
5070: For $F=0$,  $q=1$ in eqs.~(\ref{wps},\ref{wpu}).
5071: For $F<0,$ (i.e., $q>1$ in eq.~(\ref{wps})), $x_{{\rm eq}1}$ 
5072: is an unstable fixed point, while for moderately
5073: positive $F$ (i.e., $8/9<q<1$ \cite{ibb94,klar89}) it is stable.
5074: Stability of the second equilibrium point  $x_{{\rm eq}2}$
5075: is achieved by changing the sign of $F.$ For moderate
5076: fields ($q$ close to unity), the stable and the unstable fixed points 
5077: are located on
5078: opposite sides of the nucleus, and at almost the same distance from it. 
5079: As the whole analysis is classical, it has to obey the scaling laws
5080: discussed in section~\ref{scaling_laws}. Hence, all quantities describing
5081: the equilibrium points in the preceding equations scale as powers
5082: of the microwave frequency. A consequence
5083: is that there is a very simple correspondence
5084: between the scaled microwave amplitude and the dimensionless parameter $q$:
5085: \begin{equation}
5086: F_0=F\omega^{4/3} = F n_0^4 = \frac{1-q}{q^{1/3}}.
5087: \end{equation}
5088: The parameter $q$ can thus be thought of as 
5089: a convenient parametrization (leading to simpler algebraic formula)
5090: of the scaled microwave amplitude.
5091: 
5092: A fixed point in the rotating frame corresponds
5093: to a periodic orbit with exactly the period $T=2\pi/\omega$ of the
5094: microwave driving field in
5095: the original frame. The stable fixed point
5096: (periodic orbit) thus corresponds to the center of the resonance island, and to
5097: the stable equilibrium point of the pendulum in the secular approximation.
5098: Similarly, the unstable fixed point corresponds to the
5099: unstable equilibrium point of the pendulum.
5100: Note that the stable fixed point approaches
5101: $x_{\rm eq}=\omega^{-2/3}=\hat{I}_1^2$ when $F\rightarrow 0$, 
5102: i.e. the radius of the circular classical Kepler 
5103: trajectory with frequency $\omega$.
5104: Thus,
5105: the stable fixed point
5106: smoothly reaches the location  of the circular state 
5107: of the hydrogen atom, with a classical Kepler frequency equal
5108: to the driving microwave frequency.  Its energy $E_{\rm eq}=-3\omega^{2/3}/2$
5109: is the energy of the circular orbit in the rotating frame.
5110: 
5111: Since the non-dispersive wave-packets are localized in
5112: the immediate vicinity of the stable fixed point in the rotating frame, an expansion of the
5113: Hamiltonian around that position is useful. Precisely at the fixed point, all
5114: first order terms (in position and momentum) vanish. At second order,
5115: \begin{eqnarray}
5116: H_{\mathrm rot} \simeq H_{\mathrm harmonic}&&=E_{\rm eq} +
5117: \frac{{\vec{\tilde p}}^2}{2}
5118:  - \omega(\tilde{x}\tilde{p}_y-\tilde{y}\tilde{p}_x)
5119: + \frac{\omega^2 q \tilde{y}^2}{2}
5120: \nonumber\\ && - \omega^2 q \tilde{x}^2
5121: + \frac{\omega^2 q \tilde{z}^2}{2}\! ,
5122: \label{harmonic} 
5123: \end{eqnarray}
5124: where 
5125:  $(\tilde{x},\tilde{y},
5126: \tilde{z})=(x-x_{\rm eq},y-y_{\rm eq},z-z_{\rm eq})$ 
5127: (and accordingly for the
5128: momenta) denotes the displacement with respect
5129: to the fixed point. 
5130: Thus, in the harmonic approximation
5131: the motion in the $z$ direction
5132: decouples from that in the $x-y$ plane and is an oscillation with
5133: frequency $\omega\sqrt{q}.$ 
5134: The Hamiltonian for the latter,
5135: up to the additive constant $E_{\rm eq}$,  can be expressed in the
5136: standard form for a 2D, rotating anisotropic oscillator 
5137: \begin{equation}
5138: H= \frac{\tilde{p}_x^2+\tilde{p}_y^2}{2} +\frac{\omega^2
5139: (a\tilde{x}^2+b\tilde{y}^2)}{2}-\omega(\tilde{x}\tilde{p}_y-\tilde{y}\tilde{p}_x),
5140: \label{rotaibb}
5141: \end{equation}
5142: where the two parameters $a$ and $b$ are equal to $-2q$ and $q$,
5143: respectively. 
5144: This standard form has been studied in textbooks~\cite{marion}.
5145: It may be used to describe the stability of 
5146: the Lagrange equilibrium points in
5147: celestial mechanics (see \cite{ibb94,marion} and references therein).
5148: Because this Hamiltonian mixes position and momentum coordinates,
5149: it is not straightforward to determine the stability at the origin.
5150: The result is that there are two domains of stability:
5151: \begin{equation}
5152: a,b \geq 1
5153: \label{stability_region_1}
5154: \end{equation}
5155: and
5156: \begin{equation}
5157: -3 \leq a,b \leq 1,\ \ \ {\mathrm with}\ \ (a-b)^2+8(a+b) \geq 0
5158: \label{stability_region_2}
5159: \end{equation}
5160: 
5161: For the specific CP case, where $a=-2q$ and $b=q,$ only the second
5162: stability region is relevant, and the last inequality
5163: implies $8/9\leq q\leq 1$ for the fixed point to be stable.
5164: 
5165: Alternatively, one can ``diagonalize" the Hamiltonian~(\ref{rotaibb})
5166: and construct 
5167: its normal modes. The normal modes entangle position and momentum
5168:  operators due
5169: to the presence of crossed position-momentum terms in the
5170: Hamiltonian.
5171: Only along the $z$-mode (which is decoupled from the rest), the creation 
5172: and
5173: annihilation operators, $b_z^\dagger,b_z$ are
5174:  the standard combinations of $\tilde z$
5175: and $p_{\tilde z}$ operators. 
5176: In the $(x,y)$ plane, the creation and annihilation operators in the
5177: $\pm$ normal modes have complicated explicit formulae given in~\cite{delande98}. 
5178: After some algebra, one ends up with 
5179: the frequencies of the normal modes,
5180: \begin{eqnarray}
5181:  \omega_\pm &=& \omega \sqrt{\frac{2-q\pm Q}{2}},\label{omegas}\\
5182: \omega_z &=& \omega \sqrt{q},
5183: \label{ompmz}
5184: \end{eqnarray}
5185: where
5186: \begin{equation}
5187: Q=\sqrt{9q^2-8q}
5188: \label{defQ}
5189: \end{equation}
5190: and $q\geq 8/9$ for $Q$ to be real.
5191: In terms of  
5192: creation/annihilation operators, the harmonic Hamiltonian,
5193: eq.~(\ref{harmonic}),
5194: takes the form \cite{ibb94,delande95,kuba95a,delande98}
5195: \begin{equation}
5196: H_{\rm harmonic}
5197:  = E_{\rm eq} + \left(b^\dagger_+ b_+ +\frac{1}{2}\right)\omega_+ 
5198: - \left(b^\dagger_- b_-+\frac{1}{2}\right)\omega_- 
5199:  + \left(b^\dagger_z b_z+\frac{1}{2}\right)\omega_z .
5200: \label{hh}
5201: \end{equation}
5202: A minus sign appears in front of the $\omega_-$ term. This is because
5203: the fixed point is {\em not} a minimum of the Hamiltonian, although
5204: it is fully stable\footnote{In the first stability region, 
5205: eq.~(\ref{stability_region_1}),
5206: only + signs appear.}. 
5207: This is actually due to the momentum-position
5208: coupling, hence the Coriolis force. It is the same phenomenon
5209: which is responsible for the stability of the Trojan asteroids 
5210: \cite{ibb94,yeazell00,marion}
5211: and of an ion in a magnetic trap \cite{lee95} 
5212: (in the latter case, the position-momentum
5213: coupling is due to the magnetic field).
5214: 
5215: Finally, with $n_\pm, n_z$ counting
5216: the excitations in the corresponding modes, we obtain the
5217: harmonic prediction for the energies of the eigenstates in the vicinity of the fixed
5218: point: 
5219:  \begin{equation}
5220: E(n_+,n_-,n_z) = E_{\rm eq} + 
5221: \left(n_++\frac{1}{2}\right)\omega_+
5222:  - \left(n_-+\frac{1}{2}\right)\omega_-
5223: +  \left(n_z+\frac{1}{2}\right)\omega_z .
5224: \label{enharm}
5225: \end{equation}
5226: In particular, for $n_\pm=0,\ n_z=0$, we get a prediction for the ground 
5227: state of the oscillator, a Gaussian localized on top of the fixed point, i.e.,
5228: a Trojan wave-packet. In the following, we denote eigenstates in the 
5229: harmonic approximation
5230: as $|n_+,n_-,n_z\rangle$, thus the ground state non-dispersive wave-packet
5231: as $|0,0,0\rangle$. 
5232: In a 2D model, the $\omega_z$ term is dropped, the
5233: corresponding eigenstates are denoted $|n_+,n_-\rangle$ and have energies:
5234: \begin{equation}
5235: E^{2D}(n_+,n_-) = E_{\rm eq} + 
5236: \left(n_++\frac{1}{2}\right)\omega_+
5237:  - \left(n_-+\frac{1}{2}\right)\omega_-.
5238: \label{enharm2d}
5239: \end{equation}
5240: In fig.~\ref{cp2df}, we show the probability densities of the $|0,0\rangle$
5241: wave-packets obtained by exact numerical diagonalization of the  
5242: 2D Hamiltonian (\ref{hrot}), for various
5243: values of the microwave field amplitude.
5244: Clearly, for sufficiently strong microwave amplitudes,
5245: the wave-packets are well localized around
5246: the classical stable fixed point, with banana-like shapes.
5247: At very 
5248: weak fields, the stability of the
5249: fixed point gets 
5250: weaker and weaker; for a vanishing microwave field, all points on the
5251: circle with radius $\omega^{-2/3}$ are equivalent, and one has a ring of
5252: equilibrium points. Thus, when $F$ tends to zero,
5253: the non-dispersive wave-packet progressively extends along the angular direction
5254: (with the radial extension almost unchanged), ending with a doughnut shape
5255: at vanishing field. This means that, if $n_0$ is chosen
5256: as an integer,
5257: the non-dispersive wave-packet smoothly evolves into a circular
5258: state $|n_0,M=n_0\rangle$ as $F\rightarrow 0$. 
5259: The same is true for the 3D
5260: atom, where the non-dispersive wave-packet smoothly evolves into the
5261: circular state $|n_0,L=M=n_0-1\rangle.$
5262: \begin{figure}
5263: \centerline{\psfig{figure=bdzf27.ps,width=10cm}}
5264: \caption{Non-dispersive wave-packets of the two-dimensional hydrogen atom
5265: driven by a circularly polarized microwave field, for different values 
5266: of the microwave amplitude $F$.
5267: The microwave frequency is fixed at $\omega=1/(60.5)^3$,
5268: corresponding to a resonance island centered at
5269: $n_0=60.$ With increasing
5270: microwave amplitude, more states are coupled and the wave-packet becomes
5271: better localized. The scaled microwave amplitude $F_0=F\omega^{-4/3}$ 
5272: is $0.0003$, $0.0011$,
5273: $0.0111$, $0.0333$, $0.0444$, $0.0555$, from top left to bottom right.
5274: The nucleus is at the center of the figure, which extends over
5275: $\pm 5000$ Bohr radii in each direction. The microwave field is horizontal,
5276: pointing to the right.}
5277: \label{cp2df}
5278: \end{figure}
5279: 
5280: Provided the resonance island around the fixed point 
5281: is large enough,
5282: the harmonic approximation can also be used for studying properties
5283: of ``excited" states inside the resonance island.
5284: As an example, 
5285: fig.~\ref{accharm} shows the $|n_+=1,n_-=3\rangle$ 
5286: state calculated from the harmonic approximation, compared to
5287: the state obtained by exact numerical diagonalization of the 2D Floquet
5288: Hamiltonian. Obviously, the structure of the exact state is very similar to
5289: the one obtained from 
5290: its harmonic approximation. Because the creation and annihilation
5291: operators in the $\omega_{\pm}$ modes entangle position
5292: and momentum coordinates in a complicated way~\cite{delande98}, and 
5293: although the system
5294: is then completely integrable, the wave-function in the
5295: harmonic approximation is {\em not} separable in any coordinate system 
5296: (in contrast with the usual harmonic oscillator). Actually, the wave-function
5297: can be written as a product of Gaussians and Hermite polynomials of 
5298: the position
5299: coordinates, but the Hermite polynomials have to be evaluated for 
5300: complex values.
5301: This results in the unusual pattern of the probability density displayed in
5302: fig.~\ref{accharm}. An improvement over
5303: the harmonic approximation is possible,
5304: by bending the axis in the spirit of~\cite{kalinski95a}, in order to 
5305: account for the spherical symmetry of the dominant Coulomb potential. 
5306: With this
5307: improvement, the probability density, shown in the middle row of 
5308: fig.~\ref{accharm},
5309: is almost indistinguishable from the exact result. Let us repeat that
5310: this bending -- and consequently the deviation from Gaussian character of the
5311: wave-function -- does not affect at all the non-dispersive character of the 
5312: wave-packet.
5313: \begin{figure}
5314: \centerline{\psfig{figure=bdzf28.ps,width=10cm} }
5315: \caption{Comparison of exact Floquet eigenstates (bottom row) and of
5316: their harmonic
5317: approximations, for the ground state wave-packet $|n_+=0,n_-=0\rangle$ 
5318: (left column),
5319: and for the $|n_+=1,n_-=3\rangle$ excited state (right column), 
5320: for the 2D hydrogen atom
5321: driven by a circularly polarized microwave (field pointing to the
5322: right of the figure), 
5323: with amplitude $F_0=0.0333$ and resonant frequency
5324: corresponding to $n_0=60$. The nucleus is located at the center of each plot
5325: which extend over $\pm 5000$ Bohr radii. 
5326: The top row represents eigenfunctions in the harmonic
5327: approximation in $\tilde{x},\tilde{y}$ coordinates, 
5328: eq.~(\protect\ref{rotaibb}).
5329: The eigenfunctions in the middle row are obtained from the harmonic 
5330: approximation to eq.~(\protect\ref{hrot}) in polar coordinates. 
5331: They exhibit a clear bending of the electronic density along the circular 
5332: trajectory.
5333: The excited
5334: wave-packet $|1,3\rangle$ appears in 
5335: fig.~\protect{\ref{solit2}} as a straight 
5336: line (modulo small avoided crossings) 
5337: with a negative slope, meeting the state $|0,0\rangle$ in a broad avoided
5338: crossing, around $F_0\simeq 0.036$.
5339: }
5340: \label{accharm}
5341: \end{figure}
5342: 
5343: Let us now turn to the realistic 3D model of the atom.
5344: Fig.~\ref{circwp3d} shows an isovalue contour of several 
5345: non-dispersive wave-packets
5346: for the hydrogen atom driven by a microwave
5347: field with frequency
5348: $\omega=1/60^3$, i.e., roughly resonant with 
5349: the $n_0=60\to 59,61$ transitions 
5350: (see eqs.~(\ref{cprinc},\ref{resonance_condition},\ref{omk})). 
5351: The 
5352: best localized wave-packet is the ground state $|0,0,0\rangle$, while
5353: the three other states are excited by one quantum in either of 
5354: the normal modes $\omega_{\pm,z}$, 
5355: and are therefore significantly more extended in space.
5356: \begin{figure}
5357: \centerline{\psfig{figure=bdzf29.ps,width=15cm}}
5358: \caption{Isovalue plots (at 30\% of the maximum value) 
5359: of non-dispersive wave-packets 
5360: in the three-dimensional
5361: hydrogen atom driven by a circularly polarized microwave field. 
5362: Frequency of
5363: the driving $\omega=1/60^3$, amplitude 
5364: $F_0=0.04442$. 
5365: In the laboratory frame, the wave-packets 
5366: propagate -- without changing their shapes -- along a circular
5367: trajectory 
5368: centered around the nucleus indicated by a cross.  
5369: The cube edges measure 10000 Bohr radii. The microwave polarization plane
5370: is horizontal with the field pointing to the right. The
5371: four wave-packets shown represent the ground state wave-packet $|0,0,0 
5372: \rangle$ (top left), and the excited states $|1,0,0\rangle$ (bottom left),
5373: $|0,1,0\rangle$ (top right), and $|0,0,1\rangle$ (bottom right).
5374: Eventually, the microwave field will ionize such states, but their
5375: lifetimes are extremely long, of the order of thousands to millions
5376: of Kepler periods. 
5377: }
5378: \label{circwp3d}
5379: \end{figure}
5380: Again, as already mentioned in sections~\ref{LIN1D} and \ref{LIN3D}, these
5381: wave-packet eigenstates have finite, but extremely long life-times (several
5382: thousands to millions of Kepler orbits), due to the
5383: field induced ionization. For a detailed discussion of their decay
5384: properties see section~\ref{ION}.
5385: 
5386: As already demonstrated in 
5387: the LP case (see figs.~\ref{lin3d_3},\ref{lin3d_4}), the semiclassical
5388: prediction for the energies of the non-dispersive wave-packets
5389: is usually excellent. In order to stress the (small) differences,
5390: we plot in fig.~\ref{solit2} a part of the Floquet spectrum
5391: of the two-dimensional model 
5392: atom,
5393: i.e. quasi-energy levels versus the (scaled) microwave amplitude,
5394: after substraction of the prediction of the
5395: harmonic approximation around the stable fixed point, eq.~(\ref{enharm}), 
5396: for the 
5397: ground state wave-packet $|0,0\rangle$.
5398: The result is shown in units of the mean level spacing,
5399: estimated\footnote{This estimate follows from the local energy splitting, 
5400: $\sim n_0^{-3}$, divided by the number $n_0/2$ of photons needed to ionize the initial 
5401: atomic state by a resonant driving field.} to be roughly $2/n_0^4$.
5402: If the harmonic approximation was exact,
5403: the ground state wave-packet would be represented by 
5404: a horizontal line at zero.
5405: The actual result is not very far from that, which proves that the
5406: semiclassical method predicts the correct energy with an accuracy
5407: mostly better than the mean level spacing. The other states of the system
5408: appear as 
5409: energy levels which rapidly evolve with $F_0$, and which  
5410: exhibit extremely small
5411: avoided crossings -- hence extremely small couplings -- with the wave-packet.
5412: In the vicinity of such avoided
5413: crossings, the energy levels are perturbed, 
5414: the diabatic wave-functions mix (the wave-packet eigenstates 
5415: get distorted),
5416: and, typically, the lifetime of the
5417: state decreases (induced by the coupling to the closest
5418: Floquet state \cite{abuth,abu95a}, typically much less resistant against ionization, as we shall
5419: discuss in detail in sec.~\ref{ION}).
5420: 
5421: Thus, strictly
5422: speaking, when we speak of non-dispersive wave-packets as specific
5423: Floquet states, we really have in mind a generic situation, {\em far} from
5424: any avoided crossing. In particular, the examples of wave-packet states
5425: shown in the figures above correspond to such situations.
5426: \begin{figure}
5427: \centerline{\psfig{figure=bdzf30.ps,width=12cm}}
5428: \caption{Spectrum of the two-dimensional hydrogen atom in
5429: a circularly polarized microwave field 
5430: of frequency $\omega=1/(60.5)^{3}$, 
5431: as a function of the
5432: scaled microwave amplitude $F_0=F\omega^{-4/3}$. 
5433: In order to test the accuracy of the harmonic
5434: prediction, 
5435: we substract the semiclassical energy for the ground
5436: state $|0,0\rangle$ wave-packet, eq.~(\protect{\ref{enharm}}), from the result
5437: of the exact numerical diagonalization, and rescale the energy axis in units 
5438: of the mean level spacing. The
5439: almost horizontal line slightly above zero corresponds to the
5440: non-dispersive ground state wave-packet, which typically undergoes small
5441: avoided crossings with other Floquet states. A relatively large 
5442: avoided crossing
5443: occurs when two wave-packet-like states meet, as here 
5444: happens around $F_0\simeq 0.036$.
5445: The other (``colliding'') 
5446: state is the excited wave-packet $|1,3\rangle$. The dashed
5447: line indicates the harmonic prediction, eq.~(\ref{enharm}), for this state, 
5448: which is obviously
5449: less accurate. Note, however, that the slope is correctly predicted.
5450: }
5451: \label{solit2}
5452: \end{figure}
5453: The observed accuracy of the semiclassical approximation has 
5454: important practical consequences:
5455: in order to obtain the ``exact'' wave-packets numerically, 
5456: we do not need many
5457: eigenvalues for a given set of parameters. Using the
5458: Lanczos algorithm for the partial diagonalization of a matrix, it is enough to extract
5459: few (say five) eigenvalues only, 
5460: centered on 
5461: the semiclassical prediction. The accuracy of the latter
5462: (a fraction of the mean level spacing) is
5463: sufficient for a clear identification of the appropriate
5464: quantum eigenvalue.
5465:  Actually, in a real diagonalization of the Floquet Hamiltonian, we
5466: identify the wave-packet states by both their vicinity 
5467: to the semiclassical prediction
5468: for the energy, 
5469: and the large (modulus of the) slope of the
5470: level w.r.t. changes of $F_0$, induced by its large dipole
5471: moment in the rotating frame (by virtue of the Hellman-Feynman 
5472: theorem \cite{cct92}).
5473: The latter 
5474: criterion is actually also very useful for the identification of excited
5475: wave-packets in a numerically exact spectrum.
5476: The state $|1,3\rangle$ (in the harmonic approximation)
5477: presented in 
5478: fig.~\ref{accharm} is precisely the excited
5479: wave-packet which appears in fig.~\protect{\ref{solit2}} as a ``line''
5480: with a negative slope, meeting the $|0,0\rangle$ state in a broad avoided
5481: crossing around $F_0\simeq 0.036$. 
5482: From that figure, it is
5483: apparent that  the harmonic
5484: prediction for the energy is not excellent for the  $|1,3\rangle$ state.
5485: On the other hand, the slope of the Floquet
5486: state almost matches the slope given by the harmonic
5487: approximation, which confirms that the exact wave-function is still 
5488: well approximated by its harmonic counterpart.
5489: Note that also from the experimental point of view it is important to get accurate
5490: and simple semiclassical estimates of the energies of the non-dispersive
5491: wave-packets, since it may help in their preparation and unambiguous
5492: identification. For a more detailed discussion, see section~\ref{SPEC}.
5493: 
5494: It is interesting to compare the accuracy of 
5495: the harmonic approximation
5496: to the pendulum 
5497: description outlined previously in sec.~\ref{reso_an}. The latter
5498: results from lowest order perturbation theory in $F$.
5499: Taking the small $F$ limit we get $\omega_+,\omega_z
5500: \rightarrow
5501: \omega$, and $\omega_-\rightarrow \omega\sqrt{3F_0}$, for the 
5502: harmonic modes, see 
5503: eqs.~(\ref{omegas}-\ref{defQ}). 
5504: The latter result coincides -- as it should -- with the pendulum prediction,
5505: eqs.~(\ref{omega_harmonic},\ref{omega_harmonic_cp}). Similarly, the energy of the stable
5506: equilibrium point, eq.~(\ref{wps}), becomes at first order in $F$:
5507: \begin{equation}
5508: E_{\rm eq}= - \frac{3}{2n_0^2} + n_0^2\ F + O(F^2)
5509: \end{equation}
5510: which coincides with
5511: the energy of the center of the resonance island, see section~\ref{reso_an}.
5512: Thus, the prediction of the resonance analysis agrees with the harmonic
5513: approximation in the rotating frame. 
5514: For a more accurate estimate of the validity of both approaches, we have
5515: calculated -- for the 2D model of the atom, but similar
5516: conclusions are reached in 3D -- 
5517: the energy difference between the exact quantum result  
5518: and the prediction
5519: using a semiclassical
5520: quantization of the secular motion in the $(M,\phi)$ plane together with
5521: the Mathieu method in the $(\hat{I},\hat\theta )$ plane on the one side,
5522: and the prediction of the harmonic approximation around the fixed point,
5523: eq.~(\ref{enharm}), on the other side. 
5524: In fig.~\ref{semicirc}, we compare the results coming from both approaches.
5525: As expected, the semiclassical approach
5526: based on the Mathieu equation is clearly superior for very small microwave
5527: amplitudes, as it is ``exact" at first order in $F$. On the other hand,
5528: for the harmonic approximation to work well, the 
5529: island around the fixed point has to be sufficiently large.
5530: Since the size of the island increases as $\sqrt{F},$ the harmonic 
5531: approximation
5532: may become valid only for sufficiently large microwave amplitudes, when 
5533: there is at least
5534: one state trapped in the island. As seen, however,
5535: in fig.~\ref{semicirc}, the harmonic
5536: approximation yields a satisfactory prediction for the wave-packet energy
5537: (within few \% of the mean spacing) almost everywhere. 
5538: For increasing $n_0,$ the harmonic approximation is better and better
5539: and the Mathieu approach is superior only over a smaller and smaller range
5540: of $F_0=Fn_0^4$, 
5541: close to 0. Still, both
5542: approaches give very good predictions for the typical values of $F_0$ used in
5543: the following, say $F_0\simeq 0.03$. The spikes visible in the figure are
5544: due to the many small avoided crossings visible in fig.~\ref{solit2}.
5545: 
5546: \begin{figure}
5547: \centerline{\psfig{figure=bdzf31.eps,width=14cm,bbllx=0pt,bblly=50pt,bburx=600pt,bbury=400pt}}
5548: \caption{
5549: Difference between  the exact quantum energy of a non-dispersive
5550: wave-packet of a 2D hydrogen atom in a 
5551: circularly polarized microwave
5552: field, and two different semiclassical predictions, as a function of 
5553: the scaled microwave field amplitude, $F_0$.
5554: The thick lines are obtained from 
5555: a harmonic approximation 
5556: of the motion around
5557: the stable equilibrium point (in the rotating frame), eq.~(\ref{enharm2d}) 
5558: and the 
5559: thin lines use a quantum treatment (Mathieu approach, 
5560: sec.~\protect\ref{section_mathieu})
5561: of the motion in the 
5562: resonance island,
5563: combined with a semiclassical treatment of the secular motion. 
5564: From top to bottom
5565: $\omega=1/(30.5)^3,\ 1/(60.5)^3, \ 1/(90.5)^3$, corresponding to wave-packets
5566: associated with Rydberg states of principal quantum number
5567: $n_0=30,60,90$. The energy difference is expressed in units of the mean
5568: level spacing, estimated by $n_0^4/2$.
5569: The prediction of the Mathieu approach is
5570: consistently better for low $F_0$, and the harmonic approximation
5571:  becomes clearly superior for larger $F_0.$ For larger and larger $n_0,$
5572:  the harmonic approximation is better and better. Note that both
5573:  approximations make it possible to estimate the energy of the non-dispersive
5574:  wave-packet with an accuracy better than the mean level spacing, allowing
5575:  for its simple and unambiguous 
5576: extraction from exact numerical data.}
5577: \label{semicirc}
5578: \end{figure}
5579: 
5580: While we have shown some exemplary  wave-packets for few values of
5581:  $n_0$ and $F$
5582: only, they generally look very similar provided that
5583: \begin{itemize}
5584: \item $n_0$ is sufficiently large, say $n_0>30$. For smaller $n_0$,
5585: the wave-packet looks a bit distorted and one observes some
5586: deviations from the harmonic approximation (for a more detailed 
5587: discussion of this
5588: point see secs.~\ref{ION} and \ref{SPO}, where 
5589: ionization and spontaneous emission
5590: of the wave-packets are discussed);
5591: \item $F_0$ is sufficiently large, say $F_0> 0.001$, 
5592: such that
5593: the resonance island can support at least one state. For smaller
5594: $F_0,$ the wave-packet becomes more extended in the angular coordinate, 
5595: since less
5596:  atomic circular states are significantly coupled, see fig.~\ref{cp2df};
5597: \item $F_0$ is not too large, say smaller than $F_0\simeq0.065$. 
5598: Our numerical data suggest that the upper limit is not given
5599: by the limiting value of $q=8/9$,
5600: for which the fixed point is still stable. The limiting
5601: value appears to be rather linked to the  
5602: $1:2$ resonance between the $\omega_+$ and 
5603: $\omega_-$ modes, which occurs approx. at $F_0\simeq0.065$.
5604: \item In particular, the value $q=0.9562$ (i.e. $F_0\simeq 0.04442$),
5605: corresponding to 
5606: optimal 
5607: classical stability of the fixed point, advertised in~\cite{ibb94} as the
5608: optimal one, is by no means favored. A much broader range of microwave
5609: amplitudes is available (and equivalent as far as the ``quality'' of the
5610: wave-packet is concerned). What is much more relevant, is the presence of
5611: some accidental avoided crossings with other Floquet states.
5612: \end{itemize}
5613: Still, these are no very restrictive conditions, and we are left with 
5614: a broad range
5615: of parameters favoring the existence of nondispersive wave-packets, 
5616: a range which
5617: is experimentally fully accessible (see section~\ref{EXP} for a more
5618: elaborate discussion of experimental aspects). 
5619: \begin{figure}
5620: \centerline{\psfig{figure=bdzf32.ps,width=10cm}}
5621: \caption{Floquet eigenstate of the two-dimensional hydrogen atom in a
5622: circularly polarized microwave field. This state is 
5623: partially localized on the unstable
5624: equilibrium point (in the rotating frame), 
5625: see eqs.~(\ref{wp0})-(\ref{wpu}), for $F_0=0.057$ and $n_0=60$.
5626: This localization is of purely classical origin. The nucleus is at the center
5627: of the figure which extends over $\pm 5000$ Bohr radii. The microwave
5628: field points to the right.
5629: }
5630: \label{cp2unstb}
5631: \end{figure}
5632: 
5633: Finally, in analogy with the LP case, we may consider 
5634: Floquet states localized on the unstable fixed point associated 
5635: with the principal resonance island. 
5636: From
5637: the discussion following eq.~(\ref{wpu}), this point is located 
5638: opposite to the stable fixed point, on the other side of the nucleus. 
5639: An example of
5640: such a state is shown in fig.~\ref{cp2unstb}, for an amplitude of the microwave
5641: field that ensures that most of the nearby Floquet states ionize rather 
5642: rapidly.
5643: The eigenstate displayed in the figure  
5644: lives much longer (several thousands of Kepler periods). The
5645: localization in the vicinity of an unstable fixed point, in analogy
5646: to the LP case discussed previously, is of purely classical origin.
5647: As pointed out in \cite{delande97},
5648: such a localization must not be confused with scarring \cite{heller84} --
5649: a partial localization on an unstable periodic orbit embedded in a chaotic
5650: sea -- which disappears in the semiclassical limit \cite{heller84,bogomolny88a}.
5651: 
5652: \subsection{Rydberg states in elliptically polarized microwave fields}
5653: \label{EP}
5654: 
5655: The origin of non-dispersive wave-packets being their localization inside
5656: the resonance island (locking the frequency of the electronic motion
5657: onto the external drive) suggests that
5658:  such wave-packets are quite robust and 
5659: should  exist not only for CP and LP, but
5660: also for arbitrary elliptical polarization (EP). 
5661: 
5662: The possible existence
5663: of nondispersive wave-packets for EP  was mentioned 
5664: in \cite{lee97,Bth}, using the classical ``pulsating SOS'' approach. The
5665: method, however, 
5666: did not allow for quantitative predictions, and was restricted to
5667: elliptic polarisations very close to the CP case.
5668: However, the robustness of such wave-packets for arbitrary 
5669: EP is obvious 
5670: once the localization mechanism inside the resonance island 
5671: is well understood \cite{sacha98b,sachath}.
5672: 
5673: Let us  consider an elliptically polarized driving field of constant
5674: amplitude. With the ellipticity parameter $\alpha\in[0;1]$,
5675: \begin{equation}
5676: V=F(x\cos\omega t + \alpha y\sin\omega t)
5677: \label{v_ep}
5678: \end{equation}
5679: establishes a continuous transition between linear ($\alpha =0$) and circular 
5680: ($\alpha =1$) polarization treated in the two preceding 
5681: chapters\footnote{Note, however, that $\alpha =0$ defines a linearly polarized field 
5682: along the $x$-axis, i.e., {\em in} the plane of elliptical polarization for 
5683: $\alpha >0$. In sec.~\ref{LIN}, the polarization vector was chosen along
5684: the $z$-axis. The physics is of course the same, but the algebraic expressions
5685: are slightly different, requiring a rotation by an angle $\beta=\pi/2$
5686: around the $y$-axis.}.
5687: This general case is slightly more complicated than both  
5688: limiting cases LP and CP. 
5689: For LP microwaves (see section \ref{LIN3D}), the conservation of
5690: the angular momentum projection onto the polarization axis, $M$, makes
5691: the dynamics effectively two-dimensional. For the CP case,
5692: the transformation (\ref{hrot}) 
5693: to the frame rotating
5694: with the microwave frequency removes the explicit 
5695: time-dependence (see section \ref{CP}).
5696: None of these simplifications is
5697: possible in the general EP case, and the problem
5698: is truly three dimensional {\em and} time-dependent.
5699: 
5700: To illustrate the transition from LP to CP via EP, the 
5701: two-dimensional model of the atom is sufficient, and we shall restrain 
5702: our subsequent treatment to this computationally less involved case.
5703: The
5704: classical resonance analysis for EP microwave ionization has been
5705: described in detail in~\cite{sacha97,sacha98c}. It follows closely the lines described in detail
5706: in sections~\ref{LIN} and \ref{CP}. By expanding the
5707: perturbation, eq.~(\ref{v_ep}), in the action-angle coordinates $(I,M,\theta,\phi)$
5708: of the two-dimensional atom (see sec.~\ref{action-angle_hydrogen}), 
5709: one obtains the following secular Hamiltonian:
5710: 
5711: \begin{equation}
5712: {\cal H}_{\rm sec}={\hat P_t}-\frac{1}{2{\hat I}^2}-\omega {\hat I}+
5713: F\left[{\cal V}_1(\hat{I},M,\phi;\alpha)\cos\hat\theta-
5714: {\cal U}_1(\hat{I},M,\phi;\alpha)\sin\hat\theta\right],
5715: \label{EPr_0}
5716: \end{equation}
5717: with
5718: \begin{eqnarray}
5719:  {\cal V}_{1}(\hat{I},M,\phi;\alpha) &=& \cos\phi \left(X_1 + \alpha Y_1\right),\nonumber \\
5720:  {\cal U}_{1}(\hat{I},M,\phi;\alpha) &=& \sin\phi \left(Y_1 + \alpha X_1\right).
5721: \label{calUV}
5722: \end{eqnarray}
5723: 
5724: This can be finally rewritten as:
5725: \begin{equation}
5726: {\cal H}_{\rm sec}={\hat P_t}-\frac{1}{2{\hat I}^2}-\omega {\hat I}+
5727: F\chi_1(\hat{I},M,\phi;\alpha)\cos(\hat\theta+\delta_1).
5728: \label{EPr}
5729: \end{equation}
5730: Both, $\chi_1$ and $\delta_1$,
5731: depend on the shape and orientation of the   
5732: electronic elliptical trajectory,
5733: as well as on 
5734: $\alpha$,
5735: and are given by:
5736: \begin{equation}
5737: \chi_1(\hat{I},M,\phi;\alpha)= \sqrt{{\cal V}_1^2 + {\cal U}_1^2}, 
5738: \end{equation}
5739: and
5740: \begin{equation}
5741: \tan \delta_1(M,\phi;\alpha) = \frac{{\cal U}_1}{{\cal V}_1},
5742: \end{equation}
5743: which is once more the familiar form of a system with a resonance island 
5744: in the $(\hat{I},\hat{\theta})$ plane. The expressions obtained are in fact
5745: very similar to the ones we obtained for 
5746: the three-dimensional atom exposed to a
5747: circularly polarized microwave field, eqs.~(\ref{hamsccir2}-
5748: \ref{phasecirc}), in section \ref{reso_an}. This is actually not surprising:
5749: the relevant parameter for the transverse dynamics is the magnitude
5750: of the atomic dipole oscillating with the driving field, i.e. the
5751: scalar product of the oscillating atomic dipole with the polarization vector.
5752: The latter can be seen either as the projection of the 
5753: oscillating atomic dipole 
5754: onto the polarization plane or as the projection of the polarization vector
5755: on the plane of the atomic trajectory. If one considers a three-dimensional 
5756: hydrogen
5757: atom in a circularly polarized field, the projection of the polarization 
5758: vector
5759: onto the plane of the atomic trajectory is elliptically polarized
5760: with ellipticity $\alpha=L/M.$ This is another method to rediscover 
5761: the Hamiltonian~(\ref{EPr}) from Hamiltonian (\ref{hamsccir2}).
5762: \begin{figure}
5763: \centerline{\psfig{figure=bdzf33.ps,width=12cm}}
5764: \caption{The scaled effective perturbation $\chi_1/\hat{I}^2$
5765: driving the transverse/angular
5766: motion of a two-dimensional hydrogen atom exposed to a resonant,
5767: elliptically polarized microwave field, 
5768: plotted as a function of
5769: the scaled angular momentum, $M_0=M/\hat{I}$, and of the angle 
5770: $\phi$ between the
5771: Runge-Lenz vector and the major axis of the polarization ellipse.
5772: Left and right panels correspond to
5773: $\alpha=0.1$ and $\alpha=0.6$, respectively. Non-dispersive
5774: wave-packets are localized around the maxima of this effective potential,
5775: at $M_0=\pm 1,$ and are circularly co- and contra-rotating (with
5776: respect to the microwave field) around the nucleus 
5777: (see fig.~\protect\ref{ep2}).
5778: }
5779: \label{ep1}
5780: \end{figure}
5781: 
5782: 
5783: To obtain a semiclassical estimation of the energies of the nondispersive 
5784: wave-packets 
5785:  we proceed 
5786: precisely in the same way as for LP and CP. Since the radial motion in 
5787: $(\hat{I},\hat{\theta})$ is much faster than in the transverse/angular degree 
5788: of freedom defined by $(M,\phi)$, we first quantize the effective perturbation
5789: $\chi_1(\hat I,M,\phi;\alpha)$ driving the angular motion. 
5790: Fig.~\ref{ep1} shows 
5791: $\chi_1/\hat{I}^2$, as a function of $M_0=M/\hat{I}$ and 
5792: $\phi$, for two different values of the driving field ellipticity $\alpha$. 
5793: Note that $\chi_1$ becomes more symmetric as $\alpha\rightarrow 0$, since 
5794: this limit defines the LP case, where the dynamics cannot depend on the 
5795: rotational sense of the electronic motion around the nucleus. 
5796: The four extrema 
5797: of $\chi_1$ define the possible wave-packet eigenstates. Whereas the minima
5798: at $\phi=\pi/2,3\pi/2$ correspond to elliptic orbits of intermediate 
5799: eccentricity $0<e<1$ perpendicular to the driving field major axis, the 
5800: ($\phi$-independent)
5801: maxima at $M_0=\pm 1$ define circular orbits which co- or
5802: contra-rotate 
5803: with the driving field. 
5804: In the limit $\alpha\rightarrow 0$, the $M_0=\pm1$ maxima are associated with
5805: the same value of $\chi_1$. 
5806: Hence, the
5807: actual Floquet eigenstates appear as tunneling doublets in the 
5808: Floquet spectrum, 
5809: each member of the doublet being a superposition of 
5810: the co- and 
5811: contra-rotating wave-packets.
5812:  
5813: The quantization of the fast motion in the $(\hat{I},\hat{\theta})$ plane 
5814: is similar to the one already performed in the LP and CP cases.
5815: As we already observed (see fig.~\ref{lin3d_2}), the size 
5816: of the resonance island in the $(\hat{I},\hat{\theta})$ plane is 
5817: proportional to 
5818: $\sqrt{\chi_1}$. Correspondingly, also the localization properties of the 
5819: wave-packet along the classical trajectory improve with increasing $\chi_1$.
5820: We therefore conclude from fig.~\ref{ep1} that the eigenstates corresponding 
5821: to the minima of $\chi_1$ cannot be expected to exhibit strong 
5822: longitudinal localization,
5823: whereas the eigenstates localized along the circular orbits at the maxima of 
5824: $\chi_1$ can.
5825: 
5826: \begin{figure}
5827: \centerline{\psfig{figure=bdzf34.ps,width=10cm}}
5828: \caption{Energy levels of the two dimensional hydrogen atom driven by 
5829: a resonant,
5830: elliptically polarized microwave field of scaled amplitude $F_0=0.03$,
5831: for $n_0=21$, as a function of the 
5832: field ellipticity $\alpha$
5833: (the resonant frequency of the microwave is $\omega=1/(21.5)^3)$. 
5834: Full lines: semiclassical prediction; dotted lines: exact numerical result
5835: for the states originating from the $n_0=21$ hydrogenic manifold. The 
5836: non-dispersive wave-packets are the states originating from the upper doublet
5837: at $\alpha=0.$ The ascending (resp. descending) energy level is associated
5838: with the wave-packet co- (resp. contra-) rotating with the microwave field.
5839: }
5840: \label{ep3}
5841: \end{figure}
5842: 
5843: Fig.~\ref{ep3} compares the semiclassical prediction obtained by quantization
5844: of $\chi_1$ and ${\mathcal H}_{\rm sec}$ 
5845: (following the lines already described in section~\ref{LIN3D},
5846: for the ground state $N=0$ in the resonance island) 
5847: to the exact quasienergies (determined by numerical diagonalization of the
5848: Floquet Hamiltonian), for
5849: $\alpha$ varying from LP to CP.
5850: The agreement is excellent, with slightly 
5851: larger discrepancies between the semiclassical and the exact results for the
5852: states with smallest energy. For those states the resonance island is 
5853: very small
5854: (small $\chi_1$), what explains the discrepancy.
5855: \begin{figure}
5856: \centerline{\psfig{figure=bdzf35.ps,width=12cm}}
5857: \caption{Non-dispersive wave-packets of the two dimensional hydrogen 
5858: atom exposed to an elliptically polarized, resonant microwave field. 
5859: Scaled microwave amplitude $F_0=0.03$
5860: (for resonant 
5861: principal quantum number  $n_0=21$), 
5862: and ellipticity
5863: $\alpha=0.4$. Top row: non-dispersive wave-packet moving on a 
5864: circular orbit corotating with 
5865: the microwave field, for phases $\omega t=0,\pi/4,\pi/2$ 
5866: (from left
5867: to right). This wave-packet evolves into the eigenstate represented in 
5868: fig.~\protect\ref{cp2df}, under continuous increase of the ellipticity 
5869: to $\alpha=1$.
5870: Bottom row: 
5871: non-dispersive wave-packet launched along the same circular orbit, but 
5872: contra-rotating with  the driving field 
5873: (for the 
5874: same phases).
5875: Note that, while the co-rotating wave-packet almost preserves its shape 
5876: during the 
5877: temporal evolution,
5878: the contra-rotating one exhibits significant distortions during one 
5879: field cycle, as a direct
5880: consequence of its complicated level dynamics shown in 
5881: fig.~\protect\ref{ep3}. 
5882: Still being an exactly time-periodic Floquet
5883: eigenstate, it regains its shape after
5884: every period of the microwave.
5885: The size of each box extends over 
5886: $\pm 800$ Bohr radii, in both $x$ and $y$ directions,
5887: with the nucleus in the middle. The major axis of the polarization
5888: ellipse is along the horizontal $x$ axis and the microwave field points
5889: to the right at $t=0.$  
5890: }
5891: \label{ep2}
5892: \end{figure}
5893: 
5894: The highest lying state in fig.~\ref{ep3}, ascending with $\alpha$,
5895: is a non-dispersive wave-packet state located on the circular orbit and
5896: corotating with the EP field. It is shown in fig.~\ref{ep2} for $\alpha=0.4$.
5897: As mentioned above, the corresponding counterrotating wave-packet 
5898: is energetically
5899: degenerate with the co-rotating one for $\alpha=0$. Its energy decreases with
5900: $\alpha$ (compare fig.~\ref{ep3}). It is shown in the bottom row in
5901: fig.~\ref{ep2} for $\alpha=0.4$. While the corotating wave-packet
5902: preserves its shape for all $\alpha$ values (except at isolated avoided
5903: crossings) the counter-rotating wave-packet undergoes a series of strong
5904: avoided crossings for $\alpha >0.42$, progressively loosing its localized
5905: character. This is related to a strong decrease of the maximum 
5906: of $\chi_1$ at $M_0=-1$ with $\alpha$, clearly visible in fig.~\ref{ep1}.
5907: 
5908: 
5909: While we have discussed the 2-dimensional case only, 
5910: the CP situation (compare section \ref{CP}) indicates that for
5911: sufficiently large $\alpha$, the important resonant motion occurs in the
5912: polarization plane, being stable versus small deviations in the
5913: $z$ direction.
5914: Thus the calculations presented above are also
5915: relevant for the real three dimensional world, provided $\alpha$ is not far from 
5916: unity \cite{sachath,sacha98b}. For arbitrary $\alpha$, a full 3D
5917: analysis is required. While this is clearly more involved, the general scenario
5918: of a wave-packet anchored to a resonance island will certainly prevail.
5919: 
5920: \section{Manipulating the wave-packets}
5921: \label{MA}
5922: 
5923: We have shown in the previous sections that non-dispersive wave-packets 
5924: are genuine 
5925: solutions of the Floquet eigenvalue problem, eq.~(\ref{calhq}), under resonant 
5926: driving, for arbitrary polarization of the driving field. The semiclassical 
5927: approximation used to guide our exact numerical approach directly demonstrates 
5928: the localization of the electronic density in well defined regions of phase space,
5929: which protect the atom against ionization induced by the 
5930: external field (see, however, sec.~\ref{ION}). We have also seen that classical
5931: phase space does not only undergo structural changes under changes of the 
5932: driving field amplitude (figs.~\ref{lin1d_00}, \ref{lin1d_1},\ref{lin1d_4a}), 
5933: but also under changes of the driving field ellipticity (fig.~\ref{ep1}).
5934: Therefore, the creation of non-dispersive wave-packets can be conceived as an
5935: easy and efficient means of quantum control, which allows the 
5936: manipulation and the controlled transfer of quantum population accross phase space.
5937: In particular, one may imagine the creation of a wave-packet moving along 
5938: the polarization axis of a linearly
5939: polarized microwave field. A subsequent,
5940: smooth change through elliptical to finally
5941: circular polarization allows to transfer the electron to a circular
5942: orbit.
5943: 
5944: Adding additional static fields to the Hamiltonian (\ref{calhq}) provides us 
5945: with yet another handle to control the orientation and shape of highly excited 
5946: Rydberg trajectories, and, hence, to manipulate the localization properties 
5947: of nondispersive wave-packet eigenstates in configuration and phase space.
5948: The key point is that trapping inside the nonlinear resonance
5949: island is a robust mechanism which protects the
5950: non-dispersive wave-packet very efficiently from imperfections. This allows
5951: to adjust the wave-packet's properties at will, just by adiabatically
5952: changing the properties of the island itself. Moreover, when the strength of 
5953: the external
5954: perturbation increases, chaos generically invades a large part of 
5955: classical phase space, but the resonance islands most often survives. 
5956: The reason is that the phase locking phenomenon
5957: introduces various time scales in the system, which have different
5958: orders of magnitude.
5959: That makes the system quasi-integrable (for example
5960: through some adiabatic approximation {\em \`a la Born-Oppenheimer}) and
5961: -- locally -- more resistant to chaos. 
5962: 
5963: Hereafter, we discuss two possible alternatives of manipulating the 
5964: wave-packets.
5965: One
5966: is realized by adding a static
5967: electric field to the 
5968: LP microwave drive \cite{sacha98a}.
5969: Alternatively, the addition of 
5970: a static magnetic field to CP driving enhances the region of classical 
5971: stability, and extends the range of applicability of the 
5972: harmonic approximation \cite{lee97,farrelly95a,farrelly95,brunello96,cerjan97,lee95,brunello97}. 
5973: 
5974: \subsection{Rydberg states in linearly polarized microwave and static electric
5975: fields}
5976: \label{staticmw}
5977: \label{LINF}
5978: 
5979: Let us first consider a Rydberg electron driven by a resonant, linearly
5980: polarized microwave, in the presence of a static electric field.
5981: We already realized 
5982: (see the discussion in sec.~\ref{LIN3D}) that 
5983: the classical 3D motion of the driven Rydberg electron is angularly unstable
5984: in a LP microwave field. It turns out, however, 
5985: that a stabilization
5986: of the angular motion is possible by the addition of a
5987: static
5988: electric field $F_s$ parallel to the microwave polarization axis
5989: \cite{sacha98a,leopold86,leopold87}.
5990: The corresponding Hamiltonian reads:
5991: \begin{equation}
5992: H=\frac{p_x^2+p_y^2+p_z^2}{2}-\frac{1}{r}+Fz\cos\omega t +F_s z,
5993: \label{hcart}
5994: \end{equation}
5995: which we examine 
5996: in the vicinity of the $s=1$ resonance. As in sec.~\ref{LIN3D}, the 
5997: angular momentum projection $M$ on the $z$ axis remains a constant of 
5998: motion, and we shall assume $M=0$ in the following.
5999: Compared to the situation of a pure microwave field, there is an
6000: additional time scale, directly related to the static field.
6001: Indeed, in the presence of a perturbative static field alone, it is known 
6002: that the Coulomb degeneracy of the hydrogenic energy levels (in $L$)
6003: is lifted. The resulting eigenstates are combinations of the $n_0$ substates
6004: of the $n_0$ manifold (for $0 leq L \leq n_0-1$)\footnote{These states 
6005: are called ``parabolic" states, since the
6006: eigenfunctions are separable in parabolic coordinates \protect\cite{landau2}.}.
6007: The associated energy 
6008: levels are equally spaced by a quantity proportional to $F_s$ 
6009: ($3n_0F_s$ in atomic units). Classically, the trajectories are no longer
6010: closed but rather Kepler ellipses which slowly librate around the static 
6011: field axis,
6012: periodically changing their shapes, at a (small) frequency  
6013: $3n_0F_s\ll \omega_{\rm Kepler}=1/n_0^3.$
6014: Thus, the new time scale associated with the static electric field
6015: is of the order of $1/F_{s,0}$ Kepler periods, where
6016: \begin{equation}
6017: F_{s,0} = F_s n_0^4
6018: \end{equation}
6019: is the scaled static field. This is to be compared to the time scales $1/F_0$
6020: and $1/\sqrt{F_0}$, which characterize the secular time evolution in the
6021: $(L,\psi)$ and $(\hat{I},\hat{\theta})$ coordinates, respectively 
6022: (see discussion in section~\ref{LIN3D}), in the
6023: presence of the microwave field alone. To achieve confinement of the
6024: electronic trajectory in the close vicinity of the field polarization axis, we
6025: need $1/F_0\simeq 1/F_{s,0}$, with both, $F_s$ and $F$ small enough 
6026: to be treated at first order.
6027: 
6028: If we now consider the $s=1$ resonance, we deduce the secular Hamiltonian
6029: by keeping only the term which does not vanish after
6030: averaging over one Kepler period. For the microwave field, this 
6031: term was already identified in eq.~(\ref{hamscfin}).
6032: For the static field, only the static Fourier component of the atomic dipole,
6033: eqs.~(\ref{x0},\ref{y0}),
6034: has a non-vanishing average over one period.
6035: Altogether, this finally leads to
6036: \begin{equation}
6037: {\mathcal H}_{\rm sec}=\hat P_t -\frac{1}{2\hat{I}^2} -\omega \hat{I}
6038: + F_s X_0(\hat{I},L) \cos \psi  
6039:  + F \chi_1(\hat{I},L,\psi) \cos(\hat \theta+\delta_1).
6040: \end{equation}
6041: Since the last two terms of this Hamiltonian depend differently on
6042: $\hat{\theta}$, it is no more possible, as it was in the pure LP case (see
6043: section \ref{LIN3D}), to perform the quantization of the slow LP motion
6044: first. Only the secular approximation \cite{cct92} which consists in
6045: quantizing first the fast variables $(\hat{I},\hat{\theta})$, and subsequently
6046: the slow variables $(L,\psi)$, remains an option for the general
6047: treatment. However, since we are essentially interested in the wave-packet
6048: eigenstate with optimal localization properties, we shall focus on the ground
6049: state within a sufficiently large resonance island induced by a microwave
6050: field of an appropriate strength. This motivates the harmonic expansion of the
6051: secular Hamiltonian around the stable fixed point at 
6052: \begin{equation}
6053: \hat{I}=\hat{I}_1=\omega^{-1/3},\ \ \ \ \hat\theta = -\delta_1.
6054: \end{equation}
6055: with the characteristic
6056: frequency, see eq.~(\ref{omega_harmonic}):
6057: \begin{equation}
6058: \Omega(\hat{I}_1L,\psi) = \frac{\sqrt{3F\chi_1(\hat{I}_1,L,\psi)}}{\hat{I}_1^2}.
6059: \end{equation} 
6060: Explicit evaluation of the ground state energy of the locally harmonic
6061: potential yields 
6062: the effective Hamiltonian for the slow motion in 
6063: the $(L,\psi)$ plane:
6064: \begin{equation}
6065: H_{\rm eff}= -\frac{\Omega(L,\psi)}{2} +F\chi_1(L,\psi)
6066: +F_sX_0(L) \cos \psi ,
6067: \label{slow}
6068: \end{equation}
6069: where all quantities are evaluated at $\hat{I}=\hat{I_1}.$
6070: For the determination of the angular localization properties of the wave-packet
6071: it is now sufficient to inspect the extrema of $H_{\rm eff}$.
6072: 
6073: \begin{figure}
6074: \centerline{\hbox{
6075: \psfig{figure=bdzf36.ps,width=12cm}}}
6076: \caption{Contours of the effective Hamiltonian,
6077: eq.~(\protect{\ref{slow}}), in the
6078: $(L_0,\psi)$ plane (with $L_0=L/n_0$ the scaled angular momentum,
6079: and $\psi$ the angle between the major axis of the elliptical trajectory
6080: and the field axis). 
6081: The potential surface generates the slow 
6082: evolution
6083: of the angular coordinates of the 
6084: Kepler trajectory of a Rydberg electron exposed to collinear,
6085: static and resonant microwave electric fields.  
6086: Initial atomic principal quantum number $n_0=60$; 
6087: scaled microwave amplitude $F_0=Fn_0^4=0.03$, and scaled static field amplitude
6088: $F_{s,0}=0.12F_0<F_{s,c}$ (a),
6089: $F_{s,0}=0.25F_0>F_{s,c}$ (b), with $F_{s,c}$ the critical static field 
6090: amplitude defined in eq.~(\protect\ref{ec}). 
6091: The lighter the background, the higher
6092: the effective energy. 
6093: The contours are plotted at the semiclassical energies which 
6094: quantize $H_{\rm eff}$
6095: according to eq.(\ref{ebkp}), and thus represent the
6096: 60 eigenenergies shown in fig.~\protect\ref{f2}.
6097: Observe the motion of the stable island along the
6098: $\psi=\pi$ line (corresponding to the energetically highest 
6099: state in the manifold), under changes of $F_s$.
6100:  }
6101: \label{f1}
6102: \end{figure}
6103: For $F_{s,0}=0$, 
6104: we recover the pure LP case with a maximum along the line $L_0=1$ 
6105: (circular state), 
6106: and a minimum at $L_0=0$, $\psi=\pi/2$ (see fig.~\ref{lin3d_1}), 
6107: corresponding to 
6108: a straight line orbit perpendicular to the field. 
6109: For increasing
6110: $F_{s,0}$, the maximum moves towards lower values of $L_0$, and contracts
6111: in $\psi$, whereas the minimum approaches $\psi=0$ for constant $L_0=0$, 
6112: see fig.~\ref{f1}. It is easy to show that there exists
6113: a critical value $F_{s,c}$ of the static field, depending on $\hat{I}_1$,
6114: \be
6115: F_{s,c}=\frac{2}{3}\left\vert F_0J'_1(1)
6116: -\frac{\sqrt{3F_0J'_1(1)}}{4\hat{I}_1}\right\vert
6117:    \simeq 0.217F_0 -0.164\frac{\sqrt{F_0}}{\hat{I}_1},
6118: \label{ec}
6119: \ee
6120: above which both fixed
6121: points reach $L_0=0.$ Then, in particular, the maximum 
6122: at $L_0=0$, $\psi=\pi$, corresponds to a straight line orbit {\em parallel} 
6123: to
6124: $\vec{F}_s$. Note that in the classical limit, $\hat{I}_1\rightarrow\infty$,
6125: eq.~(\ref{ec}) recovers the purely
6126: classical value \cite{leopold87} for angular stability of the 
6127: straight line orbit 
6128: along the polarization axis, as it should.
6129: \begin{figure}
6130: \centerline{\psfig{figure=bdzf37.ps,width=10cm}}
6131: \caption{Semiclassical energy levels of the resonantly driven manifold 
6132: $n_0=60$ of the hydrogen atom, as a function of the ratio of the scaled static 
6133: electric field strength $F_s$ to the scaled 
6134: microwave amplitude $F_0$, for fixed $F_0=Fn_0^4=0.03$. The insert shows 
6135: the scaled angular momentum $L_0=L/n_0$ of the stable fixed point 
6136: $(L_0,\psi=0)$, see fig.~\protect\ref{f1},
6137: as a function of the same variable. The corresponding trajectory evolves
6138: from a circular orbit coplanar with the polarization axis to a straight
6139: line orbit stretched along this axis, via orbits of intermediate eccentricity.
6140: For $F_{s,0}>F_{s,c}$, eq.~(\ref{ec}), the stable fixed point is stationary
6141: at $L_0=0$. The corresponding wave-packet state, localized in
6142: the vicinity of the fixed point, is the energetically 
6143: highest state in the spectrum.
6144: For $F_{s,0}>F_{s,c}$, it is a completely localized wave-packet
6145: in the three dimensional space, propagating back and forth along the
6146: polarization axis, without spreading (see fig.~\ref{linf_wp}).
6147: }
6148: \label{f2}
6149: \end{figure}
6150: Therefore, by variation of 
6151: $F_{s,0}\in[0,F_{s,c}]$, we are able to continuously tune 
6152: the position of the maximum in 
6153: the $(L_0,\psi)$
6154: plane.
6155: Consequently, application of an additional static electric field gives us 
6156: control over the trajectory traced by 
6157: the wave-packet.
6158: This is further illustrated in fig.~\ref{f2}, through the 
6159: {\em semiclassical} level dynamics of the
6160: resonantly driven manifold originating from the $n_0=60$ energy shell, as a 
6161: function of $F_s$. In the limit $F_s=0$, the spectrum is equivalent to the one
6162: plotted in fig.~\ref{lin3d_3}(b). 
6163: As the static field is ramped up, the highest lying state
6164: (maximum value of the semiclassical quantum number $p$, for $F_s=0$, see 
6165: fig.~\ref{lin3d_3}, and the right column of fig.~\ref{lin3d_5}) gets stretched 
6166: along the static field direction and finally, for $F_{s,0}>F_{s,c}$, collapses onto
6167: the quasi one dimensional wave-packet eigenstate bouncing off the nucleus along
6168: a straight line Kepler trajectory. Likewise, the energetically lowest 
6169: state of the manifold (at $F_s=0$, minimum value of the semiclassical 
6170: quantum number $p$, left column of fig.~\ref{lin3d_5}) is equally rotated 
6171: towards the direction defined by $\vec{F_s}$, but stretched in the opposite 
6172: direction.
6173:  
6174: \begin{figure}
6175: \centerline{\psfig{figure=bdzf38.ps,width=12cm}}
6176: \caption{Temporal dynamics of the non-dispersive wave-packet of a
6177: three-dimensional hydrogen atom exposed to a linearly polarized,
6178: resonant microwave field, in the presence of a parallel static electric field. 
6179: $F_0=0.03$, $F_{s,0}=0.009$, $F_{s,0}/F_0=0.3$, $n_0=60$. 
6180: Driving field phases at the different stages of the wave-packet's evolution:
6181: $\omega t= 0$ (top left), $\pi/2$ (top right),
6182: $3\pi/4$ (bottom left), $\pi$ (bottom right). 
6183: It is well localized in the three dimensions of space and repeats its shape
6184: periodically (compare fig.~\protect\ref{lin1d_2} for the analogous dynamics in 
6185: the restricted 1D model, where no additional static electric field is needed). 
6186: The nucleus is at the center of the plot which extends
6187: over $\pm 8000$ Bohr radii. The microwave polarization axis and the
6188: static field are oriented along the vertical axis.
6189: }
6190:  \label{linf_wp}
6191: \end{figure}
6192: 
6193: The existence of a non-dispersive wave-packet localized in all three
6194: dimensions of space is confirmed by a pure quantum calculation, using
6195: a numerically exact  
6196: diagonalization of the Floquet Hamiltonian. Figure~\ref{linf_wp}
6197: shows the electronic density of a single Floquet eigenstate (the
6198: highest one in fig.~\ref{f2}, for $F_{s,0}/F_0=0.3$), 
6199: at various phases of the driving field.
6200: The wave-packet is clearly localized along the field axis, and
6201: propagates along a straight line classical 
6202: trajectory, repeating its shape periodically. Its dynamics precisely reproduces 
6203: the dynamics of the 1D analogue illustrated in fig.~\ref{lin1d_2}. 
6204: Once again, as for previous
6205: examples, the finite ionization rate (see section \ref{ION}) of the 
6206: 3D wave-packet is of
6207: the order of some million Kepler periods.
6208: 
6209: 
6210: 
6211: \subsection{Wave-packets in the presence of a static magnetic field}
6212: \label{MAG}
6213: 
6214: Similarly to a static electric field which may stabilize an angularly
6215: unstable wave-packet, 
6216: the properties of non-dispersive wave-packets
6217: under circularly polarized driving, in the presence of an additional 
6218: static magnetic field normal to 
6219: the polarization plane,
6220: has been studied in a series of papers
6221: \cite{lee95,farrelly95a,farrelly95,brunello96,cerjan97,lee97,brunello97}.
6222: The Hamiltonian of the system in the coordinate frame corotating with the CP
6223: field 
6224: reads (compare with eq.~(\ref{hrot}), for the pure CP case)
6225: \be
6226: H=\frac{p_x^2+p_y^2+p_z^2}{2}-\frac{1}{r}-
6227: (\omega-\omega_{\mathrm c}/2)L_z +Fx
6228: +\frac{\omega_{\mathrm c}^2}{8}(x^2+y^2),
6229: \label{fareq1}
6230: \ee
6231: where $\omega_{\mathrm c}$ is the cyclotron frequency. In atomic 
6232: units, the cyclotron frequency $\omega_{\mathrm c}=-qB/m$
6233: equals  the magnetic field value. It can be both positive or negative,
6234: depending on the direction of the magnetic field\footnote{A different
6235: convention is used (quantization axis defined by the orientation of
6236: the magnetic field) in many papers on this subject. It leads to
6237: unnecessarily complicated equations.}.
6238: This additional parameter modifies the dynamical properties which characterize 
6239: the
6240: equilibrium points, the analysis of which may be carried out alike
6241: the pure CP case treated in sec.~\ref{CP}. 
6242: A detailed stability analysis can be found in 
6243: \cite{lee97,farrelly95,rakovic98} and we summarize here
6244: the main results only.
6245:   
6246: Since changing the sign of $F$ in eq.~(\ref{fareq1}) is equivalent to changing
6247: the sign of $x$ 
6248: from positive to negative, we only consider the 
6249: equilibrium position at $x_{\rm eq}>0$ (compare eqs.~(\ref{wp0}-\ref{wpu})).
6250: For nonvanishing magnetic field, its position is given by
6251: \be
6252: \omega(\omega-\omega_{\mathrm c})x_{\rm eq}-\frac{1}{x_{\rm eq}^2}-F=0.
6253: \label{fareq2}
6254: \ee
6255: Redefining the dimensionless parameter $q$ (see eq.~(\ref{q})) via 
6256: \be
6257: q=\frac{1}{\omega(\omega-\omega_{\mathrm c})x_{\rm eq}^3},
6258: \label{mq}
6259: \ee
6260: we obtain for the
6261: microwave amplitude 
6262: \be 
6263: F=[\omega(\omega-\omega_{\mathrm c})]^{2/3}(1-q)/q^{1/3},
6264: \label{fm}
6265: \ee
6266: and  
6267: \be 
6268: E_{\mathrm eq}=[\omega(\omega-\omega_{\mathrm c})]^{1/3}(1-4q)/2q^{2/3}
6269: \label{me}
6270: \ee
6271: for the equilibrium energy.
6272: 
6273: Harmonic expansion of
6274: eq.~(\ref{fareq1}) around 
6275: the equilibrium point $(x_{\rm eq},y_{\rm eq}=0,z_{\rm eq}=0)$ allows for 
6276: a linear stability analysis
6277: in its vicinity.
6278: Alike the pure CP case,
6279: the $z$ motion decouples from the motion in
6280: the $(x,y)$ plane. For the latter, we recover 
6281: the generic harmonic Hamiltonian 
6282: discussed
6283: in section \ref{CP}, eq.~(\ref{rotaibb}), provided we substitute
6284: \be
6285: \tilde{\omega}=\omega-\omega_{\mathrm c}/2.
6286: \label{om}
6287: \ee
6288: 
6289: When expanded at second order around the equilibrium point, the
6290: Hamiltonian (\ref{fareq1}) takes the standard form
6291: of a rotating anisotropic oscillator, eq.~(\ref{rotaibb}), with
6292: $\tilde{\omega}$ replacing $\omega$, and with the stability parameters:
6293: \bea
6294: a&=&\frac{1}{\tilde{\omega}^2}\left(\frac{\omega_{\mathrm c}^2}{4}-\frac{2}
6295: {x_{\rm eq}^3}
6296: \right), \nonumber \\
6297: b&=&\frac{1}{\tilde{\omega}^2}\left(\frac{\omega_{\mathrm c}^2}{4}+\frac{1}
6298: {x_{\rm eq}^3}
6299: \right).
6300: \label{ab_mag}
6301: \eea
6302: 
6303: The regions of 
6304: stability of the equilibrium point $(x_{\rm eq},y_{\rm eq},z_{\rm eq})$
6305: are thus obtained from the domains of stability of the
6306: 2D rotating anisotropic oscillator, given by 
6307: eqs.~(\ref{stability_region_1},\ref{stability_region_2}). They
6308: are visualized in terms of the 
6309: physical parameters $F$ and $\omega_{\mathrm c}$,
6310: (using the standard scaled electric field $F_0=Fn_0^4=F\omega^{-4/3}$)
6311: in fig.~\ref{farfig1}, with the black region corresponding
6312: to eq.~(\ref{stability_region_1}), and the grey region to eq.~(\ref{stability_region_2}).
6313: Observe that the presence of the magnetic field tends to enlarge the 
6314: region
6315: of stability in parameter space; 
6316: for $\omega_{\mathrm c}=0$ (pure CP case, no magnetic field) the
6317: stability region is quite tiny, in comparison to large values of 
6318: $|\omega_{\mathrm c}|$
6319: \footnote{As long as we are interested in long-lived wave-packets,
6320: the region of small $F_0$ is of interest only. At higher $F_0$ and for
6321: $\omega_{\mathrm c}<\omega$, the
6322: strong driving field will ionize the atom rather fast -- see section \ref{ION}.
6323:  This makes the gray
6324: region $F_0>0.1$ of little practical
6325: interest.}.
6326: On the other hand, the stability diagram does not provide us any detailed 
6327: information on the actual size of the resonance island surrounding the 
6328: equilibrium point.
6329: However, 
6330: it is precisely the size of the resonance island which is crucial for 
6331: anchoring
6332: nondispersive quantum wave-packets close to the classical equilibrium point
6333: (see sec.~\ref{sapp}). 
6334: 
6335: An alternative approach to characterize the stability properties of
6336: the classical motion near $(x_{\rm eq},y_{\rm eq},z_{\rm eq})$ has been 
6337: advertized in 
6338: \cite{lee97,farrelly95a,farrelly95,brunello96,cerjan97,lee95,brunello97}: the 
6339: concept of zero velocity surfaces (ZVS).
6340: \begin{figure}
6341: \centerline{\psfig{figure=bdzf39.eps,width=10cm,angle=-90}}
6342: \caption{(Shaded) Regions of stability of the equilibrium point 
6343: $(x_{\rm eq},y_{\rm eq},z_{\rm eq})$ for circularly polarized 
6344: driving (amplitude $F$, 
6345: frequency $\omega$) of a hydrogen atom, in the presence of a magnetic
6346: field (corresponding cyclotron frequency $\omega_{\mathrm c}$). The black and grey regions correspond 
6347: to the two regions of stability, described by 
6348: eqs.~(\protect{\ref{stability_region_1}}) and (\protect{\ref{stability_region_2}}),
6349: respectively.
6350: }
6351: \label{farfig1}
6352: \end{figure}
6353: In order to construct a ZVS, 
6354: the Hamilton function is expressed in terms of velocities  
6355: rather than canonical momenta. For the harmonic Hamiltonian
6356: (\ref{rotaibb}), the calculation yields
6357: \begin{equation}
6358: H=\frac{v_x^2+v_y^2}{2}+\frac{\omega^2}{2}\left [(a-1)x^2+ (b-1)y^2
6359: \right ].
6360: \label{ZVS1}
6361: \end{equation}
6362: Thus, the ``kinetic energy'' becomes a positive function of velocities, and one
6363: can define the ZVS as
6364: \begin{equation}
6365: S=H-\frac{v_x^2+v_y^2}{2},
6366: \label{ZVS2}
6367: \end{equation}
6368: the generalization of an effective potential for interactions which mix 
6369: position and momentum coordinates.
6370: Note that, when the 
6371: velocities coincide with the canonical momenta, $S$ is nothing 
6372: but the potential energy surface. We prefer to denote it $S$ instead of $V$,
6373: to stress the difference. As discussed in detail in \cite{lee97}, a ZVS
6374: may be used to locate the equilibrium points. However, 
6375: their stability properties are not obvious 
6376: ({\em contrary} to the potential surface, where 
6377: minima define stable
6378: fixed points, while maxima and saddle points are unstable). For a ZVS, saddles
6379: are also unstable, but maxima may either be stable {\em or} 
6380: unstable.
6381: For example, the first stability region, eq.~(\ref{stability_region_1}), 
6382: of the rotating 2D anistropic
6383: Hamiltonian is associated with
6384: a stable minimum of the ZVS. The second region of stability,
6385: eq.~(\ref{stability_region_2}), corresponds to $a,b\leq 1$ and thus
6386: to a maximum of the ZVS. However, the ZVS does not show
6387: any qualitative change whether $(a-b)^2+8(a+b)$ is positive or negative,
6388: i.e. whether the equilibrium point is stable or unstable.
6389: Thus, a ZVS is clearly inappropriate, or at least potentially
6390: dangerous, for the discussion of the classical motion close to
6391: equilibrium.
6392: As a matter of fact, this difficulty with the ZVS is crucial
6393: in our case, even for the pure CP case, without additional magnetic field.
6394: Indeed, the ZVS
6395: becomes
6396: \begin{equation}
6397: S=\left(-\frac{2q+1}{2}x^2+\frac{q-1}{2}y^2\right) \omega^2,
6398: \label{ZVS3}
6399: \end{equation}
6400: where we use the single parameter $q$ to parametrize $S$. At $q=1$
6401: (i.e. $F=0$, see eq.~(\ref{fm})), the equilibrium point 
6402: turns from a saddle
6403: (for $q>1$) into a maximum. Consequently, the ZVS correctly reflects 
6404: the change of the equilibrium
6405: point from unstable ($q>1$) to stable ($q<1$).
6406: However, for any  $q\in(0,1)$, the equilibrium remains a maximum of the ZVS,
6407: which completely misses the change of stability at $q=8/9$.
6408: Thus, the very same maximum
6409: may change its stability (which fundamentally affects the classical 
6410: motion in its
6411: vicinity) without being noticed by inspection of the 
6412: ZVS. The latter
6413: evolves very smoothly around $q=8/9$. Thus, the ZVS contours provide
6414: {\em no} information on the nature of the classical 
6415: motion in the vicinity of the
6416: equilibrium point, in {\em disaccord} with   
6417: \cite{lee97,cerjan97,farrelly95a,farrelly95,brunello96,lee95,brunello97}.
6418: Similarly, the isovalue contours of the ZVS (which are ellipses in the
6419: harmonic approximation) have no relation
6420: with the isovalue  contours of the ground-state wave-packet
6421: localized
6422: around the equilibrium point (these contours
6423: are also ellipses in the harmonic approximation where the
6424: wave-packet is a Gaussian),
6425: contrary to what is stated in \cite{lee97,farrelly95a}.
6426: For example, the aspect ratio (major axis/minor axis)
6427: of the ZVS contour lines is $\sqrt{(2q+1)/(1-q)}$ which varies smoothly 
6428: around $q=8/9$, while the aspect ratio of the isocontours 
6429: of the ground state wave-packet diverges when $q\to 8/9.$ 
6430: \footnote{While this argument has been presented here
6431: for the simplest case of the harmonic
6432: oscillator hamiltonian (\ref{rotaibb}), it carries over to 
6433: the full, nonlinear model,
6434: eq.~(\ref{fareq1}).}
6435: 
6436: Nonwithstanding, a ZVS may be used for other purposes \cite{lee95}, e.g., 
6437: to show the existence of an ionization threshold 
6438: for the Hamiltonian (\ref{fareq1}), when $\omega_{\mathrm c}>\omega$
6439: (area coded in black in fig.~\ref{farfig1}). 
6440: Clearly, due to the parabolic confinement in the 
6441: $x-y$-plane, ionization is only possible along the 
6442: $z$ direction.
6443: The threshold  is given by
6444: \cite{lee95}
6445: \be
6446: E_{\mathrm ion}=F^2/2\omega(\omega-\omega_{\mathrm c}),
6447: \label{thresh_m}
6448: \ee
6449: which 
6450: lies above the equilibrium energy $E_{\rm eq}$. Thus, for parameters 
6451: in that region,
6452: the electron -- initially placed close to the stable fixed point --
6453: cannot ionize. One may expect, therefore, that wave-packets built
6454: around the equilibrium point for $\omega_{\mathrm c}>\omega$ lead to
6455: {\em discrete} Floquet states. In other cases, e.g., for pure
6456: CP driving, non-dispersive wave-packets are rather represented by 
6457: long-living resonances (see section \ref{ION}).
6458: 
6459: Finally, it has been often argued  
6460: \cite{lee97,farrelly95a,farrelly95,brunello96,cerjan97,lee95,brunello97}
6461: that the presence of the magnetic field is absolutely necessary for the
6462: construction of non-dispersive wave-packets. The authors consider 
6463: non-dispersive
6464: wave-packets 
6465: as equivalent 
6466: to Gaussian shaped wave-functions (using
6467: equivalently the notion of coherent states). Then it is vital that the
6468: motion in the vicinity of the fixed point is locally harmonic within a
6469: region of size $\hbar$. This leads the authors to conclude that
6470: non-dispersive wave-packets may not exist for the pure CP case except in the
6471: extreme semiclassical regime. 
6472: As opposed to that, the diamagnetic term in
6473: eq.~(\ref{fareq1}) gives a stronger weight to 
6474: the harmonic term, which is the basis of the above claim\footnote{Note
6475:  that the non-harmonic terms, being entirely due
6476: to the Coulomb field, are not removed or decreased by the addition
6477: of a magnetic field. They are just hidden by a larger harmonic term.}.
6478: From our point of view, which, as already stated above, 
6479: attributes 
6480: the non-dispersive character of the wave-packet 
6481: to a classical nonlinear
6482: resonance, the accuracy of the harmonic approximation 
6483: (which, anyway, always
6484: remains an approximation) is 
6485: irrelevant for the existence of
6486: non-dispersive wave-packets. 
6487: The best proof is that \cite{farrelly95a}
6488: concludes, on the
6489: basis of the validity of the harmonic approximation,
6490: that non-dispersive wave-packets should not exist for
6491: $n_0\simeq 60$ in CP field, in complete
6492: contradiction to numerically exact experiments showing their
6493: existence down to $n_0=15$ \cite{delande95}.
6494: On the other hand, it is an interesting question
6495: how good 
6496: the harmonic approximation actually is in the pure CP case. The interested
6497: reader may find a more quantitative discussion of this point in section
6498: \ref{SPO}.
6499: 
6500: 
6501: 
6502: \section{Other resonances}
6503: \label{HOR}
6504: 
6505: \subsection{General considerations}
6506: \label{ORG}
6507: 
6508: We have so far restricted our attention to nondispersive wave-packets anchored
6509: to the principal resonance of periodically driven Hamiltonian systems. In
6510: section~\ref{GM}, we already saw that {\em any} harmonic of the unperturbed 
6511: classical motion can dominate the harmonic expansion (\ref{hamfou}) of the 
6512: classical Hamilton function, provided it is resonantly driven by the external
6513: perturbation, i.e.
6514: \begin{equation}
6515: s\theta -\omega t\simeq {\mathrm const},\ s>0\ {\mathrm integer}.
6516: \label{rescond}
6517: \end{equation}
6518: This is the case when the $s^{\mathrm th}$ 
6519: harmonic of the classical internal frequency $\Omega$ is resonant with
6520: the external driving $\omega$. As $\Omega$ depends on the classical unperturbed
6521: action, the corresponding classical
6522: resonant action is defined by
6523: \begin{equation}
6524: \Omega(I_s) = \frac{\partial H_0}{\partial I}(I_s) = \frac{\omega}{s}.
6525: \label{Is}
6526: \end{equation}
6527: At this action, 
6528: the period of the classical motion is $s$ times the period 
6529: of the external drive.
6530: Precisely like in the $s=1$ case (the principal resonance), for any integer
6531: $s>1,$ Floquet eigenstates of the driven system exist which are 
6532: localized on the 
6533: associated classical stability islands in phase space. The energy of 
6534: these eigenstates 
6535: can again be estimated through the semiclassical quantization of the secular
6536: dynamics. 
6537: To do so, we start from eqs.~(\ref{hamfou},\ref{hamext}) and transform to 
6538: slowly varying variables (the ``rotating frame") defined by
6539: \begin{eqnarray}
6540: \hat{\theta}&=&\theta-\frac{\omega t}{s}, \label{trafo_s_a} \\
6541: \hat{I}&=&I, \nonumber \label{trafo_s_b} \\
6542: \hat{P_t}&=&P_t+\frac{\omega I}{s} \nonumber.
6543: \label{trafo_s_c}
6544: \end{eqnarray}
6545: The Floquet Hamiltonian in this rotating frame now reads:
6546: \begin{eqnarray}
6547: \hat{\mathcal H} =  \hat{P}_t+H_0(\hat{I})-\frac{\omega\hat{I}}{s} 
6548: + \lambda
6549: \sum_{m=-\infty}^{+\infty}V_m(\hat{I})\left\{\cos\left(m\hat{\theta}+
6550: \left(\frac{m-s}{s}\right)\omega t\right)\right\}, 
6551: \label{ham_trafo_lin1d}
6552: \end{eqnarray}
6553: which is periodic with period $\tau=sT$ with $T=2\pi/\omega$.
6554: Passing to the rotating frame 
6555: apparently destroys the $T$-periodicity of the original Hamiltonian.
6556: This, however, is of little importance, the crucial point
6557: being to keep the periodicity  $sT$ of the internal motion.
6558: If we now impose the resonance condition (\ref{rescond}), the major 
6559: contribution to the sum in eq.~(\ref{ham_trafo_lin1d}) will come 
6560: from the slowly evolving resonant term $m=s$, while the other 
6561: terms vanish upon averaging the fast variable $t$ over one period $\tau$, 
6562: leading to the secular Hamilton function
6563: \begin{equation}
6564: {\cal H}_{\rm sec}=\hat{P}_t+H_0(\hat{I})-\frac{\omega\hat{I}}{s}
6565: +\lambda V_s(\hat{I})\cos(s\hat{\theta}).
6566: \label{sec_s}
6567: \end{equation}
6568: This averaging procedure eliminates the explicit time dependence of 
6569: $\hat{\mathcal H}$, and is tantamount to restricting the validity of 
6570: ${\mathcal H}_{\rm sec}$ to the description of those classical 
6571: trajectories 
6572: which comply with eq.~(\ref{rescond})
6573: and, hence, exhibit a periodicity with period $\tau$. This will have an 
6574: unambiguous signature in the quasienergy spectrum, as we shall see further
6575: down.
6576: The structure of the secular Hamiltonian is simple and reminds us of the
6577: result for 
6578: the principal $s=1$ resonance, eq.~(\ref{hsec_ap}). However,
6579: due to the explicit appearance of the factor $s>1$ in the argument of the
6580: $\cos$ term,  a juxtaposition of $s$ resonance
6581: islands close to the resonant action, eq.~(\ref{Is}), is created.
6582: It should be emphasized that these $s$ resonance islands
6583: are actually $s$ clones of the same island. Indeed, a trajectory trapped
6584: inside a resonance island will successively visit all the islands:
6585: after one period of the drive, $\theta$ is approximately increased
6586: by $2\pi/s,$ corresponding to a translation to the next island. 
6587: At the center of the islands, there is a single, stable resonant
6588: trajectory whose period is exactly $\tau=sT.$
6589: 
6590: At lowest order in $\lambda,$ all the quantities of interest can
6591: be expanded in the vicinity of $I_s$,
6592: exactly as for the principal resonance in section~\ref{CD}.
6593: $V_s$ is consistently evaluated at the resonant action. 
6594: The pendulum approximation of the secular Hamiltonian then reads:
6595: \begin{equation}
6596: {\cal H}_{\mathrm pend} = \hat{P}_t + H_0(\hat{I_s}) 
6597: -\frac{\omega}{s}\hat{I}_s 
6598: + \frac{1}{2}
6599: H^{''}_0(\hat{I_s})\ (\hat{I}-\hat{I_s})^2 + \lambda
6600: V_s(\hat{I_s}) \cos s\hat{\theta},
6601: \label{eqpend_s}
6602: \end{equation}
6603: with the centers of the islands located at:
6604: \begin{equation}
6605: \hat{I}=\hat{I}_s=I_s,
6606: \end{equation}
6607: \begin{equation}
6608: \left\{
6609: \begin{array}{l}
6610: \displaystyle \hat\theta = k \frac{2\pi}{s},\ \ \ \ \ {\mathrm if}\ \lambda V_s(\hat{I}_s)H^{''}_0(\hat{I}_s)<0,\\
6611: \displaystyle \hat\theta = k \frac{2\pi}{s} + \frac{\pi}{s},\ \ \ \ \ {\mathrm if}\ \lambda V_s(\hat{I}_s)H^{''}_0(\hat{I}_s)>0,
6612: \end{array}
6613: \right.
6614: \end{equation}
6615: where $k$ is an integer running from 0 to $s-1.$
6616: For $H^{''}_0(\hat{I}_s)$ -- see eq.~(\ref{second_derivative}) -- positive,
6617: these are minima of the secular Hamiltonian, otherwise they are
6618: maxima. 
6619: The extension of each resonance island is, as a direct generalization of the
6620: results of section \ref{CD}:
6621: \begin{eqnarray}
6622: \Delta\hat\theta& = &\frac{2\pi}{s},\\
6623: \Delta\hat{I} & = & 4\sqrt{\left|\frac{\lambda V_s(\hat{I_s})}{ H^{''}_0(\hat{I_s})}\right|,}
6624: \end{eqnarray}
6625: and its area
6626: \begin{equation}
6627: A_s(\lambda) = \frac{16}{s} \sqrt{\left|
6628: \frac{\lambda V_s(\hat{I_s})}{ H^{''}_0(\hat{I_s})}\right|}.
6629: \label{area_s}
6630: \end{equation}
6631: Again, the dependence of $A_s(\lambda)$ on $\sqrt{|\lambda|}$ implies that even
6632: small perturbations may
6633: induce
6634: significant changes in the phase space structure, provided the perturbation
6635: is resonant with a harmonic of the unperturbed classical motion\footnote{
6636: The situation is very different in the opposite case, when $\omega$
6637: is the $s^{\mathrm th}$ SUB-harmonic \cite{marion} of the internal frequency. A resonance
6638: island may then exist but it is typically much smaller as it comes into
6639: play only at order $s$ in perturbation theory, with a
6640: size scaling as $|\lambda|^{s/2}.$}.
6641: 
6642: The construction of a non-dispersive wave-packet is simple once the
6643: $s$-resonance structure is understood: indeed, any set of initial
6644: conditions trapped in one of the $s$-resonance islands will classically remain
6645: trapped forever. Thus, a quantum wave-packet
6646: localized initially inside a resonance island is a good candidate for
6647: building a  non-dispersive wave-packet. There remains, however, a difficulty:
6648: the wave-packet can be initially placed in any of the $s$ resonance islands.
6649: After one period of the driving, it will have jumped to
6650: the next island, meaning that it will be far from its initial position.
6651: On the other hand, the Floquet theorem guarantees the existence
6652: of states which are strictly periodic with the period of the drive 
6653: (not the period of the resonant internal motion). 
6654: The solution to this
6655: difficulty is to build eigenstates which simultaneously occupy
6656: {\em all} 
6657: $s$-resonance islands, that is, which are composed
6658: of $s$ wave-packets each localized on a different resonance
6659: island. After one period of the drive, each individual wave-packet replaces
6660: the next one, resulting in globally $T$-periodic motion of this ``composite''
6661: Floquet state.
6662:  If the system has a macroscopic size (i.e. in the
6663: semiclassical limit), 
6664: individual wave-packets will appear extremely well localized
6665: and lying far from the other ones
6666: while maintaining a well-defined phase coherence with them.
6667: For $s=2$, the situation
6668: mimics a symmetric double well potential, where even and odd solutions
6669: are linear combinations of nonstationary states, each localized in either one
6670: well \cite{landau2}.
6671: 
6672: In order to get insight in the structure of the Floquet quasi-energy spectrum,
6673: it is useful to perform the semiclassical 
6674: EBK quantization of the secular Hamiltonian (\ref{sec_s}).
6675: Quantization of the motion in 
6676: $(\hat{I},\hat{\theta})$,
6677: see section~\ref{sapp}, provides 
6678: states trapped within the
6679: resonance islands (librational motion),
6680: and states localized outside 
6681: them (rotational motion). As usual, the number of
6682: trapped states is given 
6683: by the size, eq.~(\ref{area_s}) of the
6684: resonance island:
6685: \begin{equation}
6686: {\mathrm Number\ of\ trapped\ states} \simeq \frac{8}{\pi \hbar s}
6687: \sqrt{\left|\frac{\lambda V_s(\hat{I_s})}{ H^{''}_0(\hat{I_s})}\right|}.
6688: \label{number_of_trapped_states_s}
6689: \end{equation}
6690: The quantization can be performed along the contours of 
6691: any of the $s$ clones of the resonance island,
6692: giving of course the same result. However, this does {\em not}
6693: result in a $s$-degeneracy of the spectrum: indeed, the $s$ clones
6694: belong to the same torus in phase space (see above) and do
6695: not generate $s$ independent states.
6696: 
6697: If the number of trapped states is sufficiently large,
6698: the harmonic approximation to the pendulum (or secular) Hamiltonian
6699: can be used, with the frequency of the harmonic motion
6700: around the stable resonant orbit given by:
6701: \begin{equation}
6702: \omega_{\mathrm harm} = s\sqrt{|\lambda V_s(\hat{I_s})H^{''}_0(\hat{I_s})|}.
6703: \label{omega_harmonic_s}
6704: \end{equation}
6705: 
6706: In order to get the complete semiclassical Floquet spectrum,
6707: we additionally have to perform the semiclassical quantization in the 
6708: $(t,\hat{P}_t)$ plane,
6709: giving:
6710: \begin{equation}
6711: \frac{1}{2\pi}\int_0^{\tau}\hat{P}_tdt=\frac{\hat{P}_t\tau}{2\pi}
6712: =\frac{s\hat{P}_t}{\omega} = \left(
6713: j+\frac{\mu}{4}\right)\hbar,
6714: \label{wkb_k1d}
6715: \end{equation}
6716: with $j$ integer.
6717: This finally yields the semiclassical Floquet levels (in the 
6718: harmonic approximation):
6719: \begin{equation}
6720: {\cal E}_{N,j} = H_0(\hat{I_s}) - \frac{\omega}{s} \hat{I}_s + \left(j+\frac{\mu}{4}\right)\hbar \frac{\omega}{s}
6721: - \ {\mathrm sign}(H^{''}_0(\hat{I_s}))
6722: \left[ |\lambda V_s(\hat{I_s})| -
6723: \left(N+\frac{1}{2}\right) \hbar \omega_{\mathrm harm} \right],
6724: \label{spectrum_harmonic_s}
6725: \end{equation}
6726: with $N$ a non-negative integer.
6727: The wave-packet with optimum localization in the
6728: $(\hat{I},\hat{\theta})$ plane, i.e. optimum localization
6729: along the classical unperturbed orbit, is the $N=0$ state.
6730: According to eq.~(\ref{spectrum_harmonic_s}), the
6731: semiclassical quasi-energy spectrum has
6732: a periodicity $\hbar\omega/s,$ whereas the ``quantum" Floquet theory
6733: only enforces 
6734: $\hbar\omega$ periodicity.
6735: Thus, inside a Floquet zone of width $\hbar\omega,$
6736: each state appears $s$ times (for $0\leq j < s$), at energies
6737: separated by  $\hbar\omega/s.$ Note that this property is a direct
6738: consequence of the possibility of eliminating the time dependence of
6739: $\mathcal H$ in eq.~(\ref{ham_trafo_lin1d}) by averaging over $\tau$,
6740: leading to the time independent expression (\ref{sec_s}) for
6741: ${\cal H}_{\rm sec}$. 
6742: Therefore, it will 
6743: be only approximately valid for the exact quantum Floquet
6744: spectrum. In contrast, the $\hbar\omega$ periodicity holds exactly,
6745: as long as the system Hamiltonian is time-periodic.
6746: 
6747: We will now recover the $\hbar\omega/s$ 
6748: periodicity in a quantum description of our problem,
6749: which will provide us with the formulation of an eigenvalue problem 
6750: for the wave-packet eigenstates anchored to the $s$-resonance, in terms of a 
6751: Mathieu 
6752: equation. In doing so, we shall extend the 
6753: general concepts outlined in section \ref{section_mathieu} above.
6754: 
6755: Our starting point is eq.~(\ref{coupled_equations}), which we again consider
6756: in the regime where the eigenenergies $E_n$ of the 
6757: unperturbed Hamiltonian $H_0$ are locally approximately spaced 
6758: by $\hbar\Omega.$
6759: The resonance condition (\ref{rescond}) implies
6760: \begin{equation}
6761: \left. \frac{dE_n}{dn}\right|_{n=n_0}=\hbar\frac{\omega}{s},
6762: \label{deriven}
6763: \end{equation}
6764: where, again, $n_0$ is not necessarily an integer, and is related
6765: to the resonant action and its associated Maslov index through:
6766: \begin{equation}
6767: \hat{I}_s = \left(n_0+\frac{\mu}{4}\right) \hbar.
6768: \label{def_n0_s}
6769: \end{equation}
6770: When the resonance condition is met, the only efficient
6771: coupling in eq.~(\ref{coupled_equations}) connects 
6772: states with the same value of $n-sk.$ In other words, 
6773: in the secular approximation, a given state $(n,k)$ only 
6774: couples to $(n+s,k+1)$ and $(n-s,k-1).$
6775: We therefore consider a given ladder of coupled states labeled 
6776: by $j=n-sk.$ Because of the overall $\omega$ perodicity
6777: of the spectrum, changing $j$ by $s$ units (i.e., shifting all $k$-values
6778: by 1) is irrelevant, so that it is enough to consider the $s$ independent
6779: ladders $0\leq j \leq s-1.$
6780: Furthermore,
6781: in analogy to section~\ref{section_mathieu}, eq.~(\ref{matel_mat}), 
6782: we can replace the coupling matrix
6783: elements in eq.~(\ref{ham_trafo_lin1d}) by the resonantly driven Fourier
6784: coefficients of the classical motion, 
6785: \begin{equation}
6786: \langle \phi_n|V|\phi_{n+s}\rangle\simeq\langle \phi_{n-s}|V|\phi_n\rangle\simeq V_s(\hat{I}_s).
6787: \label{coupling_s}
6788: \end{equation}
6789: With these approximations, and the shorthand notation $r=n-n_0$,
6790: eq.~(\ref{coupled_equations}) takes the form of $s$ independent
6791: sets of coupled equations, 
6792: identified by the integer $j$:\footnote{For $s=1$, this equation reduces of course
6793: to eq.~(\protect\ref{tridiag}). We here use $n_0$ instead of $\hat{I}_s$; 
6794: the two quantities differ only by the Maslov index, eq.~(\protect\ref{def_n0_s}).}
6795: \begin{equation}
6796: \left[
6797: {\cal E}-E_{n_0}+\frac{n_0-j}{s}\hbar\omega-\frac{\hbar^2 r^2}{2}H_0^{''}
6798: (\hat{I}_s)\right]
6799: d_{r}=\frac{\lambda}{2}V_s[d_{r+s}+d_{r-s}],
6800: \label{coupled_s}
6801: \end{equation}
6802: where 
6803: \begin{equation}
6804: d_{r}=c_{n_0+r,\frac{n_0+r-j}{s}},
6805: \label{d_s}
6806: \end{equation}
6807: as a generalization of the notation in eq.~(\ref{diag_mat}). Again, $r$ is
6808: not necessarily an integer, but the various $r$ values involved
6809: in eq.~(\ref{coupled_s}) are equal modulo $s$.
6810: Precisely as in the case of the principal resonance, eq.~(\ref{coupled_s}) 
6811: can be
6812: mapped on its dual space expression, via eq.~(\ref{dual}):
6813: \begin{equation}
6814: \left[
6815: -\frac{\hbar^2}{2}H_0^{''}(\hat{I}_s)\frac{d^2}{d\phi^2}+E_{n_0}
6816: -\frac{n_0-j}{s}\hbar\omega+\lambda
6817: V_s\cos(s\phi)\right] f(\phi) = {\cal E} f(\phi),
6818: \label{mathieu_s}
6819: \end{equation}
6820: and identified with the Mathieu equation (\ref{mat_eq}) through
6821: \begin{eqnarray}
6822: s\phi & = & 2v, \\
6823: a & = & \frac{8}{\hbar^2s^2H_0^{''}(\hat{I}_s)}\left[{\cal E}-E_{n_0}+(n_0-j)\hbar
6824: \frac{\omega}{s}\right],\ j=0,\ldots, s-1, \nonumber \\
6825: q & = & \frac{4\lambda V_s(\hat{I}_s)}{s^2\hbar^2H_0^{''}(\hat{I}_s)}. \nonumber
6826: \label{ident_s}
6827: \end{eqnarray}
6828: The quasienergies associated with the $s$ resonance in the pendulum
6829: approximation then follow immediately as
6830: \begin{equation}
6831: {\cal E}_{\kappa,j}=E_{n_0}-(n_0-j)\hbar\frac{\omega}{s}+\frac{\hbar^2s^2}{8}
6832: H_0^{''}(\hat{I}_s)a_{\kappa}(\nu,q),
6833: \label{qu_mat_s}
6834: \end{equation}
6835: where the index $j$ runs from $0$ to $s-1$, and $\kappa$	labels the 
6836: eigenvalues of the Mathieu equation \cite{abramowitz72}.
6837: Again, the boundary condition for the solution
6838: of the Mathieu equation is incorporated via the characteristic exponent, which
6839: reads
6840: \begin{equation}
6841: \nu=-2\frac{n_0-j}{s}\ ({\mathrm mod}\ 2),\ \ \ \ \ j=0,\ldots s-1.
6842: \label{char_s}
6843: \end{equation}
6844: The structure of this quasi-energy spectrum apparently displays the
6845: expected $\hbar\omega/s$ periodicity. However, the characteristic 
6846: exponent $\nu$
6847: -- and consequently the $a_{\kappa}(\nu,q)$ eigenvalues --
6848: depend on $j$, what makes the periodicity approximate only. It is
6849: only far inside
6850: the resonance island that the $a_{\kappa}(\nu,q)$ eigenvalues are almost
6851: independent
6852: of $\nu$ and the periodicity is recovered. Deviations from this
6853: periodicity are further discussed in section~\ref{OGB}.
6854: Finally, the asymptotic expansion of the $a_{\kappa}(\nu,q)$ for large $q$,
6855: eq.~(\ref{math_asy}),
6856: gives again (compare section \ref{section_mathieu}) 
6857: the semiclassical estimate of the energy levels in the
6858: harmonic approximation, eq.~(\ref{spectrum_harmonic_s}), where the
6859: indices $\kappa$ and $N$ coincide.
6860:  
6861: 
6862: 
6863: \subsection{A simple example in 1D: the gravitational bouncer}
6864: \label{OGB}
6865: 
6866: As a first example of non-dispersive wave-packets localized on $s>1$ primary
6867: resonances, 
6868: we consider
6869: the particularly simple 1D model of
6870: a particle
6871: moving vertically in the gravitational field, and bouncing off 
6872: a periodically driven horizontal plane.
6873: This system is known as the Pustylnikov model \cite{lichtenberg83} (or,
6874: alternatively, the ``gravitational bouncer", or the ``bubblon model"
6875: \cite{holthaus95,benvenuto91,oliveira94,flatte96})  
6876: and represents a standard example of chaotic motion. Moreover,
6877: despite its simplicity, it may find possible 
6878:  applications in the dynamical manipulation of cold atoms
6879: \cite{steane95}.
6880: A gauge transformation shows its equivalence to a periodically
6881: driven particle moving in a triangular potential well, with the Hamiltonian
6882: \begin{equation}
6883: H=\frac{p^2}{2}+V(z)+\lambda z \sin(\omega t),
6884: \label{boueq1}
6885: \end{equation}
6886: where 
6887: 
6888: \begin{equation}
6889: V(z)=\left\{ 
6890: \begin{array}{ll}
6891: z & {\rm for}\ z\geq 0 , \\
6892: \infty & {\rm for}\ z<0 .
6893: \end{array}
6894: \right.
6895: \end{equation}
6896: The strength $\lambda$ of the periodic driving is proportional 
6897: to the maximum
6898: excursion 
6899: of
6900: the oscillating surface.
6901: Classically, this system is 
6902: well approximated by the standard map \cite{lichtenberg83}
6903: (with the momentum and the phase of the driving
6904:  at the moment of the bounce as variables), with kicking amplitude
6905: $K=4\lambda$. 
6906: 
6907: Apart from a (unimportant) phase shift $\pi/2$ in $\omega t$, 
6908: eq.~(\ref{boueq1}) 
6909: is of the general type of eq.~(\ref{h_gen}) 
6910: and the scenario for the creation of 
6911: non-dispersive wave-packets described in sections~\ref{CD} and
6912: \ref{QD} is applicable. 
6913: As a matter
6914: of fact, a  careful analysis of the problem 
6915: using the Mathieu approach described
6916: in sec.~\ref{section_mathieu}, as well as the semiclassical
6917: quantization of the Floquet Hamiltonian 
6918: were already outlined in \cite{holthaus95}, 
6919: where the associated Floquet eigenstates were baptized ``flotons".
6920: We recommend \cite{holthaus95,flatte96} for 
6921: a detailed discussion of the problem,
6922: reproducing here only the main results,
6923: with some minor modifications.
6924: 
6925: Solving the classical equations of motion for the unperturbed Hamiltonian
6926: is straightforward 
6927: (piecewise uniformly accelerated
6928: motion alternating with 
6929: bounces off the mirror) 
6930: and it is easy to express the unperturbed Hamiltonian and the
6931: classical internal frequency in terms of action-angle variables:
6932: \begin{eqnarray}
6933: \displaystyle H_0&=&\frac{(3\pi I)^{2/3}}{2},\\
6934: \displaystyle \Omega &=& \frac{\pi^{2/3}}{(3I)^{1/3}},
6935: \end{eqnarray}
6936: while the full, time-dependent Hamiltonian reads:
6937: \begin{equation}
6938:  H=\frac{(3\pi I)^{2/3}}{2}+\frac{\lambda\pi I^{2/3}}{(3\pi)^{1/3}}
6939: \sin(\omega t)
6940: -\frac{2\lambda (3I)^{2/3}}{\pi^{4/3}}\sin(\omega t)\sum_{n=1}^\infty
6941:  \frac{\cos(n\theta)}{n^2}.
6942: \label{boueq2}
6943: \end{equation}
6944: 
6945: Thus, the resonant action (\ref{Is}) is given by
6946: \begin{equation}
6947: \hat{I}_s = \frac{\pi^2s^3}{3\omega^3},
6948: \label{act-s-res}
6949: \end{equation}
6950: with the associated strength of the effective coupling
6951: \begin{equation}
6952: V_s = \frac{(3I)^{2/3}}{s^2\pi^{4/3}}.
6953: \end{equation}
6954: Using the framework of
6955: sections~\ref{CD}, \ref{QD} (for $s=1$),
6956: and \ref{ORG} (for $s>1$), the reader
6957: may easily compute the various  properties  of non-dispersive wave-packets
6958: in this system\footnote{There is, however, a tricky point:
6959: the Maslov index in this system
6960: is 3, with a contribution  1 coming from the outer turning point,
6961: and 2 from $z=0$, since the oscillating plane 
6962: acts as a hard wall. Hence, the relation
6963: between the principal quantum number and the action is 
6964: $I_1=n+3/4$, see eq.~(\protect\ref{WKB}).}.  
6965: An example for $s=1$
6966: is presented in fig.~\ref{boufig2},
6967: for the resonant principal quantum number $n_0=1000,$ (i.e., 
6968: $\hat{I}_1=1000.75$) where both,
6969: the (time-periodic) probability densities in 
6970: configuration and phase space 
6971: are shown. Note that such 
6972: high $n_0$ values (or even higher) correspond to typical experimental 
6973: falling heights (around 0.1 mm for $n_0=1000$) in experiments on cold atoms \cite{steane95}.
6974: Therefore, the creation of an atomic wave-packet in such an experiment would 
6975: allow to store 
6976: the atom in a quasi classical ``bouncing mode" over arbitrarily long times, 
6977: and might find some application in the field 
6978: of atom optics \cite{oberthaler99}.
6979: 
6980: \begin{figure}
6981: \centerline{\psfig{figure=bdzf40.ps,width=14cm}}
6982: \caption{A non-dispersive $s=1$ wave-packet of 
6983: the gravitational bouncer, eq.~(\protect\ref{boueq1}). The quantum number 
6984: of the resonant state is chosen as $n_0=1000$, to match typical 
6985: experimental dimensions \protect\cite{steane95}.
6986: The left
6987: column shows the time evolution of the wave-packet 
6988: for $\omega t=0,\pi/2,\pi,3\pi/2$ (from top to bottom).
6989: The right column shows the
6990: corresponding phase space (Husimi, see eq.~(\ref{husimi_def})) 
6991: representation ($z$ axis horizontal as in the left column,
6992: momentum $p$ on the vertical axis).
6993: The parameters are 
6994: $\omega\simeq 0.1487$, $\lambda=0.025$. 
6995: The periodic, nondispersive dynamics of 
6996: the wave-packet bouncing off the mirror in the gravitational field is 
6997: apparent.}
6998: \label{boufig2}
6999: \end{figure}
7000: 
7001: 
7002: \begin{figure}
7003: \centerline{\psfig{figure=bdzf41.ps,width=12cm}}
7004: \caption{Floquet eigenstates anchored to the $\omega=2\Omega$ 
7005: resonance in the gravitational bouncer (left column), in configuration space,
7006: at $\omega t=0$. Each eigenstate exhibits two wave-packets shifted by a phase
7007: $\pi$ along the classical trajectory, to abide the Floquet
7008: periodicity imposed by eq.~(\ref{flstate}). Symmetric and antisymmetric 
7009: linear combinations 
7010: of these states isolate either one 
7011: of the wave-packets, which now evolves precisely like a classical particle
7012: (right column), periodically bouncing off the wall at $z=0$.
7013: The parameters are 
7014: $\omega\simeq 0.2974$,
7015: (corresponding to 
7016: $n_0=1000$, for the $2:1$ resonance), 
7017: $\lambda=0.025$.}
7018: \label{othfig1}
7019: \end{figure}
7020: 
7021: For $s=2$, we 
7022: expect, following the general discussion in section \ref{ORG},
7023:  two quasi-energy levels, separated 
7024: by $\omega/2$, according to the semiclassical result, 
7025: eq.~(\ref{spectrum_harmonic_s}), which are
7026: {\em both} associated with the $s=2$ resonance. 
7027: As a matter of fact, such states are born out from an exact numerical 
7028: diagonalization of the Floquet Hamiltonian derived from eq.~(\ref{boueq1}). 
7029: An exemplary situation is shown in fig.~\ref{othfig1}, for $n_0=1000$ (i.e., 
7030: $\hat{I}_2=1000.75$, in eq.~(\ref{act-s-res})). The tunneling coupling between the 
7031: individual wave-packets shown in the right column of fig.~\ref{othfig1} is 
7032: given by the tunneling splitting $\Delta$ between the energies of both 
7033: associated Floquet states (left column of fig.~\ref{othfig1}) modulo 
7034: $\omega/2$. From the Mathieu approach -- 
7035: eqs.~(\ref{qu_mat_s},\ref{char_s}) --
7036: this tunneling coupling is directly related to the variations of the
7037: Mathieu eigenvalues when the characteristic exponent is changed.
7038: In the limit where the resonance island is big enough, $q\gg 1$ in
7039: eq.~(\ref{qu_mat_s}), asymptotic expressions \cite{abramowitz72}
7040: allow for the following estimate~\cite{holthaus95}:
7041: \be
7042: \Delta=\frac{8\sqrt{2}\lambda^{3/4}}{\pi\sqrt{\omega}}
7043: \exp\left(-\frac{16\pi\sqrt{\lambda}}{\omega^3}\right)=
7044: \frac{8[3(n_0+3/4)]^{1/6}\lambda^{3/4}}{\pi^{4/3}}
7045: \exp\left(-6(n_0+3/4)\sqrt{\lambda}/\pi\right),
7046: \label{otheqn}
7047: \ee
7048: which we can test with our numerically exact quantum treatment. 
7049: The result is shown in fig.~\ref{othfig2}, for two different values of $n_0$.
7050: 
7051: \begin{figure}
7052: \centerline{\psfig{figure=bdzf42.eps,width=14cm,bbllx=0pt,bblly=50pt,bburx=600pt,bbury=400pt}}
7053: \caption{Tunneling splitting between the energies (modulo
7054: $\omega/2$) of two Floquet states of the gravitational bouncer, anchored to the
7055: $s=2$ resonance island, as a function of the driving amplitude 
7056: $\lambda$.
7057: Driving frequency 
7058: $\omega\simeq 1.0825$ (top) and $\omega\simeq 0.8034$ (bottom),
7059: corresponding to resonant states $n_0=20$ and $n_0=50$, respectively. 
7060: The dashed line reproduces   
7061: the prediction of Mathieu theory \protect\cite{holthaus95}, 
7062: eq.~(\protect{\ref{otheqn}}).
7063: Observe that the latter fits the exact 
7064: numerical data only for small values of $\lambda$. At larger 
7065: $\lambda,$ small avoiding crossings between one member of the
7066: doublet and eigenstates originating from other manifolds dominate
7067: over the pure tunneling contribution and the Mathieu prediction 
7068: is not accurate.}
7069: \label{othfig2}
7070: \end{figure}
7071: 
7072: Observe the 
7073: excellent agreement for small $\lambda$, 
7074: with an almost
7075: exponential decrease of the splitting with $\sqrt{\lambda}$, as expected from
7076: eq.~(\ref{otheqn}). However,
7077: for larger values of $\lambda$ (nontheless still in the regime of predominantly
7078: regular classical motion) the splitting saturates and then starts to fluctuate
7079: in
7080: an apparently random way. While the phenomenon has not been completely 
7081: clarified so far, 
7082: we are inclined to attribute it
7083:  to tiny avoided crossings with Floquet states localized is 
7084:  some other resonance islands
7085: (for a discusion
7086:  of related phenomena see
7087: \cite{bonci98,brodier01}).
7088: Comparison of the two 
7089: panels of fig.~\ref{othfig2} additionally indicates that the region of 
7090: $\lambda$ values where 
7091: avoided crossings become important 
7092: increases in the semiclassical limit, and that eq.~(\ref{otheqn})
7093: remains valid for small $\lambda$ only. This is easily understood: in the semiclassical
7094: limit $n_0\rightarrow \infty,$ 
7095: the tunneling splitting decreases exponentially, while the density of states
7096: increases.
7097: 
7098: Finally, fig.~\ref{othfig3} shows a Floquet eigenstate anchored to the $s=11$
7099: resonance island chain. 
7100: \begin{figure}
7101: \centerline{\psfig{figure=bdzf43.ps,width=12cm}}
7102: \caption{Single Floquet state of the gravitational bouncer at phases 
7103: $\omega t=0$ (top) and $\omega t=\pi/2$ (bottom), anchored to
7104: the $11:1$ resonance island chain in classical phase space.
7105: Driving frequency $\omega=11\Omega\simeq 0.6027$, 
7106: $n_0=20000$, and
7107: $\lambda=0.025$. In the upper plot,
7108: among 11 individual
7109: wave-packets which constitute the eigenstate,
7110: five pairs (with partners moving in opposite
7111: directions) interfere at distances $z>0$ from the mirror, whereas the 
7112: 11th wave-packet bounces off the wall and interferes with itself. At
7113: later times (bottom), the 11 wave-packets are well separated in space.}
7114: \label{othfig3}
7115: \end{figure}
7116: It may
7117: be thought of as a linear combination of 11 non-dispersive
7118: wave-packets which, at a given time, may interfere with each other, or,
7119: at another time (bottom panel), are spatially well separated. 
7120: For a Helium atom bouncing off an atom mirror in the earth's gravitational
7121: field (alike the setting in \cite{steane95}),
7122: the $z$ values for such a 
7123: state reach 5 millimeters. This non-dispersive wave-packet is
7124: thus a macroscopic object composed of 11 individual components
7125: keeping a well-defined phase coherence. 
7126: 
7127: 
7128: \subsection{The $s=2$ resonance in atomic hydrogen under linearly 
7129: polarized driving}
7130: \label{ORH}
7131: 
7132: 
7133: Let us now return to the  hydrogen atom driven by LP microwaves.
7134: The highly nonlinear character of the Coulomb interaction favours 
7135: non-dispersive wave-packets anchored to the $s:1$ resonance island, since the
7136: Fourier components $V_s$ of the coupling between the atom and the
7137: microwave decay 
7138: slowly\footnote{The very same behavior characterizes the gravitational
7139: bouncer discussed in the previous section. For the bouncer the slow
7140: inverse square dependence of $V_s$ on $s$ is due to a hard 
7141: collision with the oscillating surface.
7142:  For the Coulomb problem, the singularity at the origin
7143: is even stronger.}
7144:  with $s$ -- compare eqs.~(\ref{v_1dfou},\ref{v_3d_x}).
7145: Consider first the simpler
7146: 1D model of the atom. We  discuss the 
7147: $s=2$ case only, since similar conclusions can be obtained
7148: for higher $s$ values.
7149: The left panel in fig.~\ref{orh_0} shows the classical phase space 
7150: structure (Poincar\'e surface of section) 
7151: for $F_0=0.03$,
7152:  with the $s=2$
7153: resonance completely embedded in the chaotic sea, and well separated from the 
7154: much larger principal resonance island.
7155: \begin{figure}
7156: \centerline{
7157: \psfig{figure=bdzf44.ps,width=14cm}
7158: }
7159: \caption{Left: Poincar\'e surface of section of the one-dimensional hydrogen atom 
7160: under linearly polarized driving, eq.~(\protect\ref{ham_lin1d}), for resonant
7161: driving at twice the Kepler frequency.
7162: The scaled field strength is $F_0=0.03$, 
7163: and the phase is fixed at $\omega t=0.$
7164: The $s=2$ resonance islands are 
7165: apparent, embedded in the chaotic sea, and separated from the $s=1$ resonance
7166: by invariant tori. Right: Husimi representation \protect\cite{abu95a}
7167: of a Floquet eigenstate (for $n_0=60$)
7168: anchored to the $s=2$ resonance displayed on the left.}
7169: \label{orh_0} 
7170: \end{figure}
7171: From our experience with the principal 
7172: resonance, and from the general considerations on $s:1$ resonances above, 
7173: we expect to find 
7174: Floquet eigenstates which are localized on this classical phase space
7175: structure and mimic the temporal evolution of the corresponding 
7176: classical trajectories.
7177: Indeed, the right plot in fig.~\ref{orh_0} 
7178: displays a Floquet eigenstate obtained by 
7179: ``exact" numerical diagonalization, which precisely exhibits 
7180: the desired properties.
7181: 
7182: Again, this observation has its direct counterpart in the realistic 
7183: 3D atom, where the $2:1$ resonance allows for the
7184: construction of non-dispersive wave-packets along elliptic trajectories,
7185: as we shall demonstrate now. 
7186: We proceed as for the $s=1$ case (sec.~\ref{LIN3D}): 
7187: the secular Hamiltonian is obtained by averaging the full Hamiltonian,
7188: eq.~(\ref{hamaa}), after transformation to the ``rotating frame",
7189: eq.~(\ref{trafo_s_b}),
7190: over one period $\tau=sT$ of the resonantly driven classical
7191: trajectory:
7192: \begin{equation}
7193: {\mathcal H}_{\rm sec}=\hat{P}_t-\frac{1}{2\hat I^2}
7194: -\frac{\omega \hat I}{s}
7195: +F\sqrt{1-\frac{M^2}{L^2}}\left[-X_s(\hat{I}) \cos\psi\cos\hat{\theta} + Y_s(\hat{I}) \sin\psi\sin\hat{\theta}\right],
7196: \label{hamscfin_s}
7197: \end{equation}
7198: where $X_s$ and $Y_s$ are given by eqs.~(\ref{xm},\ref{ym}). 
7199: This can be condensed into 
7200: \begin{equation}
7201: {\mathcal H}_{\rm sec}=\hat{P}_t-\frac{1}{2\hat I^2}
7202: -\frac{\omega \hat I}{s}
7203: +F\chi_s\cos(s\hat{\theta}+\delta_s) ,
7204: \label{hamscfin2}
7205: \end{equation}
7206: with
7207: \begin{eqnarray}
7208: \chi_s(\hat{I},L,\psi) & := & \sqrt{1-\frac{M^2}{L^2}}\sqrt{X_s^2\cos^2\psi+Y_s^2\sin^2\psi},
7209: \label{substc2} \\
7210: \tan\delta_s (L,\psi)& := & \frac{Y_s}{X_s}\tan\psi = \frac{J_s(se)\sqrt{1-e^2}}{eJ_s'(se)} \tan\psi.
7211: \label{substb2}
7212: \end{eqnarray}
7213: 
7214: For simplicity, we will now discuss the case $M=0,s=2.$
7215: Fig.~\ref{orh_2} shows the equipotential lines of $\chi_2$
7216: in the $(L,\psi)$ plane, calculated from 
7217: eqs.~(\ref{hamscfin2}-\ref{substb2})\footnote{Since $\chi_s$ scales globally 
7218: as $\hat{I}^2$,
7219: the equipotential lines in fig.~\protect\ref{orh_2} do not depend
7220: on $\hat{I}.$}.
7221: For a comparison with quantum data, the equipotential lines represent
7222: the values of $\chi_2$ for $n_0=42$, quantized from the WKB prescription
7223: in the $(L,\psi)$ plane, exactly as done for the principal
7224: $s=1$ resonance in section~\ref{LIN3D}.
7225:  
7226: \begin{figure}
7227: \centerline{\psfig{figure=bdzf45.ps,width=12cm,angle=-90}}
7228: \caption{Contour plot of the effective perturbation $\chi_2$, 
7229: eq.~(\protect\ref{substc2}), generating the slow evolution of the 
7230: electronic trajectory in the $(L_0=L/\hat{I},\psi)$ plane. The secular
7231: motion in this case is topologically different from that corresponding to
7232: the $s=1$ resonance (compare with fig.~\protect\ref{lin3d_1}). 
7233: In particular, 
7234: there appear new fixed points at $L_0\simeq 0.77,\psi=\pi/2,3\pi/2$ (unstable,
7235: corresponding to unstable elliptic
7236: orbits with major axis perpendicular to the polarization axis) and at
7237: $L_0\simeq 0.65,\ \psi=0,\pi$ (stable, corresponding
7238: to stable elliptic orbits with major axis parallel and antiparallel
7239: to the polarization axis). The resonance island in the $(\hat I, \hat \theta)$ 
7240: plane is quite large for the latter stable orbits, and the associated
7241: eigenstates are
7242: non-dispersive wave-packets localized both longitudinally along
7243: the orbit (locked on the microwave phase), and in the transverse direction, 
7244: see figs.~\protect\ref{orh_5}, \protect\ref{orh_6}, \protect\ref{orh_7}.}
7245: \label{orh_2} 
7246: \end{figure}
7247: 
7248: One immediately notices that the secular
7249: motion is in this case topologically different from that corresponding to
7250: the $s=1$ resonance (compare to fig.~\ref{lin3d_1}), 
7251: with the following features:
7252: \begin{itemize}
7253: \item three different types of motion coexist, with separatrices
7254: originating from the straight line orbits parallel 
7255: ($L_0=0,\psi=0,\pi)$ to the
7256: polarization axis.
7257: \item The straight line orbits perpendicular to the polarization axis
7258: ($L_0=0,\psi=\pi/2,3\pi/2)$ lie at minima -- actually zeros -- of
7259: $\chi_2.$ At lowest order, they exhibit vanishing coupling to 
7260: the external field,
7261: as for the $s=1$ resonance. Hence, the resonance island in the
7262: $(\hat{I},\hat{\theta})$ plane will be small, and the
7263: wave-packets  localized along the corresponding orbits 
7264: are not expected to exist for moderate excitations.
7265: \item The circular orbit (in the plane containing the polarization axis,
7266: $L_0=1,$ arbitrary $\psi$) also exhibits vanishing coupling (since the 
7267: circular motion is purely harmonic, no coupling is possible
7268: at $\omega=2\Omega$).
7269: \item There are ``new" fixed points at $L_0\simeq 0.77,\psi=\pi/2,3\pi/2$
7270: (unstable), and at $L_0\simeq 0.65,\psi=0,\pi$ (stable), corresponding
7271: to elliptical orbits with major axis perpendicular and parallel to the
7272: polarization axis, respectively. The latter ones correspond
7273: to {\em maxima} of $\chi_2$, and are associated with a large
7274: resonance island in the $(\hat{I},\hat{\theta})$ plane. The motion in
7275: their vicinity is strongly confined,
7276: {\it both} in the angular ($L_0,\psi$) and 
7277: in the ($\hat I,
7278: \hat \theta$) coordinates: the corresponding eigenstates can be characterized
7279: as non-dispersive wave-packets, localized both longitudinally along
7280: the orbit (locked on the microwave phase), and in the transverse direction.
7281: \end{itemize}
7282: 
7283: In order to separate quantum states localized in different regions
7284: of the $(L_0,\psi)$ space, we
7285: show in fig.~\ref{orh_3} a comparison between the semiclassical prediction and
7286: the numerically exact 
7287: Floquet energies (obtained as
7288: in section \ref{LIN3D} for the $s=1$ resonance) originating from 
7289: this manifold, with  
7290: $N=0$ in eqs.~(\ref{ebkires},\ref{ebkinres}), at $F_0=0.04$.
7291: Observe that the 16 upmost states appear in eight quasi-degenerate pairs
7292: differing by parity. 
7293: Exact degeneracy does not happen because of 
7294: tunneling effects: the lower the doublet in energy, the larger its
7295: tunneling splitting. The tunneling process 
7296: involved here is a ``transverse" tunneling in the $(L,\psi)$ plane, 
7297: where the electron jumps from the elliptic $(L_0\simeq 0.65,\psi=0)$ 
7298: Kepler 
7299: trajectory to its image under $z$-parity, the $(L_0\simeq 0.65,\psi=\pi)$ 
7300: trajectory (compare fig.~\ref{orh_2}). This tunneling process is entirely due 
7301: to the specific
7302: form of $\chi_2$,
7303: with two {\em distinct} maxima.
7304: 
7305: \begin{figure}
7306: \centerline{\psfig{figure=bdzf46.eps,width=12cm,angle=-90}}
7307: \caption{Comparison of numerically 
7308: exact quasienergies originating from the 
7309: $n_0=42$ manifold (depicted by pluses) to the semiclassical prediction 
7310: (open symbols) based
7311: on the quantization of the $s=2$ resonance island
7312: (microwave frequency $=2 \times$ Kepler frequency), for 
7313: scaled microwave field $F_0=0.04$. Circles
7314: correspond to doubly degenerate states localized in the vicinity of maxima of 
7315: $\chi_2$, around the elliptic fixed points at $(L_0\simeq 0.65,\psi=0,\pi)$
7316: in fig.~\protect\ref{orh_2}. Triangles correspond to almost 
7317: circular states in the vicinity of the stable minimum at 
7318: ($L_0=1,\psi\ \rm arbitrary)$, while diamonds correspond to states localized 
7319: around the stable minima at $L_0=0,\psi=\pi/2,3\pi/2$. 
7320: The agreement between 
7321: the semiclassical and quantum energies is very good, provided the size of the 
7322: resonance island in the $(\hat{I},\hat{\theta})$ plane 
7323: is sufficiently large
7324: (high lying states in the manifold). For low 
7325: lying states in the manifold, the discrepancies between quantum and 
7326: semiclassical 
7327: results are significant, due to the insufficient size of the island.
7328: }
7329: \label{orh_3} 
7330: \end{figure}
7331: 
7332: The energetically highest doublet in fig.~\ref{orh_3} 
7333: corresponds to states
7334: localized as close as possible 
7335: to the fixed points $L_0\simeq 0.65,\ \psi=0,\pi$.
7336: For these states (large resonance island in the ($\hat I, \hat \theta$)
7337: plane),
7338: semiclassical quantization nicely
7339: agrees with the quantum results. On the other hand, the agreement
7340: between quantum and semiclassical results progressively degrades 
7341: for lower energies,
7342: as the size of the  island
7343: in the ($\hat I, \hat \theta$) plane becomes smaller. Still, 
7344: the disagreement between
7345: semiclassical and quantum results is at most of the order of
7346: the spacing between adjacent levels\footnote{
7347: A quantum approach based on the pendulum approximation and 
7348: the Mathieu equation would give a much better prediction
7349: for such states.}.
7350: Below the energy of the unstable fixed points at 
7351: $L_0\simeq 0.77,\psi=\pi/2,3\pi/2$,
7352: there are no more pairs of classical trajectories in the $(L,\psi)$ plane
7353: corresponding to distinct classical dynamics related by $z$-parity.
7354: Hence, the doublet structure 
7355: has to disappear, as confirmed by the
7356: exact quantum results shown in fig.~\ref{orh_3}. On the other hand, there
7357: are two disconnected regions in the $(L,\psi)$ plane which can give rise to 
7358: quantized values of
7359: $\chi_2$ (and, consequently, to quasienergies) within the same quasienergy 
7360: range: the neighbourhood of
7361: the stable fixed points $(L_0=0,\psi=\pi/2,3\pi/2)$, and the region close to
7362: $L_0=1$. In the semiclassical quantization 
7363: scheme, these regions are completely decoupled  
7364: and induce two independent, 
7365: non-degenerate series of quasienergy levels. 
7366: Consequently, the complete spectrum exhibits a rather
7367: complicated structure, caused by the interleaving of these two series.
7368: 
7369: As discussed in section~\ref{ORG}, for a $s:1$ resonance, in a Floquet 
7370: zone of width
7371: $\omega=s\Omega$, there is
7372: not a single manifold of states, but rather a set
7373: of $s$ different manifolds approximately identical and
7374: separated by $\Omega.$
7375: Deviations from the exact $\omega/s$ periodicity are due
7376: to tunneling~\cite{holthaus95}. This tunneling process is however
7377: {\em completely} different from the 
7378: ``transverse" one in the $(L,\psi)$ plane described above. 
7379: It is a case of 
7380: ``longitudinal" tunneling, where the electron jumps from one location
7381: on a Kepler orbit to another, shifted along 
7382: the {\em same} orbit. This longitudinal tunneling is 
7383: similar in origin to the tunneling described in section \ref{OGB}. 
7384: It has to be stressed that it represents a  general
7385: phenomenon in the vicinity of a $s:1$ resonance 
7386: (with $s\ge2)$, due to the
7387: phase space structure in the $(\hat I, \hat \theta$) plane, see section
7388: ~\ref{ORG},
7389:  in contrast
7390: to the ``transverse" quasi-degeneracy (discussed in fig.~\ref{orh_3})
7391: due to the specific form of $\chi_2.$
7392: Inspecting the numerically exact quantum quasienergy spectrum,
7393: we indeed find the manifold 
7394: shown in fig.~\ref{orh_4} (compared to the semiclassical prediction). 
7395: Observe that the agreement between quantum and
7396: semiclassical quasienergies is similar to that observed in fig.~\ref{orh_3},
7397: except for the low lying states. Here, incidentally, the states
7398: anchored to the resonance island are strongly perturbed by another
7399: Rydberg manifold; proper identification of the individual 
7400: quantum states is very difficult in this region, and therefore no quantum 
7401: data are shown at low energies.
7402: 
7403: \begin{figure}
7404: \centerline{\psfig{figure=bdzf47.eps,width=12cm,angle=-90}}
7405: \caption{Same as fig.~\protect\ref{orh_3}, but for the mirror manifold 
7406: shifted in energy by $\omega/2$. While, for most states, the agreement with 
7407: semiclassics is of the same quality as in fig.~\protect\ref{orh_3}, no quantum
7408: data are plotted at the bottom of the manifold. Indeed, at those energies, 
7409: another Rydberg manifold strongly perturbs the spectrum due to close 
7410: accidental degeneracy. Consequently, the unambiguous identification of 
7411: individual states is very difficult.}
7412: \label{orh_4} 
7413: \end{figure}
7414: 
7415: Finally, let us consider the localization properties of the wave-functions 
7416: associated with 
7417: the upmost states of the manifolds in figs.~\ref{orh_3} and \ref{orh_4}. 
7418: These wave-functions should localize in the vicinity of 
7419: stable trajectories of period 2, i.e. they should be 
7420: strongly localized, both in angular and orbital coordinates, 
7421: along an elliptic Kepler orbit of intermediate eccentricity. 
7422: However, because of the longitudinal quasi-degeneracy, we expect the
7423: associated Floquet eigenstates to be composed of
7424: two wave-packets on the ellipse, exchanging their positions
7425: with period $T.$ 
7426: Furthermore, due to the transverse quasi-degeneracy,
7427: we should 
7428: have combinations of the elliptic orbits labeled by 
7429: $\psi=0$ and $\psi=\pi.$ Altogether, this makes four individual
7430: wave-packets represented by each Floquet state. 
7431: Due to the azimuthal symmetry of the problem around the field polarization
7432: axis, each wave-packet actually is doughnut-shaped (compare fig.~\ref{lin3d_8}
7433: for the simpler $s=1$ case).
7434: 
7435: Exact quantum calculations fully confirm this prediction. Fig.~\ref{orh_5}
7436: shows the electronic density of the upmost Floquet state in the
7437: $n_0=42$ manifold (fig.~\ref{orh_3}), averaged over one field period. 
7438: As expected,
7439: it is localized along two symmetric Kepler ellipses ($\psi=0$ and $\psi=\pi$,
7440: respectively), but
7441: longitudinally delocalized because of the time average.
7442: In fact, there are four such Floquet states displaying very similar electronic
7443: densities. These are the energetically highest doublet 
7444: in the $n_0=42$ manifold,
7445: and the upmost doublet in the ``mirror" manifold displayed in fig.~\ref{orh_4}.
7446: 
7447: \begin{figure}
7448: \centerline{\psfig{figure=bdzf48.ps,width=12cm,angle=0}}
7449: \caption{Electronic density of the upmost eigenstate of the $n_0=42$ manifold
7450: of fig.~\protect\ref{orh_3}, averaged over one microwave period. This state 
7451: presents localization along a pair of Kepler ellipses oriented along the field
7452: polarization axis. The box measures $\pm 3500$ Bohr radii in both $\rho$ and 
7453: $z$ directions, with the nucleus at the center. 
7454: The microwave polarization axis along $z$ is 
7455: parallel to the vertical 
7456: axis of the figure. The orientation and eccentricity of the ellipse
7457: are well predicted by the classical resonance analysis.}
7458: \label{orh_5} 
7459: \end{figure}
7460: 
7461: Fig.~\ref{orh_6} shows the electronic densities of these four
7462: Floquet eigenstates at phase $\omega t=0$ of the driving field:
7463: the four doughnuts  are now clearly visible,
7464: as well as the orbital and radial 
7465: localizations along the two elliptic trajectories. 
7466: Very much in the same way as for a double well potential
7467: (or for the bouncer discussed in section \ref{OGB}, compare 
7468: fig.~\ref{othfig1}),
7469: a linear combination 
7470: of these four states allows for the selection of one single doughnut, 
7471: localized along one single classical Kepler ellipse. 
7472: This wave-packet then evolves along this 
7473: trajectory without dispersion, as demonstrated in fig.~\ref{orh_7}. 
7474: \begin{figure}
7475: \centerline{\psfig{figure=bdzf49.ps,width=12cm,angle=0}}
7476: \caption{Electronic densities of the eigenstates of the upmost doublet states 
7477: (top) of the $n_0=42$ manifold of fig.~\protect\ref{orh_3}, and of their mirror
7478: states (bottom), shifted in energy by $\omega/2$ (fig.~\ref{orh_4}), 
7479: at driving field phase $\omega t=0$. 
7480: The longitudinal localization
7481: on the Kepler ellipses (similar for all states) is apparent. On each ellipse,  
7482: four different individual wave-packets (or rather, due to azimuthal symmetry,
7483: two doughnut wave-packets) 
7484: can be distinguished, 
7485: propagating along the Kepler ellipse.
7486: Notice the phase shift of $\pi$ in the temporal evolution on the two ellipses,
7487: implied by $z$-inversion. The microwave polarization axis along $z$ is 
7488: given by the
7489: vertical 
7490: axis of the figure, with the nucleus at the center of the figure.}
7491: \label{orh_6} 
7492: \end{figure}
7493: \begin{figure}
7494: \centerline{\psfig{figure=bdzf50.ps,width=12cm,angle=0}}
7495: \caption{Temporal evolution of a convenient linear combination of the four 
7496: eigenstates 
7497: of fig.~\protect\ref{orh_6}, for phases $\omega t=0$ (top left), 
7498: $\pi/2$ (top center), $\pi$ (top right), $3\pi/2$ (bottom left), $2\pi$ (bottom
7499: right) of the driving field. Clearly, a single doughnut propagating along a 
7500: single trajectory has been selected by the linear combination. This wave-packet
7501: essentially repeats its periodic motion with period $2T=4\pi/\omega$. 
7502: It slowly 
7503: disperses, because the four states it is composed of 
7504: are not exactly degenerate 
7505: (tunneling effect), and because it ionizes (see sec.~\ref{ION}). 
7506: The microwave polarization axis 
7507: along $z$ is 
7508: parallel to the
7509: vertical 
7510: axis of the figure, with the nucleus at the center of the plot.}
7511: \label{orh_7} 
7512: \end{figure}
7513: Note, however, that 
7514: this single wave-packet is {\em not} a single Floquet state, and
7515: thus does not exactly
7516: repeat itself periodically. It slowly disappears
7517: at long times, for 
7518: at least two 
7519: reasons: firstly, because of
7520: longitudinal and transverse tunneling, the phases of the four Floquet
7521: eigenstates accumulate small differences as time evolves, what induces
7522: complicated oscillations between the four possible
7523: locations of the wave-packet, and secondly, 
7524: the ionization rates of the individual
7525: Floquet states lead to ionization and loss of phase coherence,
7526: especially if the ionization rates (see sec.~\ref{ION})
7527: of the four states are not equal.
7528: 
7529: 
7530: 
7531: \section{Alternative perspectives}
7532: \label{ALTP}
7533: 
7534: 
7535: There are several known systems where 
7536: an oscillating field is used to
7537: stabilize a specific mode of motion, such as 
7538: particle accelerators \cite{lichtenberg83}, Paul traps 
7539: \cite{paul92} for ions, etc. 
7540: In these cases, the 
7541: stabilization is a completely
7542: classical phenomenon based on 
7543: the notion of nonlinear resonances. What 
7544: distinguishes our
7545: concept of non-dispersive wave-packets discussed in the preceding chapters 
7546: from those situations is
7547: the necessity to use quantum (or semi-classical) mechanics to describe a given
7548: problem, due to relatively low quantum numbers. Still, the principle of 
7549: localization 
7550: remains the same, and consists in locking the motion of the system
7551: on the external drive.
7552: However, it is not essential that the drive be provided externally, it 
7553: may well be supplied 
7554: by a (large) part of the system to the (smaller) remainder. Note that, 
7555: rather formally, also 
7556: an atom exposed to a microwave field can be understood as one large 
7557: quantum system -- a 
7558: dressed atom, see sec.~\ref{SQ} -- where the field-component 
7559: provides the drive for 
7560: the atomic part \cite{cct92}.
7561: In the present section, we shall therefore briefly recollect 
7562: a couple of related phase-locking phenomena in slightly more complicated 
7563: quantum systems, which open additional perspectives for creating
7564: non-dispersive 
7565: wave-packets in the microscopic world.
7566: 
7567: 
7568: 
7569: \subsection{Non-dispersive wave-packets in rotating molecules}
7570: \label{MOL}
7571: 
7572: A situation closely related to atomic hydrogen exposed to CP microwaves 
7573: (sec.~\ref{CP}) is met when considering the dynamics of a single, 
7574: highly
7575: excited Rydberg electron in a
7576: rotating molecule~\cite{benvenuto94}. 
7577: In \cite{ibb96}, the following  
7578: model Hamiltonian has been proposed:
7579: \be
7580: H=\frac{\vec{p}^2}{2}-\frac{1}{|\vec{r} + \vec{a}(t)|},
7581: \label{molibb0}
7582: \ee
7583: where $\vec{a}(t)$ denotes the position of the center of the Coulomb field
7584: w.r.t. the molecular center of mass, and is assumed to rotate in the 
7585: $x-y$ plane 
7586: with constant frequency $\omega$:
7587: \begin{eqnarray}
7588: {\vec a}(t) & = & \left (
7589: \begin{array}{lc}
7590: \cos\omega t & \ \ -\sin\omega t \\
7591: \sin\omega t & \ \ \cos\omega t
7592: \end{array} \right ){\vec a}.
7593: \label{rotvec}
7594: \end{eqnarray}
7595: In the rotating frame, one obtains the Hamiltonian \cite{benvenuto94} 
7596: \be
7597: H=\frac{p_x^2+p_y^2+p_z^2}{2}-\frac{1}{r} +a\omega^2 x -\omega L_z +
7598: \frac{a^2\omega^2}{2},
7599: \label{molibb}
7600: \ee
7601: which, apart from the constant term $a^2\omega^2/2$, is equivalent to the one 
7602: describing an atom driven by a CP field
7603: (compare eq.~\ref{hrot})).  Note that the
7604: role of the microwave amplitude (which can be arbitrarily tuned in the 
7605: CP problem)
7606: is taken by $a\omega^2$, i.e., a combination of molecular parameters, 
7607: what, of course, 
7608: restricts the
7609: experimental realization of 
7610: non-dispersive wave-packets in the molecule to 
7611: properly selected molecular species \cite{ibb96}. 
7612: 
7613: With the help of the stability analysis outlined in sec.~\ref{rotating_frame}, 
7614: eqs.~(\ref{wp0}-\ref{wpu}), the 
7615: equilibrium position $x_{\mathrm eq}$ 
7616: of the molecular Rydberg electron is easily estimated according to 
7617: (assuming a small value of $a$, limited by the size 
7618: of the molecular core)
7619: \be
7620: x_{\mathrm eq}\simeq ({\cal I}/{\cal J})^{2/3},
7621: \label{x0mol}
7622: \ee
7623:  with ${\cal I}$ 
7624: the molecular momentum of inertia, 
7625: ${\cal J}=\omega{\cal I}$
7626: the rotational quantum number.
7627: To optimize the angular localization of the wave-packet, 
7628: it is necessary that $x_{\mathrm eq}$ 
7629: be sufficiently large 
7630: (from sec.~\ref{RSEF}, $x_{\mathrm eq}\sim n_0^2$, where
7631: $n_0$ is the electronic principal quantum number). Thus, for given ${\cal I}$,
7632: ${\cal J}$
7633: should be 
7634: small. In 
7635: \cite{ibb96}, 
7636: a hydrogen-tritium molecule is considered, which yields 
7637: $n_0\simeq 18$ for ${\cal J}=1$. 
7638: 
7639: Note, however, that such reasoning 
7640: is {\em not} justified. 
7641: The effective Hamiltonian (\ref{molibb0})
7642: implies a {\em classical} description of 
7643: the molecular rotation (much as the {\em classical} treatment 
7644: of the periodic drive in 
7645: eq.~(\ref{h_gen}), with a well-defined phase) defined by the position vector
7646: ${\vec a}(t).$
7647: For such an approach to
7648: be valid, ${\cal J}$ must be sufficiently 
7649: large. The molecular rotation plays
7650: the role of the microwave field in the analogous CP problem,
7651: the number of rotational quanta is just equivalent to the average number of
7652: photons defining the amplitude of the (classical) coherent state 
7653: of the driving 
7654: field. Clearly, if ${\cal J}$ is too small, the  
7655: effect  of an exchange
7656: of angular momentum between the Rydberg electron and the core
7657: on 
7658: the {\em quantum state} 
7659: of the core (and, hence, on $a\omega^2$ assumed to be constant
7660:  in eq.~(\ref{molibb})) cannot be
7661: neglected and, therefore, 
7662: precludes any semiclassical treatment, see also \cite{delande98,shirley65}. 
7663: In other words, if ${\cal J}$ is too small, the number of rotational 
7664: states of the core which are coupled 
7665: via the interaction is too small to mimic a quasi-classical evolution as 
7666: suggested by eq.~(\ref{rotvec}).
7667: 
7668: Nontheless, this caveat does not completely
7669: rule out the existence of 
7670:  molecular non-dispersive wave-packets,
7671: provided a fast
7672: rotation of a core  
7673: with large momentum of inertia (to render $x_{\mathrm eq}$
7674: sufficiently large, eq.~(\ref{x0mol}),
7675: such that the electronic wave-packet gets localized 
7676: far away from the molecular core) can be 
7677: realized, as also suggested
7678: in \cite{ibb96}. 
7679: 
7680: \subsection{Driven Helium in a frozen planet configuration}
7681: \label{HE}
7682: 
7683: 
7684: In the previous examples of non-dispersive wave-packets,
7685: the key point has been the generic appearance of a nonlinear
7686: resonance 
7687: for periodically driven quantum
7688: systems whose  unperturbed dynamics is integrable.
7689: A natural question to ask is whether the concept of non-dispersive wave-packets 
7690: can be generalized to systems which exhibit mixed regular-chaotic dynamics 
7691: even in the absence of the external perturbation. In the atomic realm, 
7692: such a situation is realized for the helium atom, where 
7693: electron-electron interactions provide 
7694: an additional source of 
7695: nonlinearity.
7696: The corresponding Hamiltonian writes in atomic units
7697: \begin{equation}
7698: H_{\rm He} = \frac{{\vec p}_1^2}{2}+\frac{{\vec p}_2^2}{2}-\frac{2}{r_1}
7699: -\frac{2}{r_2}+\frac{1}{|{\vec r}_1-{\vec r}_2|}.
7700: \label{he_free}
7701: \end{equation}
7702: As a matter of fact, the classical and quantum 
7703: dynamics of the three-body Coulomb problem generated by Hamiltonian~(\ref{he_free}) 
7704: has been a largely unexplored
7705: ``terra incognita'' until very recently \cite{tanner00}, since the dimensionality 
7706: of the phase space dynamics increases from effectively two to effectively eight 
7707: dimensions when a second electron is added to the familiar Kepler problem.
7708: Furthermore, the exact quantum mechanical treatment of the helium atom 
7709: remains a formidable task since the early days of quantum mechanics, and 
7710: considerable advances could be achieved only very recently, with the 
7711: advent of modern semiclassical and group theoretical methods 
7712: \cite{wintgen93,gremaud97,gremaudth,puttner01}.
7713: Already the classical dynamics of this system exhibits 
7714: a largely chaotic phase space structure, which typically 
7715: leads to the rapid autoionization 
7716: of the associated {\em doubly excited} 
7717: quantum states of the atom. One of the major 
7718: surprises in the analysis of the three body Coulomb problem during 
7719: the last decades has therefore been the discovery of a new, highly correlated
7720: and classically globally stable electronic configuration, the 
7721: ``frozen planet'' \cite{eichmann90,richter90}. 
7722: The appeal of this configuration resides in its 
7723: counterintuitive, asymmetric character where both electrons are located on the
7724: same side of the nucleus. Furthermore, this configuration turns out to 
7725: be the most robust of all known doubly excited two-electron configurations, in the 
7726: sense that it occupies a large volume in phase space. Its stability is 
7727: due to the strong coupling of the two electrons by the $1/|\vec{r}_1-\vec{r}_2|$ 
7728: term in the Hamiltonian (\ref{he_free}), which enforces their highly correlated motion.
7729: 
7730: The frozen planet is an ideal candidate to test the prevailance of the 
7731: concept of nondispersive wave-packets in systems with intrinsically
7732: mixed dynamics.
7733: In a recent study~\cite{schlagheck98a,schlagheck99,schlagheck99b,schlagheckth} 
7734: the response of this highly correlated two-electron 
7735: configuration to a periodic force has been investigated from a classical
7736: and from a quantum mechanical point of view. 
7737: The Hamiltonian for the driven problem writes, in the length gauge,
7738: \begin{equation}
7739: H=H_{\rm He}+F\cos(\omega t)(z_1+z_2).
7740: \label{he_driven}
7741: \end{equation}
7742: Guided by the experience 
7743: on non-dispersive wave-packets in one electron Rydberg states,
7744: the driving frequency $\omega$ 
7745: was chosen near resonant with the natural frequency
7746: $\Omega_{\rm FP}\approx 0.3n_i^{-3}$ of the frozen planet, where $n_i$ denotes the 
7747: principal
7748: quantum number of the inner electron. It was found that, for a suitably chosen 
7749: driving field amplitude $F$, a nonlinear resonance between the correlated 
7750: electronic motion and the external drive can be induced in the classical 
7751: dynamics, at least for 
7752: the collinear frozen planet where the three particles (two electrons and the
7753: nucleus) are aligned 
7754: along the polarization
7755: axis of the driving field. 
7756: 
7757: However, contrary to the situation for the 
7758: driven hydrogen atom discussed in sections \ref{LIN1D} and \ref{LIN3D}, 
7759: there is a fundamental difference between the one 
7760: dimensional model of the driven three body Coulomb problem
7761: and the full 3D problem. For the one-electron system,
7762: we have seen that the classical Kepler ellipse performs
7763: a slow precession in the angular variables, though
7764: remains bounded and does not ionize. In contrast, if one permits deviations
7765: from collinearity in the driven frozen planet dynamics, it is found that
7766: the transverse
7767: direction is generally unstable and leads to rapid ionization. This transverse
7768: ionization is simply due to the fact that the external field destroys the 
7769: intricate electron electron correlation which creates the unperturbed 
7770: frozen planet. 
7771: Notwithstanding, 
7772: it has been shown that the application of an additional, weak 
7773: static electric field allows to compensate 
7774: for the transverse instability, and to establish a classically globally 
7775: stable dynamical 
7776: situation for the frozen planet. The transverse confinement 
7777: through the static field again justifies the collinear model, 
7778: and first quantum calculations performed 
7779: for this restricted model show the existence of a 
7780: wave-packet associated with the principal resonance between the frozen planet orbit and 
7781: the driving field,
7782: which faithfully traces the classical trajectory at
7783: the period of the drive. 
7784: As for driven one-electron systems, 
7785: these nondispersive two-electron 
7786: wave-packets exhibit life times of typically $10^6$ driving field periods
7787: \footnote{Again, in contrast to the driven one electron 
7788: problem, nothing guarantees that the life times obtained for the 1D model 
7789: carry over
7790: to the real 3D object. On the contrary, first results on the bare 3D Coulomb 
7791: problem 
7792: \protect\cite{schlagheckth} indicate a strong dependence 
7793: of the life times on 
7794: the dimension of the accessible configuration space.}.
7795: 
7796: Hence, there is strong evidence that a resonant external forcing allows
7797: for the creation of quantum eigenstates with a quasi-classical
7798: temporal evolution, even in the presence of strong two-particle correlations.
7799: 
7800: 
7801: \subsection{Non-dispersive wave-packets in isolated core excitation of
7802: multielectron atoms}
7803: \label{LAMBRO}
7804: 
7805: Another example of non-dispersive wave-packets in a two-component 
7806: atomic system has 
7807: recently been proposed for two-electron 
7808: atoms \cite{hanson95,zobay96,mecking98}. 
7809: The scheme
7810: uses an isolated-core excitation in which one of the electrons is transfered
7811: to a Rydberg trajectory 
7812: by a {\em short} laser pulse, forming 
7813: an initially well-localized wave-packet.
7814: A second 
7815: source {\em continuously} drives a transition between two discrete states 
7816: of the remaining atomic 
7817: core. The latter
7818: induces Rabi oscillations (or a coherent superposition) between 
7819: two 
7820: Rydberg series to which the
7821: first electron is
7822: excited. If the 
7823: Rabi 
7824: frequency (controlled by the continuous drive of the core) is matched with the
7825: Kepler frequency of the orbit of  
7826: the 
7827: outer electron,
7828: the autoionization rate of the latter
7829: may be strongly suppressed, provided the respective
7830: phases are also matched properly: 
7831: if 
7832: the electron approaches its 
7833: inner turning radius (where 
7834: the configuration-interaction between Rydberg electron and core -- 
7835: leading to autoionization --
7836: is 
7837: strongest) while 
7838: the core is in its ground state, 
7839: autoionization becomes impossible since the  
7840: configuration-interaction 
7841: does not compensate for the ionization potential of the Rydberg electron. 
7842: On the other
7843: hand, when the electron is far from the nucleus 
7844: (and electron-electron
7845: interaction is weak), the core may be in its excited state, 
7846: without ejecting the Rydberg 
7847: electron.
7848: 
7849: Consequently, autoionization is supressed for the center of the 
7850: Rydberg wave-packet.
7851: During 
7852: time evolution, however, the wave-packet spreads, 
7853: its head and
7854: its tail desynchronize with the Rabi evolution of the core, 
7855: and eventually approach   
7856: the region close
7857: to the nucleus (where configuration-interaction is most pronounced)  
7858: when the core is not
7859: in its ground state. Then these parts of the wave-packet autoionize, 
7860: and the remaining Rydberg population 
7861: is reshaped into a localized wave-packet, since the spreading tails have been 
7862: chopped off.
7863: Hence, these wave-packets exhibit a rather rapid ``melting'' (on a time scale
7864: of at most some hundred Kepler periods) --
7865: to be compared 
7866: to hundreds of thousands or even more Kepler cycles performed 
7867: by non-dispersive 
7868: wave-packets in microwave driven hydrogen atoms studied above (which also
7869: ionize, however {\em very slowly}, see sec.~\ref{ION}).
7870: 
7871: 
7872: The present scenario is in some sense reminiscent of the one 
7873:  in sec.~\ref{MOL}, with a (quantum) two-level core replacing
7874: the rotating molecular core.
7875: As mentioned above, a two-level system alone can only exchange one quantum
7876: with the the outer electron and thus cannot provide an
7877:  exact phase locking mechanism for the highly excited Rydberg electron.
7878: However, the two-level core
7879: is here {\em driven} by an external electromagnetic field and consequently
7880: gains an additional degree of freedom which can be used for the
7881: phase locking mechanism. The drawback is that this phase locking
7882: implies losses (through autoionization). Nevertheless,
7883: the quasi-classical evolution over $\sim 100$ Kepler cycles 
7884: is still quite impressive, and presumably stems from the relatively sharp 
7885: confinement of efficient configuration-interaction within a spatial region close
7886: to the inner turning point of the Rydberg wave-packet.
7887: 
7888: 
7889: 
7890: \section{Characteristic properties of non-dispersive wave-packets}
7891: \label{CHPROP}
7892: 
7893: After presenting several examples of non-dispersive wave-packets in the previous
7894: chapters, we now study their specific properties in more detail. 
7895: Especially, several important physical
7896: processes which may affect the existence of wave-packets 
7897: have so far been hidden under the carpet~\cite{carpetmen}. 
7898: The two most important ones, at least for
7899: driven atoms, are ionization and spontaneous emission, and they will 
7900: be discussed
7901: in detail below. First, let us briefly discuss the general properties of
7902: wave-packet eigenstates 
7903: under the variation of various parameters of 
7904: the driven system 
7905: (e.g.,
7906: microwave amplitude 
7907: and frequency, the strength of 
7908: an external static 
7909: field, etc.).
7910: 
7911: 
7912: \subsection{Ionization rates and chaos assisted tunneling}
7913: \label{ION}
7914: 
7915: 
7916: Atoms driven by microwaves will eventually ionize. Therefore, the
7917: non-dispersive wave-packet states discussed up till now cannot be,
7918: rigorously speaking, discrete states, 
7919: they are rather resonances~\cite{goldberger50} with
7920: some finite life-times. Importantly, as we shall discuss in detail below,
7921: these life-times may be extremely long, of the order of millions of microwave
7922: periods. 
7923: In that sense, they are comparable to 
7924: those of highly 
7925: excited atomic Rydberg states,
7926: which also decay, by spontaneous emission, 
7927: on time scales of few millions of classical periods.
7928: Even more importantly, the life-times of the 
7929: non-dispersive wave-packets are typically orders of
7930: magnitude larger than the life-times of other 
7931: states in the Floquet spectrum: the wave-packets
7932: are particularly resistant
7933: to ionization. This is due to the 
7934: classical confinement
7935: of the
7936: electron inside the regular island. 
7937: To ionize, the electron has no other option but
7938: to tunnel
7939: out of the classically confining island,
7940: before gaining energy by diffusive excitation~\cite{casati88}. 
7941: The resonance island is strictly confining only
7942: for a one-dimensional system. For multi-dimensional systems, the 
7943: tori in the resonance islands are not fully isolating and a very
7944: slow classical diffusion process might eventually lead to ionization. 
7945: This, however, takes place on extremely long time scales and is
7946: completely negligible in atomic systems. 
7947: In practice, ionization of the wave-packet is essentially mediated by a pure
7948: quantum process, exponentially unlikely 
7949: in the
7950: semiclassical limit. As we shall see below, this tunneling
7951: process has quite interesting properties which may be quantitatively
7952: described 
7953: for microwave driven atoms. 
7954: More details 
7955: can be found in \cite{kuba95a,kuba98,kuba98a}. 
7956: 
7957: Due to the initial tunneling step, the life-times of 
7958: non-dispersive wave-packets will typically be 
7959: much longer than those of 
7960: Floquet states 
7961: localized in the chaotic
7962: sea surrounding the island \cite{abu95a}. Moreover, 
7963: since the ionization mechanism involves 
7964: chaotic
7965: diffusion, many quantum mechanical paths link the initial wave-packet to
7966: the final continuum. Thus, the life-time of the wave-packet will reflect 
7967: the interferences between those different possible paths,
7968: and will sensitively depend on parameters such as the microwave frequency or
7969: amplitude, that affect the interfering paths 
7970: through the chaotic sea. 
7971: These fluctuations,
7972: reported first in \cite{kuba95a}, are perfectly deterministic
7973: and
7974: resemble the conductance fluctuations observed in mesoscopic systems 
7975: \cite{stone85} . 
7976: \begin{figure}
7977: \centerline{\psfig{figure=bdzf51.eps,width=8cm,bbllx=100pt,bblly=150pt,bburx=520pt,bbury=600pt,angle=-90}}
7978: \caption{Typical fluctuations of the width (ionization rate)
7979: and of the energy (with respect to its averaged, smooth behavior)
7980: of the non-dispersive wave-packet of a two-dimensional
7981: hydrogen atom in a circularly polarized microwave field.
7982:  The data presented
7983: are obtained for small variations
7984: of the effective principal quantum number $n_0=\omega^{-1/3}$ around 40,
7985: and a scaled
7986: microwave electric field $F_0=0.0426$.
7987: To  show that the fluctuations 
7988: cover
7989: several orders of magnitude, we use a logarithmic vertical scale, and
7990: plot the absolute value of the shift.}
7991: \label{ionfig1}
7992: \end{figure}
7993: In fig.~\ref{ionfig1}, we show the fluctuations of the
7994: ionization rate (width) of the non-dispersive wave-packet
7995: of the two dimensional hydrogen atom in a circularly
7996: polarized microwave field. The energy levels
7997: and widths
7998: are obtained as explained in sec.~\ref{RSEF}, by numerical
7999: diagonalization of the complex rotated Hamiltonian. 
8000: All the data presented in this section have been obtained 
8001: in the regime where the typical ionization rate is smaller than the
8002: mean energy spacing between consecutive levels, so that the
8003: ionization can be thought as a small perturbation acting
8004: on bound states.
8005: The width (although 
8006: very small) displays
8007: strong fluctuations over several orders of magnitude.
8008: Similarly, the real part of the energy (i.e., the center of the
8009: atomic resonance) displays wild fluctuations. The latter
8010: can be observed only if the smooth variation
8011: of the energy level with the control parameter (following 
8012: approximately the semiclassical prediction given
8013: by eq.~(\ref{estharm}))
8014: is substracted.
8015: Therefore, we fitted the numerically obtained energies by a smooth function
8016: and substracted this fit to obtain the displayed fluctuations. Note that
8017: these fluctuations are so small that an accurate fit is 
8018: needed\footnote{In particular, 
8019: the semiclassical expression is not sufficiently
8020: accurate for such  a fit.}. This can be easily seen in fig.~\ref{solit2}
8021: where, on the scale of the mean level spacing, these fluctuations
8022: are invisible by eye (the level appears as a straight horizontal line).
8023: 
8024: The explanation  
8025: for the fluctuations is the following: 
8026: in a quantum language, they are due to the coupling
8027: between the localized wave-packet and states
8028: localized in the chaotic sea surrounding the resonance island. 
8029: While the energy 
8030: of the wave-packet is a smooth function of the
8031: parameters $F$ and $\omega,$  the energies
8032: of the chaotic states  display a complicated behavior characterized
8033: by level repulsion and large avoided crossings. It happens often
8034: that -- for some parameter values -- there is a quasi-degeneracy
8035: between the wave-packet eigenstate and a chaotic state, see the numerous
8036: tiny avoided crossings in fig.~\ref{solit2}. 
8037: There, the two
8038: states are efficiently mixed, the wave-packet captures
8039: some part of the coupling of the chaotic state to the continuum
8040: and its ionization
8041: width increases (see also \cite{abu95a}). 
8042: This is the very origin
8043: of the observed fluctuations. Simultaneously, the 
8044: chaotic state repels the wave-packet state leading to a deviation of the 
8045: energy from its smooth behavior,
8046: and thus to the observed fluctuations. 
8047: This mechanism is similar 
8048: to 
8049: ``chaos assisted tunneling'', described in the literature
8050: \cite{LB90,LB92,grossmann91,GDJH91b,GDJH93,plata92,bohigas93,bohigas93b,tomsovic94,shudo95,leyvraz96}
8051: for both, driven one-dimensional and two-dimensional autonomous systems.
8052: There, the tunneling rate between two symmetric
8053: islands -- which manifests itself through the splitting between
8054: the symmetric and antisymmetric states of a doublet -- 
8055: may be strongly enhanced by the
8056: chaotic transport between the islands. We have then a ``regular''
8057: tunneling escape 
8058: from one island, a chaotic diffusive transport from the vicinity of
8059: one island to the other (many paths, leading to interferences and resulting
8060: in large fluctuations of the splitting),
8061:  and another ``regular'' tunneling penetration into the
8062: second island. In our case, the situation is even simpler -- we have a
8063: ``regular'' tunneling escape supplemented by a chaotic diffusion and
8064: eventual ionization. Thus, instead of the level splitting, we observe 
8065: a shift of the energy level and a finite width.
8066: 
8067: Since these fluctuations stem from the coupling between the regular
8068: wave-packet state and a set of chaotic states, it is quite natural
8069: to model such a situation via a Random Matrix model \cite{kuba98}, 
8070: the
8071: approach being directly motivated by a similar treatment of the
8072: tunneling splitting in \cite{tomsovic94}. For details,  
8073: we refer the 
8074: reader to the original work~\cite{kuba98}. It suffices to say here that
8075: the model is characterized by three real parameters: $\sigma$ -- which
8076: characterizes the mean strength of the coupling between the regular state 
8077: and the chaotic levels, $\gamma$ -- which measures the decay of the chaotic
8078: states (due to ionization;  direct ionization transitions from
8079: the wave-packet state to the continuum are negligible),
8080: and $\Delta$ -- which is the mean level spacing of chaotic levels.
8081: The two physically relevant, dimensionless 
8082: parameters are  $\gamma/\Delta$ and $\sigma/\Delta$.
8083: In the perturbative regime $(\gamma/\Delta,\sigma/\Delta\ll 1)$ it is possible
8084: to
8085: obtain analytical \cite{kuba98} 
8086: predictions for the statistical distribution of the energy
8087: shifts $P(s)$ (of the 
8088: wave-packet's energy from its unperturbed value)
8089: and 
8090: for the distribution of its widths $P(\Gamma)$. $P(s)$ turns out to
8091: be a Cauchy distribution (Lorentzian), similarly to the tunneling
8092: splitting distribution found in \cite{tomsovic94}. The distribution of the 
8093: widths 
8094: is a bit more complicated (it is the square root of $\Gamma$ which
8095: is approximately 
8096: Lorentzian distributed). 
8097: The perturbative approach fails for the 
8098: asymptotic behavior of the 
8099: tails of the distributions, where an exponential cut-off is expected and
8100: observed in numerical studies \cite{kuba98,tomsovic94}. By fitting 
8101: the predictions of the Random Matrix model to the numerical data of 
8102: fig.~\ref{ionfig1}, we may finally extract the values of 
8103: $\gamma/\Delta$, the strength of the decay, and of $\sigma/\Delta,$  
8104: the coupling between the regular and the chaotic states.
8105: \begin{figure}
8106: \centerline{\psfig{figure=bdzf52.eps,width=8cm,bbllx=100pt,bblly=120pt,bburx=320pt,bbury=500pt,angle=-90}}
8107: \caption{The distribution of energy shifts
8108: (a) and 
8109: ionization widths 
8110: (b) for the non-dispersive wave-packet of a two-dimensional hydrogen atom
8111: in a circularly polarized field, obtained by numerical diagonalization
8112: of the Hamiltonian (large bins), compared to 
8113: the random matrix model (small bins). 
8114: Both distributions
8115: are shown on a double logarithmic scale to better visualize the behavior
8116: over a large range of shift and width values. 
8117: Since the energy shift 
8118: may be 
8119: positive or negative, we show the distribution of its modulus. 
8120: The random matrix model fits very well the numerical results,
8121: with both distributions showing 
8122: regions of algebraic behavior followed by an exponential cut-off.}
8123: \label{ionfig2}
8124: \end{figure}
8125: An example of such a fit is shown in fig.~\ref{ionfig2}. 
8126: The numerical data
8127: are collected around some mean values of $n_0$ and $F_0$, typically
8128: 1000 data points were used for a single fit \cite{kuba98}. This
8129: allowed us to study the dependence of the parameters 
8130: $\gamma/\Delta,\sigma/\Delta$ on $n_0$ and $F_0.$
8131: \begin{figure}
8132: \centerline{\psfig{figure=bdzf53.eps,width=8cm,bbllx=100pt,bblly=150pt,bburx=520pt,bbury=600pt,angle=-90}}
8133: \caption{ 
8134: Effective tunneling rate $\sigma/\Delta$  
8135: of the wave-packet (a), 
8136: as a function of the
8137: effective quantum number $n_0=\omega^{-1/3}$
8138: (the inverse of the effective Planck constant), 
8139: for fixed classical dynamics, $F_0=0.0426$. Note the
8140: exponential decrease for sufficiently high $n_0$ (the vertical scale is
8141: logarithmic). The corresponding effective chaotic 
8142: ionization rate $\gamma/\Delta$
8143: (b)
8144: smoothly increases with $n_0$, approximately as $n_0^2$.}
8145: \label{ionfig3}
8146: \end{figure}
8147: The dependence 
8148: on $n_0$
8149: is shown in fig.~\ref{ionfig3}.
8150: Clearly, the tunneling rate $\sigma/\Delta$ decreases 
8151: {\em exponentially} with $n_0.$
8152: Since $n_0$ is the inverse of the effective Planck constant in our problem
8153: (see the discussion in section~\ref{LIN1D} and eq.~(\ref{effective})), 
8154: this shows that
8155: \begin{equation}
8156: \sigma/\Delta \propto \exp \left (-\frac{S}{\hbar_{\mathrm eff}}\right),
8157: \label{dec_exp_tun}
8158: \end{equation}  
8159: where $S$, corresponding to some effective imaginary action 
8160: \cite{shudo95}, is found to be given for our specific choice of
8161: parameters by 
8162:  $S\simeq 0.06 \pm 0.01$ (as fitted from the plot). Such an exponential
8163:  dependence is a hallmark of a tunneling process, thus confirming that
8164:  the wave-packets are strongly localized in the island and communicate
8165:  with the outside world via tunneling.
8166: The $n_0$ dependence of the dimensionless chaotic ionization
8167: rate $\gamma/\Delta$ is very different: it shows a slow, algebraic 
8168: increase with $n_0$. A simple analysis based on a Kepler map
8169: \cite{casati88} description 
8170: would yield a linear increase with $n_0$, whereas our 
8171: data seem to suggest a quadratic function of $n_0$. This discrepancy
8172: is not very surprising, bearing in mind the simplicity of the
8173: Kepler map approach.
8174: 
8175: Similarly, we may study, for fixed $n_0$, the dependence of $\gamma/\Delta$
8176: and $\sigma/\Delta$ on $F_0$, i.e. on the microwave field strength. Such
8177: studies, performed for both linear and circular polarizations, have 
8178: indicated that, not very surprisingly, the chaotic 
8179: ionization rate
8180: $\gamma/\Delta$ increases rather smoothly with the microwave amplitude
8181: $F_0$. On the other hand, the tunneling rate $\sigma/\Delta$ shows
8182: pronounced non-monotonic variations with $F_0,$ see fig.~\ref{ionfig4}.
8183: \begin{figure}
8184: \centerline{\psfig{figure=bdzf54.eps,width=8cm,bbllx=100pt,bblly=150pt,bburx=520pt,bbury=600pt,angle=-90}}
8185: \caption{The tunneling rate $\sigma/\Delta$ (panel (a)) and the chaotic
8186: ionization rate $\gamma/\Delta$ (panel (b)), as a function of the scaled
8187: microwave amplitude $F_0$, for wave-packet eigenstates of a two-dimensional 
8188: hydrogen atom in a circularly polarized
8189: microwave field. 
8190: Observe the oscillatory behavior of the tunneling rate. The bumps
8191: are due to secondary nonlinear resonances in the classical dynamics of the
8192: system.}
8193: \label{ionfig4}
8194: \end{figure}
8195: This unexpected behavior can nontheless be explained \cite{kuba98}. 
8196: The bumps in $\sigma/\Delta$ occur at microwave field
8197: strengths where secondary nonlinear resonances emerge 
8198: within the 
8199: resonance island in
8200: classical phase space. 
8201: For circular polarization, this 
8202: corresponds to some resonance
8203: between two eigenfrequencies $\omega_+$ and $\omega_-$ (see
8204: sec.~\ref{CP} and figs.~\ref{solit2},\ref{solfig}) of the dynamics in the classical resonance island. 
8205: Such resonances 
8206: strongly perturb
8207: the classical dynamics and necessarily affect the
8208: quantum transport 
8209: from 
8210: the island.
8211: 
8212: Let us
8213: stress finally that, even for rather strong microwave fields (say $F_0=0.05$),
8214: where most of the other Floquet states have life-times of few tens or hundreds
8215: of microwave
8216: periods, and irrespective of the polarization of the driving field or of the 
8217: dimension of the accessible configuration space (1D, 2D or 3D), the life-time 
8218: (modulo fluctuations) 
8219: of a non-dispersive wave-packet is typically of the order of
8220: $10^5$ Kepler periods, for $n_0\simeq 60$. This may be used for their possible
8221: experimental detection, see section~\ref{LIFE}. 
8222: 
8223: 
8224: \subsection{Radiative properties}
8225: \label{SPO}
8226: 
8227: 
8228: So far, we have considered the interaction of 
8229: the atom with 
8230: the coherent driving field
8231: only. However, this is not the full story. 
8232: Since the driving field couples
8233: excited atomic states, it remains to be seen to which extent 
8234: spontaneous emission 
8235: (or, more precisely, the coupling to other, initially unoccupied modes of
8236: the electromagnetic field) affects the wave-packet properties. 
8237: This is very important, since the non-dispersive wave-packets
8238: are supposed to be long living objects, and spontaneous
8239: emission obviously limits their life-time.
8240: Furthermore, we have here an example of decoherence effects due to 
8241: interaction with
8242: the environment.
8243: More generally, the interaction of non-dispersive wave-packets
8244: with an additional weak external electromagnetic field may
8245: provide 
8246: a useful tool to probe 
8247: their properties.
8248: In particular, 
8249: their localization within the resonance island
8250: implies that 
8251: an external probe will couple them efficiently 
8252: only to neighboring states within the island. In turn, that   should
8253: make their experimental characterization easy and unambiguous.
8254: Of course, 
8255: external drive (microwave field) and probe must 
8256: not be treated
8257: on the same footing. 
8258: One should rather consider the 
8259: atom 
8260: dressed by the external 
8261: drive
8262: as a strongly coupled system,
8263: or use the Floquet picture described above,
8264: and 
8265: treat the additional mode(s) of the probe (environment) 
8266: as a perturbation. 
8267: We first start with the simplest situation, where a single mode
8268: of the environment is taken into account.
8269: 
8270: \subsubsection{Interaction of a non-dispersive wave-packet
8271: with a monochromatic probe field}
8272: \label{spofloq}
8273: Let us first consider the addition of a monochromatic probe field
8274: of frequency $\omega_p.$ The situation is very similar to the 
8275: probing of a time-independent system by a weak monochromatic field,
8276: with the only difference  
8277: that the Floquet Hamiltonian replaces
8278: the usual time-independent Hamiltonian. Thus, the
8279: weak probe field may induce a transition between two Floquet states
8280: if it is {\it resonant} with this transition, i.e. if
8281: $\hbar \omega_p$ is equal to the quasi-energy difference between
8282: the two Floquet states. According to Fermi's Golden Rule,
8283: the transition probability is proportional to the
8284: square of the matrix element coupling the initial Floquet state
8285: $|{\cal E}_i\rangle$ to the final one $|{\cal E}_f \rangle .$
8286: 
8287: Using 
8288: the Fourier representation of Floquet states,
8289: \begin{equation}
8290: |{\cal E}(t)\rangle = \sum_k{\exp (-ik\omega t) |{\cal E}^k\rangle},
8291: \end{equation}
8292: and averaging over one driving field cycle $2\pi/\omega$ we get
8293: \begin{equation}
8294: \langle {\cal E}_f | {\cal T} | {\cal E}_i\rangle =
8295: \sum_{k=-\infty}^{\infty}
8296: {\langle {\cal E}_f^k | {\cal T} | {\cal E}_i^k \rangle}.
8297: \label{matelm_floquet}
8298: \end{equation}
8299: $\cal T$ denotes the transition operator, usually some
8300: component of the dipole operator depending on the polarization
8301: of the probe beam.
8302: If the quasi-energy levels are not bound states but
8303: resonances with finite life-time (for example because of
8304: multiphoton transition amplitudes to the continuum) 
8305: this approach is easily extended \cite{cct92}, yielding
8306: the following expression for the photoabsorption cross-section of the
8307: probe field at frequency $\omega_p$:
8308: \begin{equation}
8309: \sigma(\omega_p) = \frac{4 \pi \omega_p \alpha}{c} 
8310: {\mathrm Im} \sum_{f}{|\langle {\cal E}_f | {\cal T} | {\cal E}_i \rangle |^2
8311: \left [\frac{1}{{\cal E}_f-{\cal E}_i-\omega_p}+\frac{1}{{\cal E}_f-{\cal E}_i+\omega_p} \right ]}
8312: \label{sigma_probe}
8313: \end{equation}
8314: where $\alpha$ is the fine structure constant, 
8315: and where ${\cal E}_f$ and ${\cal E}_i$ are the complex energies 
8316: of the initial and final
8317: Floquet states. The sum extends over all the Floquet states of the system.
8318: Although the Floquet energy spectrum is itself $\omega$-periodic
8319: (see sec.~\ref{QD}), this is {\em not} the case for the photoabsorption
8320: cross-section. Indeed, the Floquet states at energies ${\cal E}_f$ and
8321: ${\cal E}_f+\omega$ have the same Fourier components, but shifted
8322: by one unit in $k$, resulting in different matrix elements. 
8323: The photoabsorption spectrum is thus
8324: composed of series of lines separated by $\omega$ with unequal intensities.
8325: When the driving is weak, each Floquet state has a dominant Fourier component.
8326: The series then appears as a dominant peak accompanied by side bands
8327: shifted in energy by an integer multiple of the driving frequency $\omega.$
8328: In the language of the scattering theory \cite{cct92,goldberger50,faisal87}, 
8329: these side bands
8330: can be seen as the scattering of the probe photon assisted
8331: by one or several photons of the drive. In any case, the Floquet
8332: formalism is well suited, since it contains 
8333: this weak driving regime as a limiting case, as well as the strong 
8334: driving regime
8335: needed to 
8336: generate a non-dispersive wave-packet.
8337:    
8338: \subsubsection{Spontaneous emission from a non-dispersive wave-packet}
8339: \label{spospo}
8340: We now 
8341: address the situation where no probe field is added to
8342: the microwave field. 
8343: Still, photons of the driving field can be scattered in the 
8344: (initially empty) 
8345: remaining modes of
8346: the electromagnetic field. This is thus some kind of spontaneous emission
8347: or rather resonance fluorescence of the atom under coherent driving. 
8348: It can be seen as spontaneous
8349: emission of the dressed atom, where 
8350: an initial
8351: Floquet state decays spontaneously to another Floquet 
8352: state with a lower quasi-energy,
8353: the energy difference being carried by the spontaneous photon.
8354: As an immediate consequence, the spectrum of the emitted photons
8355: is composed of the resonance frequencies of the Floquet system,
8356: the same that are involved in eq.~(\ref{sigma_probe}).
8357: The decay rate along a transition depends
8358: on the dipole matrix element connecting the initial and the final states,
8359: but also on the density of modes for the emitted photons.
8360: If we consider, for simplicity, the case of free atoms,
8361: one obtains:
8362: \begin{equation}
8363: \Gamma_{if} = \frac{4\alpha^3 ({\cal E}_i-{\cal E}_f)^3}{3} 
8364: |\langle {\cal E}_f|{\cal T}|{\cal E}_i\rangle|^2
8365: \label{gamma}
8366: \end{equation}
8367: where ${\cal E}_i-{\cal E}_f$ is the positive energy difference 
8368: between the initial 
8369: and final Floquet states. As the matrix element of the dipole 
8370: operator $\cal T$
8371: is involved, clearly the localization properties of the Floquet states 
8372: will be of 
8373: primordial importance for the spontaneous emission process.
8374: 
8375: The total decay rate (inverse of the life-time) of a 
8376: state $|{\cal E}_i\rangle$
8377: is obtained by summing the partial rates
8378: $\Gamma_{if}$ connecting the initial state to all states with lower
8379: energy. It is not straightforward to determine
8380: which Floquet states contribute most to the decay rate --
8381: the two factors in eq.~(\ref{gamma}) compete:
8382: while $|\langle {\cal E}_f|{\cal T}|{\cal E}_i\rangle|^2$ tends to favor states localized close
8383: to the initial state (maximum overlap), the factor $({\cal E}_i-{\cal E}_f)^3$
8384: (due to the density of modes in free space)
8385: favors transitions to much less excited states. Which factor wins depends
8386: on the polarization of the driving field.
8387: 
8388: \subsubsection{Circular Polarization}
8389: \label{spocp} 
8390: Consider first a circularly polarized microwave field. 
8391: A first analysis of spontaneous emission
8392: has been given in \cite{ibb97b}, where the rotating frame 
8393: (see sec.~\ref{CP}) approach was used. The driven problem becomes
8394: then time-independent, and the analysis of spontaneous emission 
8395: appears to be simple.
8396: This is, however, misleading, and it is quite easy to omit some
8397: transitions with considerable 
8398: rate. 
8399: The full and correct analysis, both in the rotating and in the standard
8400: frame \cite{delande98},  discusses this problem extensively. The
8401: reader should consult the original papers for details.
8402:  
8403: A crucial point is to realize that the Floquet spectrum of the Hamiltonian
8404: in CP splits into separate blocks, all of  them being identical,
8405: except for a shift by an integer multiple of the driving frequency $\omega.$ 
8406: Each block corresponds to a fixed
8407: quantum number $\kappa=k+M$ where $k$ labels the photon block 
8408: (Fourier component) in the Floquet approach, while $M$ is the azimutal 
8409: quantum number. This merely signifies that the absorption
8410: of a driving photon of circular polarization $\sigma^+$ increases $M$ by
8411: one unit. In other words, $\kappa$ is nothing but the total
8412: angular momentum (along the direction of propagation of the microwave field)
8413: of the 
8414: entire system comprising the atom and the driving field.
8415: The separate $\kappa$ blocks 
8416: are coupled by spontaneous emission.
8417: Since, again, the spontaneously emitted photon carries 
8418: one quantum of angular
8419: momentum, spontaneous emission couples states within the 
8420: same $\kappa$-block (for
8421: $\pi$ polarization of the emitted photon w.r.t. the $z$ axis, 
8422: which leaves $M$ invariant) or 
8423: in neighboring $\kappa$-blocks $\kappa'=\kappa \pm 1$ (see fig.~\ref{spon1}).
8424: \begin{figure}
8425: \centerline{\psfig{figure=bdzf55.eps,width=9cm,bbllx=100pt,bblly=100pt,bburx=420pt,bbury=400pt}}
8426: \caption{Spontaneous emission transitions for the non-dispersive wave-packet
8427: of an atom 
8428: in
8429: a circularly polarized microwave field. The quasi-energy levels of the
8430: Floquet Hamiltonian can be split in series labelled by $\kappa$
8431: (total angular momentum of the atom and of the microwave field). 
8432: The various series
8433: are identical, except for an energy shift equal to an integer multiple
8434: of the microwave frequency $\omega.$
8435: The arrows indicate possible spontaneous transitions 
8436: leaving the initial state $|0,0,0\rangle$.
8437: Only arrows drawn with solid lines are allowed in the harmonic approximation.
8438: The position of the $|0,0,0\rangle$ wave-packet in each Floquet ladder is
8439: indicated by 
8440: the fat lines.}
8441: \label{spon1}
8442: \end{figure}
8443: $\sigma^+$ polarization of the emitted photon gives rise
8444: to higher frequency photons
8445: since -- for the same initial and final Floquet states --
8446: the energy difference in the $\sigma^+$ channel is larger by
8447: $\hbar \omega$ than in the $\pi$ channel (and by
8448: $2\hbar \omega$ than in the $\sigma^-$ channel), as immediately
8449: observed in fig.~\ref{spon1}. As the emission
8450: rate, eq.~(\ref{gamma}), changes with the cubic power of the energy difference,
8451: spontaneous photons with $\sigma^+$ polarization
8452: are expected to be dominant.
8453: 
8454: In the absence of any further approximation, the spontaneous emission
8455: spectrum is  
8456: fairly complicated - it consists of three series with
8457: different polarizations, $\sigma^{\pm}$ and $\pi$.
8458: We may use, however, the harmonic approximation,
8459: discussed in detail in sec.~\ref{CP}. The Floquet
8460: states localized in the vicinity of the stable fixed point
8461: may be labelled
8462: by three quantum numbers $|n_+,n_-, n_z\rangle$, 
8463: corresponding to the various excitations in
8464: the normal modes. The non-dispersive wave-packet we are most interested in
8465: corresponds to the ground state $|0,0,0\rangle$. The dipole operator (responsible for the spontaneous transition)
8466: may be expressed as a linear combination of the creation and annihilation 
8467: operators in these normal modes. Consequently, we obtain strong selection
8468: rules for dipole transitions between $|n_+,n_-,n_z\rangle$ states belonging
8469: to different ladders (at most $\Delta n_i=0,\pm 1$ with not all possibilities
8470: allowed -- for details see \cite{delande98}). The situation is even simpler
8471: for 
8472: $|0,0,0\rangle$,
8473: which may 
8474: decay 
8475: only via three transitions, all 
8476: $\sigma^+$ polarized 
8477: (i.e., from the $\kappa$ block 
8478: to the $\kappa-1$ block):
8479: \begin{itemize}
8480: \item a transition to the 
8481: $|0,0,0\rangle$ state 
8482: in the $\kappa-1$
8483: block. By definition, this occurs precisely at the microwave frequency 
8484: of the drive. One
8485: can view this process as elastic scattering of the microwave photon;
8486: \item a transition to the state $|1,0,0\rangle$,  
8487: at frequency $\omega - \omega_+$;
8488: \item a transition to the state $|0,1,0\rangle$, 
8489: at frequency $\omega + \omega_-$.
8490: \end{itemize}
8491: In the harmonic approximation, 
8492: explicit analytic expressions can be obtained for
8493: the corresponding transition rates \cite{delande98}. It suffices to say 
8494: here that in the semiclassical
8495: limit the elastic component becomes dominant, since its intensity scales as 
8496: $\omega^{5/3}$,
8497: while the intensities of the other two components 
8498: are proportional to $\omega^2,$ i.e., are typically
8499: weaker by a factor $n_0=\omega^{-1/3}$.
8500: This implies that the non-dispersive wave-packet decays exclusively
8501: (in the harmonic approximation) to its immediate neighbor
8502: states, 
8503: emitting a photon with frequency in the microwave range, comparable
8504: to the driving frequency. Direct decay to the atomic $|n=1,L=M=0\rangle$
8505: ground state
8506: or to weakly excited states of the system is forbidden by the selection rules
8507: of the dipole operator. This is easily understood: the CP nondispersive ground
8508: state wave-packet $|0,0,0\rangle$
8509: is built essentially from states with large angular momentum (of the
8510: order of $n_0$), and as it can lose only one unit of angular
8511: momentum per spontaneous emission event, it can decay only to similar
8512: states. When the harmonic approximation breaks down, additional lines
8513: may appear, but, for the same reason, in the microwave range only. 
8514: Another important observation is that the inelastic component at
8515: $\omega + \omega_-$ is by far stronger than the one at $\omega - \omega_+$.
8516: This is entirely due to the cubic power of the transition frequency entering
8517: the expression for the rate (\ref{gamma}). Note the sign difference, due to
8518: the sign difference between $\pm$ modes 
8519: in the harmonic hamiltonian, eq.~(\ref{hh}). 
8520: 
8521: In the semiclassical limit $\omega=n_0^{-3} \to 0,$ the decay is dominated
8522: by the elastic component, and the total decay rate is \cite{delande98}:
8523: \begin{equation}
8524: \Gamma = \frac{2 \alpha^3 \omega^{5/3} q^{-2/3}}{3},
8525: \label{jacks}
8526: \end{equation}
8527: what, multiplied by the energy $\omega$ of the 
8528: spontaneous photon, gives the energy loss 
8529: due to spontaneous emission:
8530: \begin{equation}
8531: \frac{dE}{dt} = \frac{2 \alpha^3 \omega^{8/3}
8532: q^{-2/3}}{3}=\frac{2\alpha^3\omega^4|x_{\rm eq}|^2}{3},
8533: \label{jacks2}
8534: \end{equation}
8535: where we used eq.~(\ref{q}).
8536: This is nothing but the result 
8537: obtained from classical electrodynamics 
8538: \cite{jackson} for a point charge moving
8539: on a circular orbit of radius $|x_{\rm eq}|$ with frequency $\omega$.
8540: Since the charge loses energy, it cannot survive on a circular orbit and
8541: would eventually fall onto the nucleus following a spiral trajectory. 
8542: This model
8543: stimulated Bohr's original 
8544: formulation
8545: of quantum mechanics. 
8546: Let us notice that the non-dispersive
8547: wave-packet is the first physical realization of the Bohr model. There is no
8548: net loss of energy since, in our case, the electron is driven by the microwave
8549: field and an emission at frequency $\omega$ occurs in fact
8550: as an elastic scattering of a microwave photon. Thus
8551: the non-dispersive wave-packet is a cure of the long-lasting
8552: Bohr paradox.
8553: 
8554: Figure \ref{spon2} shows the square of the dipole matrix elements
8555: connecting the non-dispersive wave-packet, for
8556: $n_0=60$ (i.e., microwave frequency $\omega=1/60^3$) and
8557: scaled microwave field 
8558: $F_0=0.04446$, to other Floquet states
8559: with lower energy. These are the results 
8560: of an exact numerical diagonalization of the full Floquet Hamiltonian.
8561: They are presented as a 
8562: stick spectrum because the widths of the important lines are very narrow
8563: on the scale of the figure, which is given as a function of 
8564: the energy difference
8565: between the initial and the final state, that is the frequency of the 
8566: scattered 
8567: photon. Thus, this figure shows the lines that could be observed
8568: when recording the photoabsorption of a weak microwave probe field.
8569: As expected, there is a dominant line at the frequency $\omega$ of the
8570: microwave, and two other lines at frequencies $\omega-\omega_-$
8571: and $\omega+\omega_+$ with comparable intensities, while
8572: all other lines are at least 10 times weaker. This means that
8573: the harmonic approximation works here very well; its predictions,
8574: indicated by the crosses in the figure, are in good quantitative
8575: agreement with the exact result (apart from tiny shifts  
8576: recognizable in the figure,
8577: which correspond to the mismatch between the exact and the 
8578: semiclassical energies 
8579: observed in figs.~\ref{solit2} and \ref{solfig}). This is not completely
8580: surprising as the energy levels themselves are well reproduced
8581: by this harmonic approximation, see sec.~\ref{rotating_frame}.
8582: However, the photoabsorption  spectrum probes the {\em wave-functions} 
8583: themselves 
8584: (through the overlaps) which are well known to be much more sensitive
8585: than the energy levels. The good agreement for both the energy spectrum
8586: and the matrix elements is a clear-cut proof of the
8587: reliability of the harmonic approximation for physically accessible principal
8588: quantum numbers, say $n_0<100$; in fact, it is good down to
8589: $n_0 \simeq 30$, and the non-dispersive wave-packet exists even
8590: for lower $n_0$ values (e.g. $n_0=15$ in \cite{delande95}) although
8591: the harmonic approximation is not too good at such low quantum numbers. 
8592: There were repeated claims in the literature 
8593: \cite{lee97,farrelly95a,farrelly95,brunello96,cerjan97,lee95}
8594: that the stability island 
8595: as well as the effective potential 
8596: are necessarily
8597: unharmonic in the vicinity of the equilibrium point, and that the
8598: unharmonic terms will destroy the stability of the non-dispersive wave-packets.
8599: The present results prove that these claims are doubly wrong:
8600: firstly, as explained in sec.~\ref{rotating_frame}, harmonicity is
8601: {\em not} a requirement for non-dispersive wave-packets to exist 
8602: (the only condition is the existence of a sufficiently large
8603: resonance island); secondly, the harmonic approximation
8604: is clearly a very good approximation even for moderate values
8605: of $n_0.$
8606: 
8607: \begin{figure}
8608: \centerline{\psfig{figure=bdzf56.eps,width=12cm,angle=-90}}
8609: \caption{Square of the dipole matrix element (scaled w.r.t. $n_0$, i.e.
8610: divided by $n_0^2$)
8611: connecting the $|0,0,0\rangle$ non-dispersive 
8612: wave-packet of a three-dimensional hydrogen atom
8613: in a circularly polarized microwave field with
8614: other Floquet states, as a function of the
8615: energy difference between
8616: the two states. The stick spectrum  is the exact result obtained
8617: from a numerical diagonalization 
8618: with $\omega=1/60^3$, corresponding to a principal
8619: quantum number $n_0=60$, and scaled amplitude
8620: $F_0=0.04446$; in natural units, the microwave frequency $\omega/2\pi$ 
8621: is 30.48 GHz,
8622: and the
8623: microwave amplitude 17.6 V/cm. The crosses represent  the
8624: analytic prediction within the harmonic approximation \protect\cite{delande98}.
8625: There are three dominant lines ($\sigma^+$ polarized) discussed in the text,
8626: other transitions (as well as transitions with
8627: $\sigma^-$ or $\pi$ polarizations) are negligible, what proves
8628: the validity of the harmonic approximation. 
8629: If a weak probe field (in the microwave domain) is applied to the system 
8630: in addition
8631: to the driving field, its absorption spectrum should therefore
8632: show the three dominant lines, allowing an unambiguous
8633: characterization of the non-dispersive wave-packet.
8634: }
8635: \label{spon2}
8636: \end{figure}
8637: 
8638: \begin{figure}
8639: \centerline{\psfig{figure=bdzf57.eps,width=12cm,angle=-90}}
8640: \caption{
8641: Same as figure \protect{\ref{spon2}}, but for the
8642: decay rates along the different transitions and in natural units.
8643: The density of modes
8644: of the electromagnetic field completely kills the transition at 
8645: frequency $\omega-\omega_+$,
8646: invisible in the figure. One can see a  small line at frequency approximately
8647: $\omega+2\omega_-$ (see arrow), an indication of a weak breakdown of the harmonic
8648: approximation.
8649: }
8650: \label{spon3}
8651: \end{figure} 
8652: 
8653: Multiplication by the free space density of states transforms fig.~\ref{spon2}
8654: in fig.~\ref{spon3}, which shows that the corresponding 
8655: spontaneous decay rates 
8656: are very low, of the order of 100 Hz at most. They are
8657: few orders of magnitude smaller than the ionization rates and thus may be
8658: difficult to observe. With increasing $n_0,$ the spontaneous rate decreases
8659: algebraically while the ionization rate decreases {\it exponentially}, 
8660: see sec.~\ref{spolp}. 
8661: Thus for large $n_0,$ the spontaneous emission may be the
8662: dominant process. For $F_0\simeq 0.05$ the cross-over may be expected around
8663: $n_0=200$. However, for smaller $F_0,$ the ionization rate decreases
8664: considerably, and for $F_0\simeq 0.03$ both rates become comparable around
8665: $n_0=60$. Still, a rate of few tens of photons (or electrons
8666: in the case of ionization) per second 
8667: may be quite hard to observe experimentally.
8668: 
8669: To summarize, resonance fluorescence of 
8670: non-dispersive wave-packets in 
8671: circularly polarized microwave occurs only in the microwave range
8672: (close to the driving frequency). In 
8673: particular, the elastic component (dominant in the semiclassical limit)
8674: does not destroy the wave-packet, the wave-packet merely converts the 
8675: microwave photon into a photon emitted with the same polarization, but in
8676: a different direction.
8677: Let us stress that we assumed the free space density of modes in this
8678: discussion. Since 
8679: the microwave field may be also 
8680: supplied to the atom by
8681: putting the latter in a microwave cavity, it should be 
8682: interesting to 
8683: investigate how the density of modes in such a cavity affects the
8684: spontaneous emission rate either by increasing
8685: or decreasing it (see \cite{haroche92} for a review)
8686: or, for special cavities (waveguides), even
8687: invalidates the concept of a decay rate \cite{lewenstein88,lewenstein88b}. 
8688:   
8689: \subsubsection{Linearly polarized microwave}
8690: \label{spolp}
8691:   
8692: Let us now discuss the spontaneous emission of non-dispersive wave-packets
8693: driven by a linearly polarized microwave field. 
8694: The situation becomes complicated since we should consider
8695: different wave-packets corresponding to (see fig.~\ref{lin3d_5})
8696: extreme librational states ($p=0$, located perpendicularly to the polarization
8697: axis), separatrix states elongated along the polarization axis, and 
8698: extreme rotational (maximal $p$,  doughnut shaped) states
8699: of the resonantly driven manifold. Clearly, all these wave-packet states have
8700: different localization properties and spontaneous emission will couple
8701: them to different final states. No systematic analysis of the effect has
8702: been presented until now, only results based on the simplified 
8703: one-dimensional model are available \cite{hornberger98}. 
8704: Those are of relevance 
8705: for the spontaneous emission of the separatrix based wave-packet
8706: and are reviewed below. Quantitatively we may, however, expect that
8707: the spontaneous emission properties of the extreme rotational wave-packet
8708: will resemble those of the non-dispersive wave-packet in circular
8709: polarization. Indeed, the linearly polarized wave 
8710: can be decomposed into two circularly
8711: polarized waves, and the extreme
8712: rotational state (fig.~\ref{lin3d_8}) is a coherent superposition of
8713: two circular wave-packets -- each locked on one 
8714: circularly polarized component -- moving in the opposite sense. 
8715: The decay of each
8716: component can then be
8717: obtained from the preceding discussion.
8718:  
8719: A completely different picture emerges for other wave-packets. Their electronic
8720: densities averaged over one period are concentrated along either the $\rho$
8721: (extreme librational) or $z$ axes and do not vanish close to the 
8722: nucleus (fig.~\ref{lin3d_5}).
8723: They have non-negligible dipole elements with Floquet states 
8724: built on low lying
8725: atomic states (i.e. states practically unaffected by the driving field).
8726: Because of the cubic power dependence of the decay rate, eq.~(\ref{gamma}),
8727: on the
8728: energy of the emitted photon, these will dominate the spontaneous emission. 
8729: Thus, in contrast to the CP case, 
8730: spontaneous decay will lead to the destruction of the wave-packet. 
8731: A quantitative analysis confirms this qualitative picture.
8732: To this end, 
8733: a general master equation formalism can be developed \cite{hornberger98}, which
8734: allows 
8735: to treat 
8736: the
8737: ionization process induced by the driving field exactly,
8738: while 
8739: the spontaneous emission is treated perturbatively, 
8740: as in the preceding section. 
8741: Applied 
8742: to the one-dimensional model of the atom (sec.~\ref{LIN1D}), 
8743: with the density of
8744: field modes of the real three-dimensional world, it is possible 
8745: to approximately model 
8746: the behaviour of the separatrix states of the three-dimensional atom.
8747:  
8748: A non-dispersive wave-packet may decay
8749: either by ionization or by spontaneous emission, the total 
8750: decay rate being the sum of the two rates
8751: \cite{hornberger98}. 
8752: Like
8753: in the CP case, the  
8754: decay rate to the atomic continuum decreases exponentially with $n_0$ 
8755: (since it is essentially a tunneling process, see sec.~\ref{ION}, eq.~(\ref{dec_exp_tun}))
8756: while the spontaneous decay rate depends 
8757: algebraically on $n_0$ \cite{bethe77}. The wave-packet is a coherent
8758: superposition of
8759: atomic states with principal quantum number close to $n_0=\omega^{-1/3}$, and
8760: the dipole matrix element between an atomic state $n$ and a weakly 
8761: excited state scales as $n^{-3/2}$\cite{bethe77}. 
8762: Since the energy of the emitted
8763: photon is of order one (in atomic units), eq.~(\ref{gamma})
8764: shows
8765: that the spontaneous emission rate should decrease like 
8766: $\alpha^3/n_0^3.$
8767: The numerical results, presented in fig.~\ref{spon4}, fully confirm
8768: this $1/n_0^3$ prediction.
8769: \begin{figure}
8770: \centerline{\psfig{figure=bdzf58.eps,width=12cm,angle=-90}}
8771: \caption{
8772: Comparison of the spontaneous decay rate, the 
8773: ionization rate, and their sum, for a
8774: non-dispersive wave-packet in a linearly polarized microwave field,
8775: as a 
8776: function of the principal quantum number $n_0$. 
8777: Microwave amplitude $F_0=0.04442,$ decay rates in atomic units. 
8778: The full decay rate exhibits a cross-over 
8779: from a dominantly coherent (ionization)
8780: to a dominantly incoherent (spontaneous emission) regime. 
8781: The fluctuations of the
8782: rate present in the coherent regime are suppressed in the incoherent regime.
8783: The data presented here are obtained by an exact numerical calculation
8784: on the one-dimensional model of the atom \protect\cite{hornberger98}, 
8785: see sec.~\ref{LIN1D}.
8786: }
8787: \label{spon4}
8788: \end{figure}
8789: However, the spontaneous decay of real, 3D wave-packets with near 1D
8790: localization properties (see fig.~\ref{lin3d_5}, middle column, and 
8791: fig.~\ref{linf_wp}) is
8792: certainly slower. 
8793: Indeed, these states are combinations of atomic states
8794: with various total angular momenta $L$; among them, only the
8795: low-$L$ values decay rapidly to weakly excited states, 
8796: the higher $L$
8797: components being coupled only to higher excited states. 
8798: In other words, they are dominantly composed by extremal parabolic Rydberg
8799: states, which have well-known decay properties \cite{bethe77}.
8800: Altogether,
8801: their decay rate is decreased by a factor of the order of $n_0,$  
8802: yielding a $n_0^{-4}$ law instead of  $n_0^{-3}$. 
8803: 
8804: On the other hand, since the ionization process is dominated by tunneling in
8805: the direction of the microwave polarization axis
8806: \cite{abuth,abu98a,abu97,abu97b}, the ionization rate in 3D remains globally
8807: comparable to the ionization rate in 1D, for the wave-packet launched along
8808: straight line orbits. This remains true even if the generic
8809: fluctuations of the ionization rate (see section~\ref{ION}) may induce locally
8810: (in some control parameter) large deviations between individual 3D and 1D
8811: decay rates\footnote{A similar behaviour is observed
8812: in circular polarization for the 2D and 3D non-dispersive wave-packets: they
8813: exhibit comparable ionization rates, but distinct fluctuations
8814: \protect{\cite{kuba95a}}.}. 
8815: Therefore, the transition from dominant ionization to dominant spontaneous
8816: decay will shift to slightly higher values of $n_0$ in 3D.
8817: As in the CP case studied above, this cross-over may be moved to smaller 
8818: values of $n_0$  by reducing the
8819: ionization rate, i.e., by decreasing $F_0$.
8820: 
8821: \subsection{Non-dispersive wave-packet as a soliton}
8822: \label{SOLI}
8823: 
8824: 
8825: 
8826: The non-dispersive character of the wave-packets discussed in this review
8827: brings to mind solitons, i.e. solutions of {\it nonlinear} wave equations
8828: that propagate without deformation: the non-linearity is there
8829: essential to overcome the spreading of the solution. The non-dispersive
8830: wave-packets discussed by us are, on the other hand, solutions of
8831: the {\it linear} Schr\"odinger wave equation, and it is not some
8832: non-linearity of the wave equation which protects them from spreading, but
8833: rather the periodic driving. Thus, at first glance,
8834: there seems to be no link  between both phenomena. This is
8835: not fully correct. One may 
8836: conceive non-dispersive wave-packets as solitonic
8837: solutions of particular nonlinear equations, 
8838: propagating not in time, but in parameter 
8839: space \cite{kuba97b}.
8840: The 
8841: evolution of energy levels in such a space, 
8842: called ``parametric level dynamics'',
8843: has been extensively studied (see 
8844: \cite{haake90,nakamura93} for 
8845: reviews), both for time-independent
8846: and  for periodically time-dependent systems. In the latter case, 
8847: the energy levels are the quasi-energies 
8848: of the Floquet Hamiltonian (see sec.~\ref{QD}). 
8849: 
8850: For the sake of simplicity, 
8851: we consider here the two-dimensional hydrogen atom 
8852: exposed to a circularly polarized microwave (sec.~\ref{2d_model}),
8853: where the explicit time dependence can be removed by 
8854: transforming to
8855: the rotating frame (see sec.~\ref{rotating_frame}), but completely
8856: similar results are obtained for the Floquet Hamiltonian of any 
8857: periodically time-dependent system.
8858: The Hamiltonian, given by eq.~(\ref{hrot}),
8859: \be
8860: H=\frac{{\vec p}^2}{2}-\frac{1}{r}+Fx-\omega L_z,
8861: \label{solieq0}
8862: \ee
8863: may be thought of as an example of a generic system of the form
8864: \be
8865: H(\lambda)=H_0+\lambda V,
8866: \label{solieq1}
8867: \ee
8868: where $\lambda$ is a parameter. In our case, 
8869: for example, the
8870: microwave amplitude may be tuned, leading to $V=x$ and $\lambda=F$. 
8871: The interesting quantities are then 
8872: the eigenvalues $E_i(\lambda)$ and the eigenfunctions
8873: $|\psi_i(\lambda)\rangle$ of eq.~(\ref{solieq1}). 
8874: Differenciating the
8875: Schr\"odinger equation with respect to $\lambda$,
8876: one shows (with some algebra)
8877: \cite{haake90} that the behavior of $E_i(\lambda)$ with
8878: $\lambda$ may be viewed as the motion of $N$ fictitious
8879: classical particles
8880: (where $N$ is
8881: the dimension of the Hilbert space) with positions $E_i$ and momenta
8882: $p_i=V_{ii}=\langle\psi_i|V|\psi_i\rangle$, governed by the
8883: Hamiltonian
8884: \begin{equation}
8885: {\cal H}_{\rm cl}=\sum_{i=1}^N\frac{p_i^2}{2}+\frac{1}{2}\sum_{i=1}^N\sum_
8886: {j=1,j\neq i}^N
8887: \frac{\mid{\cal L}_{ij}\mid^2}{(E_j-E_i)^2},
8888: \label{hclas}
8889: \end{equation}
8890: where
8891: ${\cal L}_{ij}=(E_i-E_j)\langle\psi_i|V|\psi_j\rangle$
8892: are additional independent variables obeying the general Poisson brackets
8893: for angular momenta. The resulting dynamics,
8894: although nonlinear, is integrable \cite{haake90}.
8895: 
8896: Let us now consider the parametric motion of some eigenstate 
8897: $\mid n_+,n_- \rangle$, 
8898: for example of the ground state wave-packet $|0,0\rangle$.
8899: Its coupling to other states is quite weak -- because
8900: of its localization in a well defined region of phase space -- 
8901: and the corresponding ${\cal L}_{ij}$ are consequently very small.
8902: If we 
8903: first suppose
8904: that the wave-packet state is well isolated (in energy)
8905: from other wave-packets (i.e., states with low values of $n_+,n_-$), 
8906: the fictitious particle associated with 
8907: $|0,0\rangle$ 
8908: basically ignores the other 
8909: particles and propagates freely at constant
8910: velocity. It preserves its properties across the
8911: successive interactions with neighboring states, 
8912: in particular its shape:
8913: in that sense, it is a solitonic solution of the
8914: equations of motion generated by Hamiltonian~(\ref{hclas}).
8915: 
8916: Suppose that, in the vicinity of some $F$ values, 
8917: another wave-packet 
8918: state (with low $n_+,n_-$ quantum numbers) becomes quasi-degenerate 
8919: with $|0,0\rangle$. 
8920: In the harmonic approximation, see sec.~\ref{rotating_frame}, the two states
8921: are completely uncoupled; 
8922: it implies that the corresponding
8923: ${\cal L}_{ij}$ vanishes and the two levels cross. 
8924: The coupling
8925: between
8926: the two solitons 
8927: stems from the {\em difference} between the exact hamiltonian
8928: and its harmonic approximation, 
8929: i.e. from third order or higher terms, beyond the harmonic
8930: approximation. Other
8931: $|n_+,n_-\rangle$ states having different slopes w.r.t. $F$ induce 
8932: ``solitonic collisions'' 
8933: at some other values of $F$.  
8934: To illustrate the effect,
8935: part of the spectrum of the two-dimensional hydrogen atom in 
8936: a CP microwave
8937: is shown in fig.~\ref{solfig}, as a function of the scaled microwave
8938: amplitude $F_0$.
8939: For the sake of clarity, the
8940: energy of the ground state wave-packet 
8941: $|0,0\rangle$ calculated
8942: in the harmonic approximation, eq.~(\ref{estharm}) is substracted,
8943: such that it appears as an almost
8944: horizontal line.
8945: Around $F_0=0.023$, it is crossed by another solitonic solution, 
8946: corresponding to
8947: the $|1,4\rangle$ wave-packet\footnote{Similarly to the 
8948: wave-packet $|1,3\rangle$ discussed in fig.~\protect\ref{solit2},
8949: the semiclassical harmonic
8950: prediction for the energy of $|1,4\rangle$ is not satisfactory.
8951: However, the slope of the energy level is well reproduced as a 
8952: function of $F_0$.}, 
8953: what represents the collision
8954: of two solitons. Since this
8955: avoided crossing 
8956: is narrow and well isolated from
8957: other avoided crossings, the wave-functions before and after 
8958: the crossing preserve
8959: their shape and character, as typical for an isolated two-level system.
8960: This may be further verified by wave-function plots before and after
8961: the collision (see 
8962: \cite{kuba97b} for more details). 
8963: \begin{figure}
8964: \centerline{\psfig{figure=bdzf59.ps,width=10cm}}
8965: \caption{
8966: The quasi-energy spectrum of a two-dimensional hydrogen atom 
8967: in a circularly polarized microwave field,
8968: as a function of the scaled microwave amplitude $F_0$ (for
8969: $n_0=60$). In order to emphasize the dynamics of wave-packet states,
8970: the semiclassical prediction, eq.~(\protect\ref{estharm}) for the
8971: ground state wave-packet energy is substracted from the
8972: numerically calculated energies. Consequently, the
8973: ground state wave-packet $|0,0\rangle$ is represented by 
8974: the almost horizontal line. 
8975: The dashed line represents the
8976: semiclassical prediction for the $|1,4\rangle$ wave-packet. Although it is 
8977: rather far from the exact result, the slope
8978: of the energy level is well reproduced. The size
8979: of the avoided crossing between the ``solitonic"
8980: levels $|0,0\rangle$ and  $|1,4\rangle$ is a direct measure
8981: of the failure of the harmonic approximation.
8982:  } 
8983: \label{solfig}
8984: \end{figure}
8985: The avoided crossings become larger 
8986: (compare fig.~\ref{solit2}) with increasing $F_0$. In fact, as mentioned 
8987: in sec.~\ref{rotating_frame},
8988: we have numerically verified that the solitonic character of the ground
8989: state wave-packet practically disappears at
8990: the  $1:2$ resonance, close to $F_0\simeq 0.065$ \cite{kuba97b}. For larger
8991: $F_0,$ while one may still find nicely
8992: localized wave-packets for {\em isolated} values of $F_0,$ 
8993: the increased size of the avoided crossings makes it difficult
8994: to follow the wave-packet 
8995: when sweeping $F_0.$ 
8996: For such strong fields, the ionization rate of
8997: wave-packet states becomes appreciable, comparable to the level
8998: spacing between consecutive states and the simple solitonic model breaks down.
8999: In order to understand the variations of the 
9000: (complex) energies of the resonances with $F$, a slightly
9001: more complicated model -- level dynamics in the complex plane --
9002: should be used~\cite{haake90}.
9003: 
9004: 
9005: 
9006: 
9007: \section{Experimental preparation and detection of non-dispersive wave-packets}
9008: \label{EXP}
9009: 
9010: In the preceding chapters, we have given an extensive
9011: theoretical description of the characteristic 
9012: properties of non-dispersive wave-packets in driven Rydberg
9013: systems. We have seen that these surprisingly robust ``quantum particles''
9014: are ubiquitous in the interaction of electromagnetic radiation with 
9015: matter. However, any theoretical analysis needs to be confronted with 
9016: reality, and we have
9017: to deal with the question of creating and identifying non-dispersive
9018: wave-packets in a laboratory experiment.
9019: In our opinion, none of the currently operational experiments on the
9020: interaction of Rydberg atoms with microwave fields 
9021: allows for an unambiguous identification of non-dispersive wave-packets,
9022: although some of them \cite{koch95b} 
9023: certainly have already populated such states.
9024: In the following, we shall therefore start out with 
9025: a brief description of the typical approach of state of the art experiments, 
9026: and subsequently extend on various alternatives to create and to probe  
9027: non-dispersive wave-packets in a real experiment.
9028: We do not aim at a comprehensive 
9029: review on the interaction 
9030: of Rydberg atoms with microwave fields, but rather refer to 
9031: \cite{abuth,delande94,sachath,koch95b,casati87b}
9032: for a detailed treatment of various aspects of this intricate
9033: problem. Here, we strictly focus on
9034: issues pertinent to our specific purpose.
9035: 
9036: \subsection{Experimental status}
9037: \label{EXSQ}
9038: 
9039: 
9040: The theoretical interest in the interaction of Rydberg states of atomic
9041: hydrogen
9042: with low frequency electromagnetic
9043: fields has been triggered by early experiments \cite{bayfield74}
9044: which showed a surprisingly
9045: efficient excitation and subsequent ionization of the atoms
9046: by the field. More precisely, a microwave field
9047: of frequency $\omega$
9048: comparable to the energy
9049: difference between the initial atomic state and its nearest neighbor
9050: was observed to induce appreciable ionization, 
9051: for atom-field interaction times of 
9052: approx. $100$ driving field cycles, and for 
9053: field amplitudes 
9054: beyond a certain threshold value
9055: (of the order of $5-10\%$ of the
9056: Coulomb field experienced by the Rydberg electron on its unperturbed
9057: Kepler orbit). This threshold behavior of the
9058: ionization probability as a function of the driving field amplitude
9059:  rather than
9060: of the driving frequency -- in apparent contradiction to the photoeffect --
9061: motivated a theoretical analysis of the classical dynamics of the Rydberg
9062: electron under external driving. It turned out that
9063: the ionization threshold marks the transition
9064: from  regular
9065: to  chaotic {\em classical dynamics} of the driven electron \cite{leopold78}.
9066: 
9067: The microwave ionization of atomic Rydberg states was thus identified
9068: as
9069: an experimental testing ground for quantum transport
9070: under the conditions of classically mixed regular chaotic
9071: dynamics, where the transport was simply measured by the experimentally
9072: observed ionization yield, or -- with some additional experimental effort --
9073: by the time dependent redistribution of the atomic population over the
9074: bound states \cite{bayfield88a,bluemel89a,bluemel91a,abu91}. 
9075: Depending on the precise value of the scaled frequency $\omega_0$ -- the ratio
9076: of the microwave frequency $\omega$ 
9077: to the Kepler frequency $\Omega_{\rm Kepler}$
9078: of the initially excited Rydberg atom, eq.~(\ref{omega0}) --
9079: of the driving field, theory soon predicted essentially
9080: ``classical'' ionization yields ($\omega_0<1.0$), or some quantum suppression
9081: of chaotic ionization ($\omega_0>1.0$) 
9082: \cite{casati84}, 
9083: mediated by the quantum mechanical
9084: interference effect known as 
9085: {\em dynamical localization}, 
9086: analogous to 
9087: Anderson localization in the electronic transport
9088: through disordered solids  
9089: \cite{fishman82,grempel84,brenner96,abu98b,wimbergerda,wimberger01}.
9090: The physical process involved in chaotic ionization is 
9091: classically deterministic
9092: diffusion, 
9093: therefore essentially statistical in nature, and 
9094: insensitive to the details of the transport process.
9095: Correspondingly, the mere ionization probability
9096: condenses all details of the ionization process in one single number,
9097: without revealing details on 
9098: individual local
9099: structures in phase space. It reflects the statistical characteristics
9100: of the excitation process, rather than the population of some well defined
9101: individual atomic  states in its course \cite{abuth,abu95a}.
9102: Hence, state of the art experiments are ``blind'' for the details of the
9103: atomic excitation process on the way to ionization, and therefore not
9104: suitable for the unambiguous identification of individual eigenstates of the
9105: atom in the field, notably of non-dispersive wave-packets. 
9106: The case is getting
9107: worse with additional complications which are unavoidable in a real experiment,
9108: such as the unprecise definition of the initial state the atoms are prepared
9109: in 
9110: \cite{koch95b,galvez88,sirko96,sirko93a,sauer92b,leeuwen85,koch95a,koch92b,koch92a,koch89}, 
9111: the experimental uncertainty on the envelope of the amplitude of the
9112: driving field experienced by the atoms as they enter the interaction
9113: region with the microwave 
9114: (typically a microwave cavity or wave guide) \cite{bluemel91a,sauer92b,noel00},
9115: stray electric fields due to contact potentials in the interaction region,
9116: and finally uncontrolled noise sources which may affect the
9117: coherence effects involved in the quantum mechanical transport process
9118: \cite{bayfield91}.
9119: On the other hand, independent experiments on the microwave ionization
9120: of Rydberg states of atomic hydrogen \cite{bayfield89,galvez88}, 
9121: as well as on hydrogenic initial states 
9122: of lithium \cite{noel00}, did indeed provide hard evidence
9123: for the relative stability of the atom against ionization when driven by a
9124: resonant field of scaled frequency $\omega_0\simeq 1.0$.
9125: Furthermore, in the hydrogen experiments, this stability was observed 
9126: to be insensitive
9127: to the polarization of the driving field, be it linear,
9128: circular or elliptical \cite{bellermann96}.
9129: These experimental findings suggest that some atomic dressed states 
9130: anchored to
9131: the principal resonance island in the classical phase space
9132: are populated by switching on the microwave field,
9133: since these states  tend to be more stable
9134: against ionization than  states localized in the chaotic sea~\cite{abu95a}, 
9135: see section \ref{ION}.
9136: 
9137: Consequently, for an unambiguous preparation and identification of
9138: non-dispersive wave-packets launched along well defined classical trajectories
9139: the experimental strategy has to be refined. 
9140: We suggest
9141: two 
9142: techniques for their preparation:
9143: \begin{itemize}
9144: \vskip -10pt
9145: \item The direct, selective optical excitation from
9146: a low lying state {\em in the presence of the microwave field}. 
9147: This approach actually 
9148: realizes some kind of  
9149: ``Floquet state absorption spectroscopy'' \cite{abu96}.
9150: \item The preparation of the appropriate atomic initial 
9151: state -- optionally in the presence of a static field --
9152: followed by  
9153: switching  
9154: the microwave field on the appropriate time scale 
9155: to the desired maximum field amplitude \cite{kuba95b}.
9156: \end{itemize}
9157: 
9158: Two, possibly complementary methods should allow 
9159: for an efficient detection of such wave-packets: 
9160: \begin{itemize}
9161: \item Floquet spectroscopy  -- this time 
9162: involving either microwave or optical transitions between states dressed
9163: by the microwave field;
9164: \item Measurements of the time-dependence of the ionization yield 
9165: of the non-dispersive
9166: wave-packet. This 
9167: requires the ability to vary the
9168: interaction time between the atoms and the microwave by more 
9169: than one order of magnitude
9170: \cite{abuth,abu95a,noel00,abu95b,arndt91,benson95}.
9171: \end{itemize}
9172: 
9173: All these techniques 
9174: are experimentally well-developed and actually realized in different, 
9175: currently
9176: operational experimental settings \cite{noel00,abu95b,noel00b}.
9177: The only prerequisite for an unambiguous identification of non-dispersive 
9178: wave-packets therefore remains an experimental set-up which allows 
9179: to follow these
9180: complementary strategies simultaneously.
9181: 
9182: 
9183: \subsection{Direct preparation}
9184: \label{SPEC}
9185: 
9186: 
9187: The 
9188: most straightforward way to populate a non-dispersive
9189: wave-packet state is its
9190: direct optical excitation
9191: in the presence of the driving field,
9192: from a 
9193: weakly excited state of the system at energy $E_0$.
9194: In section \ref{spofloq}, we have discussed how a weak electromagnetic probe 
9195: field can induce transitions between Floquet states.
9196: This is particularly easy if  
9197: one of the states involved is in an energetically low
9198: lying state. 
9199: Such a state
9200: is practically unaffected by the driving
9201: microwave field (which is weak as compared to the Coulomb field 
9202: experienced by a deeply 
9203: bound state, and very  far
9204: from any resonance), such that the corresponding Floquet state is almost
9205: exactly 
9206: identical with the time-independent atomic state. In other words,
9207: all the Fourier components of the Floquet state vanish,
9208: except the $k=0$ component, 
9209: which represents the 
9210: unperturbed atomic state
9211: $|\phi_0\rangle$.
9212: In such a case, the photoexcitation cross-section (\ref{sigma_probe}) becomes
9213: (neglecting the anti-resonant term):
9214: 
9215: \begin{equation}
9216: \sigma(\omega_p) = \frac{4 \pi \omega_p \alpha}{c}
9217: {\mathrm Im} \sum_{f}{|\langle {\cal E}_f^0 | {\cal T} | \phi_0 \rangle |^2
9218: \frac{1}{{\cal E}_f-E_0-\omega_p}},
9219: \label{sigma_probe_simple}
9220: \end{equation}
9221: where the sum extends over all Floquet states with energy ${\cal E}_f$ and
9222: involves only their $k=0$ Fourier component
9223: \footnote{Alternatively, the sum could be rewritten as a sum
9224: over one Floquet zone only, with all the Fourier
9225: components $|\phi_f^k\rangle$ involved, with the denominator
9226: replaced by ${\cal E}_f+k\omega-E_0-\omega_p.$}.
9227: 
9228: Eq.~(\ref{sigma_probe_simple}) shows that the excitation probability 
9229: exhibits a maximum any
9230: time the laser is scanned across a frequency which is resonant with
9231: the transition from the 1s ground state $|\phi_0\rangle$ 
9232: to a specific dressed state of the atom
9233: in the field. Fig.~\ref{abs_spec} shows
9234: an example for the photoabsorption probability from the ground state
9235: of atomic hydrogen in the presence of 
9236: a microwave field and a parallel static electric field. The microwave
9237: frequency is resonant with
9238: atomic transitions in the region of $n_0\simeq 60$.
9239: Clearly, the
9240: cross-section shows extremely narrow peaks each of which corresponds to
9241: a Floquet eigenstate of the atom in the field. As a matter of fact, the
9242: state marked by the arrow is similar to the dressed state of the 3D atom
9243: displayed in fig.~\ref{linf_wp}, 
9244: a wave-packet periodically moving along the
9245: field polarization axis. As obvious from the figure, this state can be
9246: efficiently reached by direct excitation from the ground state.
9247: Furthermore, due to its sharp signature in
9248: $\sigma(\omega_p)$, it is easily and unambiguously identified.
9249: On the other hand, this kind of preparation of the wave-packet is
9250: obviously reserved to those dressed states 
9251: which
9252: have nonvanishing overlap with the deeply bound atomic states. 
9253: For wave-packets tracing circular or elliptical orbits far from the
9254: nucleus, another strategy is needed, that
9255: is discussed below.
9256: 
9257: \begin{figure}
9258: \centerline{\psfig{figure=bdzf60.eps,width=9cm,angle=-90}}
9259: \caption{Photo-excitation of highly excited Rydberg states of the
9260: hydrogen atom in the presence of a linearly polarized microwave field
9261: of frequency $\omega/2\pi=30.48$ GHz and amplitude
9262: $F= 7.93$ V/cm and a parallel static electric field $F_s=2.38$ V/cm.
9263: The initial state is the ground state of the atom, and the
9264: polarization of the probe beam is parallel to the static
9265: and microwave fields. The spectrum 
9266: is displayed
9267: in the region where Floquet states are mainly composed of Rydberg states
9268: with principal quantum number around $n_0=60$, while the microwave driving 
9269: is resonant with the Kepler frequency of such states. Hence, some of
9270: the Floquet states are trapped in the non-linear resonance island
9271: and behave as non-dispersive wave-packets. Of special interest
9272: is the state marked with an arrow, which is similar to the non-dispersive
9273: wave-packet displayed in fig.~\protect{\ref{linf_wp}}, localized
9274: in all three dimensions of space (the field amplitude is
9275: slightly different). Its large photo-excitation probability
9276: should make a direct experimental preparation possible. At the scale of the figure,
9277: the width of the various lines is very small, and the spectrum is
9278: almost a pure stick spectrum.
9279: }
9280: \label{abs_spec}
9281: \end{figure}
9282: A slightly adapted spectroscopic approach should 
9283: be equally 
9284: useful for the unambiguous identification of wave-packet eigenstates.
9285: Instead of probing the dressed spectrum from a weakly excited
9286: state using a laser field, one may equally well probe the local structure
9287: of the dressed spectrum in the vicinity of the wave-packet eigenstate
9288: by inducing transitions from the wave-packet to neighboring states
9289: by a second, weak microwave field of linear or circular 
9290: polarization \cite{bluemel90a}.
9291: Such stimulated transitions will be mediated by the dipole matrix elements
9292: given in equation~(\ref{matelm_floquet}), and allow to measure
9293: the energy spacings in the
9294: immediate vicinity of the wave-packet state directly. Hence, microwave probe
9295: spectroscopy should be an extremely
9296: sensitive probe, since it allows for the unambiguous
9297: identification of the wave-packet via the characterization of its local
9298: spectral environment.
9299: 
9300: Given the spectroscopic
9301: resolution which is nowadays available in the optical as well as in the
9302: microwave domain, the spectroscopic approach outlined above seems to be the method
9303: of choice for an unambiguous identification, and -- where possible --
9304: for an efficient launch of nondispersive wave-packets along a periodic
9305: orbit of the classical dynamics. What it requires, however, is a
9306: precise determination of Floquet spectra from the accompanying
9307: quantum calculation. Fortunately, both for hydrogen and for alkali
9308: atoms, the necessary theoretical quantum data may be obtained from 
9309: already existing
9310: software \cite{krug00,krug01}.
9311: 
9312: Although we elaborated in this review paper only the case of the hydrogen
9313: atom, the general concepts are also fruitful for non-hydrogenic atoms.
9314: Indeed, the major difference between the Rydberg electron in a hydrogen
9315: atom and in a non-hydrogenic atom is the existence in the latter case
9316: of an ionic core which affects the classical and quantum dynamics of the
9317: Rydberg electron. On the scale of a Rydberg atom, the ionic core is 
9318: a extremely small object which will thus induce a very local perturbation.
9319: As long as the Rydberg electron does not approach the ionic core,
9320: it behaves completely similarly in hydrogen or non-hydrogenic atoms.
9321: Thus, the properties of non-dispersive wave-packets tracing circular
9322: or elliptical classical orbits are essentially independent of the ionic core,
9323: and the hydrogenic analysis holds. 
9324: For orbits which come close to the nucleus, the ionic core may scatter
9325: the Rydberg electron. Thus, instead of being indefinitely trapped on a
9326: torus inside a resonance island, it may happen that the Rydberg electron
9327: hops from a torus to another one when it gets close to the nucleus.
9328: This of course will affect the long time classical and quantum dynamics.
9329: Nevertheless, it remains true that most of the time the classical
9330: dynamics -- and consequently the phase locking phenomenon responsible
9331: for the existence of non-dispersive wave-packets -- is identical
9332: to the hydrogenic dynamics. From the quantum point of view, the ionic
9333: core is responsible for the existence of non-zero quantum defects
9334: in the low angular momentum channels. The energy levels, mixed by the microwave
9335: driving, will thus be significantly shifted from their
9336: hydrogenic positions. However, the {\em structure} of the energy levels --
9337: grouped in manifolds -- will essentially survive, see \cite{krug00}.
9338: It is likely that some non-dispersive wave-packets also exist in
9339: non-hydrogenic atomic species.
9340: 
9341: 
9342: \subsection{Preparation through tailored pulses}
9343: \label{PTP}
9344: 
9345: Another, indirect method for preparing  non-dispersive wave-packets
9346: is also available. This will be the method of choice for wave-packets
9347: moving along classical orbits of large angular momentum (small
9348: eccentricity). Such states, obviously, are not accessible to a direct
9349: optical excitation from weakly excited, low angular momentum states.
9350: The same general scheme may be also applicable to high eccentricity
9351: wave-packets although in that case we expect that the direct
9352: excitation may be more efficient and flexible.
9353: The method to be discussed here
9354: consists of two stages.
9355: We first prepare the atom in a well chosen and well defined initial
9356: highly excited state, and then turn the microwave field on relatively slowly,
9357: from zero amplitude to its plateau value $F_{\rm max}$.
9358: 
9359: A non-dispersive wave-packet is a single eigenstate $|{\cal E}\rangle$
9360: of the Floquet Hamiltonian
9361: describing the driven system at fixed driving field amplitude
9362: $F.$
9363: As shown in sections  \ref{LIN3D}, \ref{CP} and \ref{SOLI},
9364:  the evolution of the quasienergies
9365: of the driven atom with an external control parameter like the driving field
9366: amplitude is rather complicated.
9367:  It reflects the dramatic transformation
9368: of the structure of classical phase space, manifesting in an
9369: abundance of avoided crossings of various sizes in the level dynamics.
9370: Still, as exemplified in figs.~\ref{fig_mathieu}, \ref{solit2}, and \ref{solfig},
9371: the wave-packet states may be followed 
9372: rather easily 
9373: under changes of $F$ (parametrized by $t$, during the switching of the pulse)
9374: in the level dynamics, 
9375: in agreement with their ``solitonic''
9376: character (see section \ref{SOLI}).
9377: Nontheless, the very same figures illustrate clearly that the targeted 
9378: wave-packet state undergoes
9379: many avoided crossings as the  microwave amplitude is 
9380: swept. To remain in 
9381: a single eigenstate, the avoided crossings should be passed
9382: either adiabatically or diabatically, with a
9383: branching ratio at an individual crossing
9384: being described by the well known Landau-Zener scenario
9385: \cite{landau2,breuer89b}.
9386: Consequently, if we want to populate an individual wave-packet eigenstate
9387: from a field free atomic state $|\psi_0\rangle$,
9388: we need some knowledge
9389: of the energy level dynamics. Then it is possible to identify those
9390: field free states which are connected to the wave-packet via
9391: adiabatic and/or diabatic transitions in the network of energy levels,
9392: and subsequently
9393: to design $F(t)$ such as to transfer population
9394: from $|\psi_0\rangle$ to $|{\cal E}\rangle$ most efficiently.
9395: A precise experimental preparation of $|\psi_0\rangle$
9396: is the prerequisite of any such approach.
9397: 
9398: When the driving field is increased from zero, the major modification
9399: in the classical phase space is the emergence
9400:  of the resonance
9401: island (see figs.~\ref{lin3d_4}, \ref{lin3d_9}, \ref{lin3d_10}). 
9402: Quantum mechanically, the states with initial
9403: principal quantum number close to $n_0=\omega^{-1/3}$ will
9404: enter progressively inside the resonance island.
9405: For a one-dimensional system, the Mathieu equation, discussed
9406: in section \ref{section_mathieu}, fully describes the evolution of the
9407: energy levels in this regime. As shown for example
9408: in figure~\ref{fig_mathieu}, the non-dispersive wave-packet
9409: with the best localization, i.e. $N=0,$ 
9410: is -- in this simple situation -- adiabatically connected 
9411: to the field-free state with principal quantum number closest
9412: to $n_0,$ i.e. the eigenstate $\kappa=0$ of the
9413: Mathieu equation. When the Mathieu parameter $q,$ eq.~(\ref{map_mathieu3}), 
9414: is of the order
9415: of unity, the state of interest is trapped in the resonance island, which
9416: happens at field amplitudes given by eq.~(\ref{ftrapping})
9417: for the one-dimensional atom, and for 
9418: linear polarization of the microwave field.
9419: A similar scaling
9420: is expected for other polarizations, too.
9421: In the interval
9422: $F\leq F_{\rm trapping}$, the field has to be increased slowly enough such
9423: as to avoid losses from the ground state to the excited states of
9424: the Mathieu equation, at an energy separation of the order of $n_0^{-4}$.
9425: The most favorable situation is then the case of
9426: ``optimal'' resonance (see section~\ref{QD}),
9427: when $n_0$ is an integer, the situation in figure~\ref{fig_mathieu}.
9428: The wave-packet state is always separated from the other
9429: states by an energy gap comparable to its value at $F=0$,
9430: i.e. of the order of $3/(2n_0^4)$.
9431: The situation is less favorable if $n_0$ is not an integer,
9432: because the energy gap between the wave-packet
9433: of interest and the other states is smaller when $F \to 0.$
9434: The worst case is met when $n_0$ is half-integer: the free
9435: states $n_0+1/2$ and $n_0-1/2$ are quasi-degenerate, and selective
9436: excitation of a single wave-packet is thus more difficult.
9437: 
9438: The appropriate time scale for switching on the field is given 
9439: by the inverse of the
9440: energy splitting, i. e. for ``optimal" resonance
9441: \begin{equation}
9442: \tau_{\rm trapping}\sim n_0^4=n_0\times 2\pi/\omega,
9443: \label{t_smallF}
9444: \end{equation}
9445: or $n_0$ driving field periods.
9446: 
9447: Once trapped in the resonance island, the coupling to states localized
9448: outside the island will be residual -- mediated by quantum mechanical
9449: tunneling, see section \ref{ION} -- and the size of the
9450: avoided crossings between the trapped and the untrapped states is exponentially
9451: small. After adiabatic switching into the resonance island on a time 
9452: scale of $n_0$
9453: Kepler orbits, we now have to switch diabatically from $F_{\rm trapping}$ to
9454: some final
9455: $F$ value, in order to avoid adiabatic
9456: losses from
9457: the wave-packet into other states while passing through the avoided crossings.
9458: 
9459: 
9460: The preceding discussion is based on a one-dimensional model
9461: and the Mathieu equation. Taking into account the other
9462: ``transverse'' degrees of freedom is not too difficult. Indeed,
9463: as noticed in sections \ref{LIN3D}, \ref{CP}, 
9464: the various time scales of the problem
9465: are well separated. The transverse motion is slow and can be
9466: adiabatically separated from the fast motion in $(\hat{I},\hat\theta)$. 
9467: Instead of getting
9468: a single set of energy levels, one gets a family of sets,
9469: the various families being essentially uncoupled. An example
9470: for the 3D atom in a linearly polarized microwave field is shown
9471: in fig.~\ref{lin3d_4}.
9472:  It follows that the estimate for the trapping field
9473: and the
9474: switching time are essentially the same as for 1D systems.
9475: Inside the resonance island, the situation becomes slightly more
9476: complicated, because there is not a single frequency for
9477: the secular motion, but several frequencies along the
9478: transverse degrees of freedom. For example, in CP,
9479: it has been shown that there are three eigenfrequencies, 
9480: eqs.~(\ref{omegas},\ref{ompmz}), in the harmonic
9481: approximation -- see section \ref{rotating_frame}.
9482:  This results in a large number of excited energy levels
9483: which may have avoided crossings with the ``ground state", i.e.
9484: the nondispersive wave-packet we want to prepare.
9485: As shown in section \ref{SOLI}, most of these avoided
9486: crossings are extremely small and can be easily crossed diabatically.
9487: However, some of them are rather large, especially when there
9488: is an internal resonance between two eigenfrequencies. Examples are
9489: given in figs.~\ref{solit2} and \ref{solfig}, where $\omega_+=3\omega_-$ and 
9490: $\omega_+=4\omega_-$, respectively. 
9491: 
9492: These
9493: avoided crossings are large and dangerous, because the states
9494: involved lie {\em inside} the resonance island, which thus
9495: loses its protective character.
9496: They are mainly due to the unharmonic character of the
9497: Coulomb potential.
9498: Their size  may be qualitatively analyzed as we do below on the
9499: CP example, 
9500: expecting similar
9501: sizes of the avoided crossings for any polarization.
9502: 
9503: The unharmonic corrections to the
9504: harmonic approximation around the stable fixed point
9505: $x_e,p_e$ in the center of the nonlinear resonance -- as outlined in
9506: section \ref{CP} -- are due to the higher order terms
9507: $(-1)^j\tilde{x}^j/j!x_e^{j+1}$ in the
9508: Taylor series of the Coulomb potential, where
9509: $\tilde{x},\tilde{y}$ are excursions from
9510: the equilibrium position.
9511: $\tilde{x}$ and $\tilde{y}$
9512: can be expressed as linear combinations of $a_{\pm}^{\dagger}$ and
9513: $a_{\pm}$ operators \cite{delande98} giving
9514: $\tilde{x},\tilde{y}\sim \omega^{-1/2}\sim n_0^{3/2}$. Furthermore,
9515: the equilibrium distance from the nucleus scales as the size of the atom,
9516: $x_e\sim n_0^2$, and, therefore,
9517: \begin{equation}
9518: \tilde{x}^j/x_e^{j+1}\sim n_0^{\left(-2-\frac{j}{2}\right)}.
9519: \label{abu_ptp_scaleac}
9520: \end{equation}
9521: 
9522: A state $|n_+,n_-\rangle$ is obtained
9523: by the excitation of
9524: $N_+$ quanta in the $\omega_+$-mode and of $N_-$ quanta in the
9525: $\omega_-$-mode, respectively, i. e.
9526:  by the application of the
9527: operator product $(a_+^{\dagger})^{n_+}(a_-^{\dagger})^{n_-}$ on
9528: the wave-packet state $|0,0\rangle$, which -- by virtue of
9529: eq.~(\ref{abu_ptp_scaleac}) -- will be subject to an unharmonic correction
9530: scaling like
9531: \begin{equation}
9532: \Delta E_{\rm unharmonic}\sim n_0^{\left(-2-\frac{n_++n_-}{2}\right)}.
9533: \label{abu_ptp_unharm}
9534: \end{equation}
9535: Hence, the size of the
9536: avoided crossings between the wave-packet eigenstate and
9537: excited states of the local potential around the stable fixed point
9538: decreases with the number of quanta in the excited modes.
9539: 
9540: In addition, we can determine the width $\Delta F_{\rm unharmonic}$
9541: of such avoided crossings in
9542: the driving field amplitude $F$, by differentiation of the energy
9543: (\ref{enharm})
9544: of $|n_+,n_-\rangle$ with respect to $F$.
9545: Then, the difference between the energies of two eigenstates localized
9546: in the resonance island is found to scale like $Fn_0$. Defining
9547: $\Delta F_{\rm unharmonic}$ by the requirement that $Fn_0$ be of the order
9548: of $\Delta E_{\rm unharmonic}$, we find
9549: \begin{equation}
9550: \Delta F_{\rm unharmonic}\sim n_0^{\left(-3-\frac{n_++n_-}{2}\right)}.
9551: \label{abu_ptp_fun}
9552: \end{equation}
9553: Since we want to switch the field to a maximum value
9554: $F_{\rm max}\sim n_0^{-4}$, the Landau-Zener formula
9555: \begin{equation}
9556: \tau\sim \frac{F_{\rm max}}{\Delta E \Delta F}
9557: \label{abu_ptp_lz}
9558: \end{equation}
9559: yields
9560: \begin{equation}
9561: \tau_{\rm unharmonic}\sim n_0^{n_-+n_++1}\sim n_0^{n_-+n_+-2}\ \ \
9562: {\mathrm microwave\ periods} 
9563: \label{abu_ptp_tunharm}
9564: \end{equation}
9565: for the scaling behavior of the time scale which guarantees diabatic
9566: switching through avoided crossings of the wave-packet eigenstate with
9567: excited states of the elliptic island.
9568: Let us stress that this is only a very rough estimate of the
9569: switching time,  some numerical
9570: factors (not necessarily close to unity) are not taken into account.
9571: 
9572: \begin{figure}
9573: \centerline{\psfig{figure=bdzf61.ps,width=14cm}}
9574: \caption{
9575: Snapshots of the electronic density for a two-dimensional
9576: hydrogen atom exposed to a circularly polarized microwave field
9577: with increasing amplitude. The  microwave
9578: field amplitude is switched on
9579: according to eq.~(\protect\ref{abu_ptp_pulse}), with maximum scaled field
9580: $F_{0,\rm max}= 0.03$ and
9581: $T_{\rm switch}=400\times 2\pi/\omega$, where $\omega$ is resonant with
9582: $n_0=60$ (frequency $\omega=1/(60.5)^3$).
9583: The evolution of the initial circular state $n=M=60$ is numerically
9584: computed by solving the time dependent Schr\"odinger equation in a
9585: convenient Sturmian basis.
9586: Top-left -- $t=0$ (initial circular state);
9587: top-middle -- $t=20$ microwave periods;
9588: top-right -- $t=60$ periods, bottom-left -- $t=100$ periods;
9589: bottom-middle -- the final state, $t=400$ periods;
9590: bottom-right -- the non-dispersive wave-packet (exact Floquet
9591: eigenstate): it is almost indistinguishable from the previous wave-function,
9592: what proves that the excitation process efficiently
9593: and almost exclusively populates the state of interest. The box extends
9594: over 10000 Bohr radii in both directions, with the nucleus
9595: at the center. The microwave field is along the horizontal axis,
9596: pointing to the right.}
9597: \label{ptp_fig1}
9598: \end{figure}
9599: 
9600: The above predictions can be checked, e.g., by a numerical integration of the
9601: time dependent Schr\"odinger equation for a  microwave-driven 
9602: atom, taking into account the  
9603: time-dependent amplitude of the field. An
9604: exemplary calculation on the two-dimensional model atom 
9605: (see sec.~\ref{2d_model}) 
9606: can be found in \cite{kuba97a}, for 
9607: CP driving.
9608: Fig.~\ref{ptp_fig1} shows the evolution
9609: of the electronic density
9610: of the atomic wave-function (initially prepared in the circular Rydberg
9611: state $n=M=60$) during the rising part of the driving
9612: field envelope, modeled by 
9613: \begin{equation}
9614: F(t)=F_{\rm max}\sin^2\left(\frac{\pi t}{2T_{\rm switch}}\right).
9615: \label{abu_ptp_pulse}
9616: \end{equation}
9617: 
9618: The driving field frequency was chosen according to the resonance condition
9619: with the $n_0=60$ state,
9620: with a maximum scaled amplitude $F_{0,\rm max}=0.03$.
9621: Inspection of 
9622: fig.~\ref{solfig} shows that, for this value of $F_{\rm max}$, the crossing
9623: between the wave-packet eigenstate and the state $|n_+=1,n_-=4\rangle$ 
9624: has to
9625: be passed diabatically after adiabatic
9626: trapping within the principal resonance. By virtue of the above estimations
9627: of the adiabatic and the diabatic time scales, the switching
9628: time (measured in driving field cycles)
9629: has to be chosen such that $n_0<T_{\rm switch}<n_0^3$
9630: (in microwave periods).
9631: Clearly, the pulse 
9632: populates the desired wave-packet 
9633: once the driving field
9634: amplitude 
9635: reaches its maximum value. More quantitatively, 
9636: the overlap of the final state after  
9637: propagation of the time-dependent Schr\"odinger equation 
9638: with the
9639: wave-packet eigenstate of the driven atom in the field
9640: (bottom-right panel) amounts to $94\%$. Since losses of atomic
9641: population due to ionization
9642: are negligible on the time scales considered in the figure, $6\%$ of the
9643: initial atomic population is lost during the switching process.
9644: The same calculation, done for the realistic three-dimensional
9645: atom with $n_0=60$ gives the same result, proving that
9646: the $z$-direction (which is neglected in the 2D model) 
9647: is essentially irrelevant in this problem.
9648: \begin{figure}
9649: \centerline{\psfig{figure=bdzf62.eps,width=12cm,angle=-90}}
9650: \caption{Overlap between the wave-function obtained at the end of
9651: the microwave turn-on and the exact target state representing the
9652: non-dispersive wave-packet, as a function of
9653: the switching time $T_{\rm switch},$ for the two-dimensional
9654: hydrogen atom (circles). The filled squares
9655: indicate the results obtained for a fully
9656: three-dimensional atom. $F_{\rm max}$, $\omega$, and $n_0$  
9657: as in fig.~\protect\ref{ptp_fig1}.}
9658: \label{ptp_fig2}
9659: \end{figure}
9660: Figure~\ref{ptp_fig2}
9661: shows the efficiency of the proposed switching scheme
9662: as a function of the switching time $T_{\rm switch}$, expressed
9663: in units of microwave periods. Observe that too
9664: long switching times tend to be less effective, since the avoided crossings
9665: passed during the switching stage  
9666: are not traversed diabatically. The rough estimate, eq.~(\ref{abu_ptp_tunharm}),
9667: overestimates the maximum switching time by one order of magnitude.
9668: On the other hand, 
9669: too short switching times do not allow the wave-packet 
9670: to localize inside the resonance island. However, a wide range
9671: of switching times remains where good efficiency is achieved.
9672: 
9673: For a given initial state $|\psi_0\rangle$
9674: of the atom, only the
9675: resonance condition defining the driving field frequency is to some extent
9676: restrictive, as depicted in fig.~\ref{ptp_fig3}.
9677: It is crucial that the initially excited field-free state
9678: is adiabatically connected 
9679: (through the Mathieu equation) to the ground state wave-packet. 
9680: The best choice is ``optimal resonance", but the 
9681: adiabaticity is preserved 
9682: if $n_0$ is changed by less than one half, see section~\ref{QD}. 
9683: This corresponds to a relative 
9684: change of $\omega$ of the order of $3/2n_0.$
9685: Given the spectral resolution of presently available
9686: microwave generators, the definition of the
9687: frequency with an accuracy of less than 1\%
9688: is not a limitation. 
9689: The exact numerical calculation displayed in fig.~\ref{ptp_fig3}
9690: fully confirms that efficient excitation is possible as long as 
9691: $n_0=\omega^{-1/3}-1/2$ matches the effective principal 
9692: quantum number of the
9693:  initially excited field-free state within a margin of $\pm 1/2$ 
9694: (in the range [59.5,60.5]).
9695: 
9696: \begin{figure}
9697: \centerline{\psfig{figure=bdzf63.eps,width=12cm,angle=-90}}
9698: \caption{Overlap between the wave-function obtained at the end of
9699: the microwave turn-on and the exact target state representing the
9700: non-dispersive wave-packet as a function of $n_0$,
9701: obtained for the two-dimensional
9702: hydrogen atom (circles). $F_{\rm max}$ is 
9703: as in fig.~\protect\ref{ptp_fig1}, and the switching time 
9704: is $T_{\rm switch}=250$ microwave periods. 
9705: The initial state corresponds to a circular
9706: $n=60$ state of a 2D hydrogen atom.}
9707: \label{ptp_fig3}
9708: \end{figure}
9709: 
9710: In conclusion,
9711: the preparation of non-dispersive wave-packets by excitation
9712: of a Rydberg state followed by careful switching of the
9713: microwave field can be considered as an efficient method, provided
9714: a clean experimental preparation of the atomic initial state can be
9715: achieved. Furthermore, the boundaries --
9716: eqs.~(\ref{t_smallF}) and (\ref{abu_ptp_tunharm}) --
9717: imposed on the time scale for the
9718: switching process leave a sufficient flexibility for the experimentalist 
9719: to efficiently 
9720: prepare the wave-packet.
9721: A final word is in place on the homogeneity of the driving field amplitude
9722: experienced by the atoms in the ``flat top region'' of the interaction,
9723: i.e. after the switching from the field free state into
9724: the wave-packet state at $F(t)=F_{\rm max}$. In any laboratory experiment,
9725: a slow drift of the amplitude will be unavoidable over the interaction volume.
9726: Hence, slightly different non--dispersive wave-packets will coexist
9727: at various spatial positions. Since the ionization rate
9728: of nondispersive wave-packets is rather sensitive
9729: with respect to detailed values of the parameters, see
9730: section \ref{ION}, this should manifest itself
9731: by a deviation of the time dependence of the ionization yield from purely
9732: exponential decay.
9733: 
9734: 
9735: \subsection{Life time measurements}
9736: \label{LIFE}
9737: 
9738: 
9739: Given the above, rather efficient experimental 
9740: schemes for the population of non-dispersive wave-packet eigenstates --
9741: either via direct optical Floquet absorption or through an appropriate
9742: switching procedure --
9743: we still need some means to prove that we
9744: really {\em did} populate the wave-packet. As a matter of fact, to provide 
9745: unambiguous experimental evidence, one has to test various 
9746: characteristic properties of the wave-packet, so as to
9747: exclude accidental coincidences. A natural way is Floquet spectroscopy 
9748: (see sec.~\ref{SPEC}),
9749: i.e. probing the structure of the Floquet quasi-energy levels,
9750: in either
9751: the optical or the microwave regime
9752: (via absorption, stimulated emission, Raman spectroscopy etc.). Another
9753: possibility is to explore unique properties of wave-packet Floquet states.
9754: For example, as discussed in sec.~\ref{ION}, these states exhibit 
9755: extremely small ionization rates. 
9756: Hence, an experimentally accessible quantity to identify
9757: these states is the time-dependence of their survival probability, i.e. 
9758: of the probability not to ionize during an interaction time $t.$
9759: It is
9760: given by \cite{abuth,abu95c}
9761: \begin{equation}
9762: P(t)=\sum_{\epsilon}|c_{\epsilon}|^2\exp(-\Gamma_{\epsilon}t),
9763: \label{abu_lt_pion}
9764: \end{equation}
9765: where
9766: the  $c_{\epsilon}$ denote the expansion coefficients of the initial 
9767: field-free 
9768: state in 
9769: the Floquet basis,
9770: at a given value of microwave amplitude $F$.
9771: 
9772: If the selective population of the wave-packet is successful, 
9773: only one Floquet state contributes to $P(t)$, and the 
9774: decay of the population to the atomic continuum should manifest in 
9775: its exponential decrease, as opposed to a multiexponential  
9776: decrease in the case of a broad distribution of the 
9777: $c_{\epsilon}$ over the Floquet states 
9778: \cite{abuth,abu95a,abu98b,wimbergerda,wimberger01,abu95b}. 
9779: Of course, the distinction between 
9780: an exponential and an algebraic decay law requires the variability of 
9781: the experimental interaction time over more than one order of magnitude. 
9782: This is a nontrivial task in experiments on atomic Rydberg states of 
9783: hydrogen, since the typical velocities of the atomic beam are of the order
9784: of $1000\ \rm m/s$. That significantly restricts the interval on which
9785: the interaction time may be changed, taking into account the typical
9786: size (in the cm-range) of
9787: the atom-field interaction region 
9788: \cite{koch95b,bayfield85}. 
9789: However, the feasibility of such measurements has already been demonstrated
9790: in microwave experiments on rubidium Rydberg states, where the interaction
9791: time has been scanned from approx. $100$ to approx. $100000$ field cycles,
9792: i.e., over three orders of magnitude \cite{abu95b,arndt91}.
9793: Note that, whereas the dynamics
9794: of the driven Rydberg electron 
9795: along a Kepler ellipse of large eccentricity will 
9796: certainly be affected by the presence of a non-hydrogenic core, 
9797: non-dispersive wave-packets as the ones discussed in 
9798: sections \ref{CP} and \ref{EP} 
9799: can certainly be launched along circular trajectories, since the 
9800: Rydberg electron of the rubidium atom essentially experiences a Coulomb
9801: field on such a circular orbit. 
9802: 
9803: To use the character of the decay as a means 
9804: to identify the wave-packet,
9805: the microwave field amplitude should be sufficiently large
9806: to guarantee that other Floquet states localized in the chaotic sea (see sec.~\ref{ION}) 
9807: decay rapidly. Otherwise, the observation
9808: of a mono-exponential decay simply suggests that we succeeded in populating
9809: a single Floquet state - not necessarily a wave-packet \cite{abu95b}.
9810: The appropriate choice of the driving 
9811: field amplitude $F$, such that appreciable ionization  
9812: is achieved for the 
9813: longer experimentally accessible 
9814: interaction times,
9815: should therefore allow for the experimental identification of the 
9816: mono-exponential decay from the wave-packet to the atomic continuum, but also
9817: -- by varying $F$ -- of the variations of the decay rate with $F,$
9818: which is predicted to fluctuate wildly over several orders of magnitude,
9819: see section \ref{ION}. 
9820: Note, however, that this requires an excellent homogeneity of the
9821: microwave field (e.g., provided by a high quality microwave cavity), 
9822: as the fluctuations take place over
9823: rather small intervals of $F.$
9824: 
9825: 
9826: \section{Conclusions}
9827: \label{CONC}
9828: 
9829: 
9830: In this report, we have shown that novel and highly robust eigenstates 
9831: of periodically 
9832: driven quantum systems -- non-dispersive wave-packets -- are born 
9833: out of classically mixed regular-chaotic dynamics. As much as a mixed phase 
9834: space is generic for classical Hamiltonian systems, non-dispersive 
9835: wave-packets are a generic manifestation thereof on the quantum level, given
9836: a sufficiently high density of states (needed to resolve finite-size phase 
9837: space structures). While we described their semiclassical properties and
9838: their experimental preparation, manipulation, and identification during the 
9839: largest part of this report for a specific system -- atomic Rydberg states
9840: driven by a microwave field -- it is clear from our approach 
9841: that such ``quantum particles'' can be anchored to any nonlinear resonance 
9842: between a periodic drive and a periodic trajectory of a Hamiltonian 
9843: system. As an alternative example, we have briefly touched upon the 
9844: atomic realization of the gravitational bouncer, though many other 
9845: realizations in simple quantum optical or atomic and molecular systems can be
9846: thought of. Let us only mention unharmonic traps for ions, atoms, or BEC
9847: condensates, periodically kicked atoms \cite{reinhold01},
9848: as well as molecular dynamics \cite{gerber98,gerber01} 
9849: on adiabatic potential surfaces
9850: (the driven frozen planet briefly discussed in section \ref{HE} may 
9851: be conceived
9852: as opening a perspective in this direction). Nontheless, atomic 
9853: Rydberg states remain arguably the best objects to study the 
9854: fundamental properties of non-dispersive wave-packets as the realization of
9855: Schr\"odinger's dream \cite{schroe26}.
9856:  On one hand, they are 
9857: microscopic realizations of the Keplerian motion and of Bohr's orbitals using
9858: a well understood non-linear dynamical system. 
9859: On the other hand, they possess the essential complication which open 
9860: quantum systems add to bounded Hamiltonian dynamics -- the 
9861: driving-induced, coherent coupling to the atomic continuum of free electronic
9862: states. On top of that, all these features can be controlled in 
9863: real laboratory experiments, and we might actually dream of probing the
9864: characteristic properties of nondispersive wave-packets on single, trapped 
9865: atoms or ions, using novel experimental approaches yet to come. Let us finally 
9866: dare to speculate on the potential use of non-dispersive wave-packets in 
9867: coherent control: given their spectacular robustness -- which we abundantly 
9868: illustrated in this report -- it is clear that they provide a means to store 
9869: and to ``ship'' quantum probability densities in and across phase space,
9870: e.g., under adiabatic changes of the driving field polarization and/or of the 
9871: strength or orientation of additional static fields. Given the recent 
9872: advances in coherent control of molecular reactions employing laser fields
9873: \cite{gerber98} -- which so far do not explore the unique perspectives of 
9874: nonlinear dynamics -- it looks like a promising (and challenging) program to 
9875: systematically study non-dispersive wave-packets in molecular reaction 
9876: dynamics.
9877: 
9878: 
9879: \section{Acknowledgments}
9880: 
9881: It is a pleasure 
9882: to acknowledge a longstanding and fruitful collaboration with Robert
9883: G\c{e}barowski, Beno\^{\i}t Gr\'emaud, 
9884: Klaus Hornberger, Andreas Krug, Romek Marcinek, 
9885: Krzysiek Sacha, Peter Schlagheck, and Sandro
9886: Wimberger on non-dispersive wave-packets and related topics 
9887: over the past five years.
9888: 
9889: We acknowledge support of bilateral collaborations via
9890: programmes Procope (German-French) and Polonium (Polish-French).
9891: J.Z. acknowledges support by Polish Committee for Scientific Research
9892: under grant 2P03B00915. Laboratoire Kastler Brossel is laboratoire 
9893: de l'Universit{\'e} Pierre et Marie
9894: Curie et de l'Ecole Normale Sup{\'e}rieure, unit{\'e} mixte de
9895: recherche 8552 du CNRS. 
9896: CPU time on various computers has been provided by IDRIS (Orsay) and 
9897: RZG (Garching).
9898: 
9899: 
9900: 
9901: \newpage
9902: \begin{thebibliography}{100}
9903: 
9904: \bibitem{englert95}
9905: J.~A. Bergou and B.~G. Englert, J. Mod. Opt. {\bf 45},  701  (1998).
9906: 
9907: \bibitem{schroe26}
9908: E. Schr\"odinger, \natw {\bf 14},  664  (1926).
9909: 
9910: \bibitem{lichtenberg83}
9911: A.~J. Lichtenberg and M.~A. Lieberman, {\em Regular and Stochastic Motion},
9912:   Vol.~38 of {\em Applied Mathematical Sciences} (Springer, Berlin, 1983).
9913: 
9914: \bibitem{raman97}
9915: C. Raman, T.~C. Weinacht, and P.~H. Bucksbaum, \pr A {\bf 55},  R3995  (1997).
9916: 
9917: \bibitem{yeazell90}
9918: J.~A. Yeazell, M. Mallalieu, and J. C.~R.~Stroud, \prL {\bf 64},  2007  (1990).
9919: 
9920: \bibitem{alber91}
9921: G. Alber and P. Zoller, \phr {\bf 199},  231  (1991).
9922: 
9923: \bibitem{landau2}
9924: L.~D. Landau and E.~M. Lifschitz, {\em Quantum Mechanics} (Pergamon, Oxford,
9925:   1977).
9926: 
9927: \bibitem{parker86}
9928: J. Parker and J. C.~R.~Stroud, \prL {\bf 56},  716  (1986).
9929: 
9930: \bibitem{alber86}
9931: G. Alber, H. Ritsch, and P. Zoller, \pr {\bf 34},  1058  (1986).
9932: 
9933: \bibitem{averbukh89}
9934: I.~S. Averbukh and N.~F. Perelman, \pla {\bf 139},  449  (1989).
9935: 
9936: \bibitem{yeazell91}
9937: J.~A. Yeazell and J. C.~R.~Stroud, \pr {\bf 43},  5153  (1991).
9938: 
9939: \bibitem{hillery84}
9940: M. Hillery, R.~F. O'Connell, M.~O. Scully, and E.~P. Wigner, \phr {\bf 106},
9941:   12  (1984).
9942: 
9943: \bibitem{moyal49}
9944: J.~E. Moyal, Proc. Camb. Phil. Soc. Math. Phys. Sci. {\bf 45},  99  (1949).
9945: 
9946: \bibitem{landau1}
9947: L.~D. Landau and E.~M. Lifschitz, {\em Mechanics} (Pergamon, Oxford, 1994).
9948: 
9949: \bibitem{haake90}
9950: F. Haake, {\em Quantum Signatures of Chaos}, Vol.~54 of {\em Springer Series in
9951:   Synergetics} (Springer, Berlin, 1991).
9952: 
9953: \bibitem{glauber63}
9954: R.~J. Glauber, \pr {\bf 131},  2766  (1963).
9955: 
9956: \bibitem{mandel90}
9957: L. Mandel and E. Wolf, {\em Coherence and Quantum Optics} (Cambridge University
9958:   Press, Cambridge, 1995).
9959: 
9960: \bibitem{cct92}
9961: C. Cohen-Tannoudji, J. Dupont-Roc, and G. Grynberg, {\em Atom-Photon
9962:   Interactions: Basic Processes and Applications} (John Wiley and Sons, New
9963:   York, 1992).
9964: 
9965: \bibitem{cct73}
9966: C. Cohen-Tannoudji, B. Diu, and F. Lalo\"e, {\em M\'ecanique quantique}
9967:   (Hermann, Paris, 1973).
9968: 
9969: \bibitem{liboff80}
9970: R.~L. Liboff, {\em Introductory Quantum Mechanics} (Holden-Day, San Francisco,
9971:   1980).
9972: 
9973: \bibitem{husimi40}
9974: K. Husimi, Proc. Phys. Math. Soc. Japan {\bf 22},  264  (1940).
9975: 
9976: \bibitem{bethe77}
9977: H.~A. Bethe and E.~E. Salpeter, {\em Quantum Mechanics of One- and Two-Electron
9978:   Atoms} (Plenum Publishing Corp., New York, 1977).
9979: 
9980: \bibitem{goldberger50}
9981: M.~S. Goldberger and K.~M. Watson, {\em Collision Theory} (Wiley, New York,
9982:   1964).
9983: 
9984: \bibitem{faisal87}
9985: F.~H.~M. Faisal, {\em Theory of Multiphoton Processes} (Plenum Press, New York,
9986:   1987).
9987: 
9988: \bibitem{yeazell88}
9989: J.~A. Yeazell and J. C.~R.~Stroud, \prL {\bf 60},  1494  (1988).
9990: 
9991: \bibitem{marmet94}
9992: L. Marmet {\it et~al.}, \prL {\bf 72},  3779  (1994).
9993: 
9994: \bibitem{mallalieu94}
9995: M. Mallalieu and J. C.~R.~Stroud, \pr {\bf A49},  2329  (1994).
9996: 
9997: \bibitem{weinacht99}
9998: T.~C. Weinacht, J. Ahn, and P.~H. Bucksbaum, Nature {\bf 397},  233  (1999).
9999: 
10000: \bibitem{ottb}
10001: E. Ott, {\em Chaos in Dynamical Systems} (Cambridge University Press,
10002:   Cambridge, 1993).
10003: 
10004: \bibitem{lee97}
10005: E. Lee, A.~F. Brunello, and D. Farrelly, \pr A {\bf 55},  2203  (1997).
10006: 
10007: \bibitem{schroe68}
10008: E. Schr\"odinger,  in {\em Sources of Quantum Mechanics}, edited by B.~L.
10009:   van~der Waerden (Dover, New York, 1968).
10010: 
10011: \bibitem{delande94}
10012: D. Delande and A. Buchleitner, \amop {\bf 35},  85  (1994).
10013: 
10014: \bibitem{abu95d}
10015: A. Buchleitner and D. Delande, \prL {\bf 75},  1487  (1995).
10016: 
10017: \bibitem{ibb94}
10018: I. Bia{\l}ynicki-Birula, M. Kalinski, and J.~H. Eberly, \prL {\bf 73},  1777
10019:   (1994).
10020: 
10021: \bibitem{ahn00}
10022: J. Ahn, T.~C. Weinacht, and P.~H. Bucksbaum, Science {\bf 287},  463  (2000).
10023: 
10024: \bibitem{meyer00}
10025: D.~A. Meyer, Science {\bf 289},  1431a  (2000).
10026: 
10027: \bibitem{kwiat00}
10028: P.~G. Kwiat and R.~J. Hughes, Science {\bf 289},  1431a  (2000).
10029: 
10030: \bibitem{bucksbaum00}
10031: P.~H. Bucksbaum, J. Ahn, and T.~C. Weinacht, Science {\bf 289},  1431a  (2000).
10032: 
10033: \bibitem{berman77}
10034: G.~P. Berman and G.~M. Zaslavsky, Phys. Lett. {\bf 61A},  295  (1977).
10035: 
10036: \bibitem{henkel92}
10037: J. Henkel and M. Holthaus, \pr {\bf A45},  1978  (1992).
10038: 
10039: \bibitem{holthaus94}
10040: M. Holthaus, Prog. Theor. Phys. Supplement {\bf 116},  417  (1994).
10041: 
10042: \bibitem{holthaus95}
10043: M. Holthaus, \csf {\bf 5},  1143  (1995).
10044: 
10045: \bibitem{abuth}
10046: A. Buchleitner, Ph.D. thesis, Universit\'e Pierre et Marie Curie, Paris, 1993.
10047: 
10048: \bibitem{farrelly95a}
10049: D. Farrelly, E. Lee, and T. Uzer, \prL {\bf 75},  972  (1995).
10050: 
10051: \bibitem{ibb95}
10052: I. Bia{\l}ynicki-Birula, M. Kalinski, and J.~H. Eberly, \prL {\bf 75},  973
10053:   (1995).
10054: 
10055: \bibitem{farrelly95}
10056: D. Farrelly, E. Lee, and T. Uzer, \pla {\bf 204},  359  (1995).
10057: 
10058: \bibitem{kalinski95a}
10059: M. Kalinski, J.~H. Eberly, and I. Bia{\l}ynicki-Birula, \pr {\bf A52},  2460
10060:   (1995).
10061: 
10062: \bibitem{kalinski95b}
10063: M. Kalinski and J.~H. Eberly, \pr {\bf A52},  4285  (1995).
10064: 
10065: \bibitem{delande95}
10066: D. Delande, J. Zakrzewski, and A. Buchleitner, Europhys. Lett. {\bf 32},  107
10067:   (1995).
10068: 
10069: \bibitem{kuba95a}
10070: J. Zakrzewski, D. Delande, and A. Buchleitner, Phys. Rev. Lett. {\bf 75},  4015
10071:    (1995).
10072: 
10073: \bibitem{ibb96}
10074: I. Bia{\l}ynicki-Birula and Z. Bia{\l}ynicka-Birula, Phys. Rev. Lett. {\bf 77},
10075:    4298  (1996).
10076: 
10077: \bibitem{eberly96}
10078: J.~H. Eberly and M. Kalinski,  in {\em Multiphoton Processes 1996, Proceedings
10079:   of the 7th International Conference on Multiphoton Processes,
10080:   Garmisch-Partenkirchen, Germany, October 1996}, Vol.~154 of {\em Institute of
10081:   Physics Conference Series}, edited by P. Lambropoulos and H. Walther
10082:   (Institute of Physics, Bristol and Philadelphia, 1997), pp.\ 29--36.
10083: 
10084: \bibitem{kalinski96a}
10085: M. Kalinski and J.~H. Eberly, \prL {\bf 77},  2420  (1996).
10086: 
10087: \bibitem{brunello96}
10088: A.~F. Brunello, T. Uzer, and D. Farrelly, \prL {\bf 76},  2874  (1996).
10089: 
10090: \bibitem{kalinski96b}
10091: M. Kalinski and J.~H. Eberly, \pr {\bf A53},  1715  (1996).
10092: 
10093: \bibitem{kalinski97}
10094: M. Kalinski and J.~H. Eberly, \prL {\bf 79},  3542  (1997).
10095: 
10096: \bibitem{kalinski98}
10097: M. Kalinski, \pr {\bf 57},  2239  (1998).
10098: 
10099: \bibitem{ibb97a}
10100: I. Bia{\l}ynicki-Birula and Z. Bia{\l}ynicka-Birula, Phys. Rev. Lett. {\bf 78},
10101:    2539  (1997).
10102: 
10103: \bibitem{ibb97b}
10104: Z. Bia{\l}ynicka-Birula and I. Bia{\l}ynicki-Birula, \pr {\bf A56},  3629
10105:   (1997).
10106: 
10107: \bibitem{delande97}
10108: D. Delande, J. Zakrzewski, and A. Buchleitner, \prL {\bf 79},  3541  (1997).
10109: 
10110: \bibitem{kuba97a}
10111: J. Zakrzewski and D. Delande, \jpb {\bf 30},  L87  (1997).
10112: 
10113: \bibitem{cerjan97}
10114: C. Cerjan, E. Lee, D. Farrelly, and T. Uzer, \pr {\bf A55},  2222  (1997).
10115: 
10116: \bibitem{kuba97b}
10117: J. Zakrzewski, A. Buchleitner, and D. Delande, Z. Phys. {\bf B103},  115
10118:   (1997).
10119: 
10120: \bibitem{abu96}
10121: A. Buchleitner, D. Delande, and J. Zakrzewski,  in {\em Multiphoton Processes
10122:   1996, Proceedings of the 7th International Conference on Multiphoton
10123:   Processes, Garmisch-Partenkirchen, Germany, October 1996}, Vol.~154 of {\em
10124:   Institute of Physics Conference Series}, edited by P. Lambropoulos and H.
10125:   Walther (Institute of Physics, Bristol and Philadelphia, 1997), pp.\ 19--28.
10126: 
10127: \bibitem{kuba98}
10128: J. Zakrzewski, D. Delande, and A. Buchleitner, Phys. Rev. {\bf E57},  1458
10129:   (1998).
10130: 
10131: \bibitem{kuba98a}
10132: J. Zakrzewski, D. Delande, and A. Buchleitner, Acta Physica Polon. {\bf A 93},
10133:   179  (1998).
10134: 
10135: \bibitem{abu98a}
10136: A. Buchleitner, K. Sacha, D. Delande, and J. Zakrzewski, Eur. Phys. J. D {\bf
10137:   5},  145  (1999).
10138: 
10139: \bibitem{delande98}
10140: D. Delande and J. Zakrzewski, \pr A {\bf 58},  466  (1998).
10141: 
10142: \bibitem{sachath}
10143: K. Sacha, Ph.D. thesis, Jagellonian University, Krak\'ow, 1998, unpublished.
10144: 
10145: \bibitem{hornbergerda}
10146: K. Hornberger, Master's thesis, Ludwig-Maximilians-Universit\"at, M\"un\-chen,
10147:   1998.
10148: 
10149: \bibitem{hornberger98}
10150: K. Hornberger and A. Buchleitner, Europhys. Lett. {\bf 41},  383  (1998).
10151: 
10152: \bibitem{sacha98a}
10153: K. Sacha, J. Zakrzewski, and D. Delande, Eur. Phys. J. {\bf D1},  231  (1998).
10154: 
10155: \bibitem{sacha98b}
10156: K. Sacha and J. Zakrzewski, \pr A {\bf 58},  3974  (1998).
10157: 
10158: \bibitem{sacha99a}
10159: K. Sacha and J. Zakrzewski, \pr A {\bf 59},  1707  (1999).
10160: 
10161: \bibitem{yeazell00}
10162: D. Farrelly,  in {\em Physics and Chemistry of Wave Packets}, edited by J.~A.
10163:   Yeazell and T. Uzer (John Wiley \& Sons, New York, 2000).
10164: 
10165: \bibitem{Berry_WKB}
10166: M.~V. Berry and K.~E. Mount, Rep. Prog. Phys. {\bf 35},  315  (1972).
10167: 
10168: \bibitem{ozorio88}
10169: A.~M.~O. de~Almeida, {\em Hamiltonian Systems: Chaos and Quantization}
10170:   (Cambridge University Press, Cambridge, 1988).
10171: 
10172: \bibitem{einstein17}
10173: A. Einstein, Verh. d. Dtsch. Phys. Ges.  82  (1917).
10174: 
10175: \bibitem{bornwolf}
10176: M. Born and E. Wolf, {\em Principles of Optics} (Academic Press, Oxford, 1999).
10177: 
10178: \bibitem{breuer91}
10179: H.~P. Breuer and M. Holthaus, \anp {\bf 211},  249  (1991).
10180: 
10181: \bibitem{heller89}
10182: E.~J. Heller,  in {\em Chaos and Quantum Physics}, Vol.~Session LII of {\em Les
10183:   Houches} (North-Holland, Amsterdam, 1991), p.\ 547.
10184: 
10185: \bibitem{stoeckmann99}
10186: H.~J. St\"ockmann, {\em Quantum Chaos: An Introduction} (Cambridge University
10187:   Press, Cambridge, 1999).
10188: 
10189: \bibitem{jensen89b}
10190: R.~V. Jensen, M.~M. Sanders, M. Saraceno, and B. Sundaram, \prL {\bf 63},  2771
10191:    (1989).
10192: 
10193: \bibitem{leopold94}
10194: J.~G. Leopold and D. Richards, \jpb {\bf 27},  2169  (1994).
10195: 
10196: \bibitem{zaslavsky81}
10197: G.~M. Zaslavsky, \phr {\bf 80},  157  (1981).
10198: 
10199: \bibitem{ringot00}
10200: J. Ringot, P. Szriftgiser, J.~C. Garreau, and D. Delande, \prL {\bf 85},  2741
10201:   (2000).
10202: 
10203: \bibitem{abu97}
10204: A. Buchleitner and D. Delande, Phys. Rev. {\bf A55},  R1585  (1997).
10205: 
10206: \bibitem{chirikov59}
10207: B.~V. Chirikov, Doc. Ac. Sci. USSR {\bf 125},  1015  (1959).
10208: 
10209: \bibitem{chirikov79}
10210: B.~V. Chirikov, \phr {\bf 52},  263  (1979).
10211: 
10212: \bibitem{floquet1883}
10213: M.~G. Floquet, Ann. \'Ecole Norm. Sup. {\bf 12},  47  (1883).
10214: 
10215: \bibitem{mermin76}
10216: N.~W. Ashcroft and N.~D. Mermin, {\em Solid State Physics} (Saunders College
10217:   Publishing, Fort Worth, 1976), college edition.
10218: 
10219: \bibitem{shirley65}
10220: J.~H. Shirley, \pr {\bf 138},  B979  (1965).
10221: 
10222: \bibitem{marion}
10223: J.~B. Marion, {\em Classical Dynamics of Particles and Systems}, 2nd ed.
10224:   (Academic Press, New York, 1970).
10225: 
10226: \bibitem{loudon}
10227: R. Loudon, {\em The Quantum Theory of Light}, 2nd ed. (Clarendon Press, Oxford,
10228:   1986).
10229: 
10230: \bibitem{abramowitz72}
10231:  in {\em Handbook of Mathematical Functions}, edited by M. Abrammowitz and
10232:   I.~A. Stegun (Dover, New York, 1972).
10233: 
10234: \bibitem{shakeshaft88}
10235: R. Shakeshaft, Z. Phys. D {\bf 8},  47  (1988).
10236: 
10237: \bibitem{abu95c}
10238: A. Buchleitner, D. Delande, and J.~C. Gay, \josab {\bf 12},  505  (1995).
10239: 
10240: \bibitem{cormier96}
10241: E. Cormier and P. Lambropoulos, J. Phys. B {\bf 29},  1667  (1996).
10242: 
10243: \bibitem{javanainen88}
10244: J. Javanainen, J.~H. Eberly, and Q. Su, \pr {\bf A38},  3430  (1988).
10245: 
10246: \bibitem{su91}
10247: Q. Su and J.~H. Eberly, \pr {\bf A44},  5997  (1991).
10248: 
10249: \bibitem{reed91}
10250: V.~C. Reed, P.~L. Knight, and K. Burnett, \prL {\bf 67},  1415  (1991).
10251: 
10252: \bibitem{burnett92}
10253: K. Burnett, V.~C. Reed, J. Cooper, and P.~L. Knight, \pr {\bf A45},  3347
10254:   (1992).
10255: 
10256: \bibitem{grobe93}
10257: R. Grobe and J.~H. Eberly, \pr {\bf A47},  R1605  (1993).
10258: 
10259: \bibitem{grobe94}
10260: R. Grobe, K. Rz\c{a}\.zewski, and J.~H. Eberly, J. Phys. B {\bf 27},  L503
10261:   (1994).
10262: 
10263: \bibitem{kusta65}
10264: P. Kustaanheimo and E. Stiefel, J. Reine Angew. Math. {\bf 218},  204  (1965).
10265: 
10266: \bibitem{delande84}
10267: D. Delande and J.-C. Gay, \jpb {\bf 17},  335  (1984).
10268: 
10269: \bibitem{rath88}
10270: O. Rath and D. Richards, \jpb {\bf 21},  555  (1988).
10271: 
10272: \bibitem{griffiths92}
10273: J.~A. Griffiths and D. Farrelly, \pr {\bf A45},  R2678  (1992).
10274: 
10275: \bibitem{gebarowski95}
10276: R. G{\c{e}}barowski and J. Zakrzewski, \pr {\bf A51},  1508  (1995).
10277: 
10278: \bibitem{schlagheck99}
10279: P. Schlagheck and A. Buchleitner, \phd {\bf 131},  110  (1999).
10280: 
10281: \bibitem{barut79}
10282: A.~O. Barut, C.~K.~E. Schneider, and R. Wilson, J.\ Math.\ Phys.\ {\bf 20},
10283:   2244  (1979).
10284: 
10285: \bibitem{chen80}
10286: A.~C. Chen, \pr {\bf A 22},  333  (1980).
10287: 
10288: \bibitem{chen81}
10289: A.~C. Chen, \pr {\bf A 23},  1655  (1981).
10290: 
10291: \bibitem{delandeth}
10292: D. Delande, Ph.D. thesis, Universit\'e de Paris, Paris, 1988, th\`ese de
10293:   doctorat d'etat.
10294: 
10295: \bibitem{lanczos}
10296: C. Lanczos, J. Res. Nat. Bur. Standards, Sect B {\bf 45},  225  (1950).
10297: 
10298: \bibitem{delande91}
10299: D. Delande, A. Bommier, and J.-C. Gay, \prL {\bf 66},  141  (1991).
10300: 
10301: \bibitem{ericsson80}
10302: T. Ericsson and A. Ruhe, Math. Comput. {\bf 35},  1251  (1980).
10303: 
10304: \bibitem{grimes94}
10305: R.~G. Grimes, J.~G. Lewis, and H.~D. Simon, SIAM J. Matrix Anal. Appl. {\bf
10306:   15},  228  (1994).
10307: 
10308: \bibitem{balslev71}
10309: E. Balslev and J.~M. Combes, Commun.\ math.\ Phys. {\bf 22},  280  (1971).
10310: 
10311: \bibitem{graffi85}
10312: S. Graffi, V. Grecchi, and H.~J. Silverstone, Ann.\ Inst.\ Henri Poincar\'e
10313:   {\bf 42},  215  (1985).
10314: 
10315: \bibitem{yajima82}
10316: K. Yajima, Comm.\ Math.\ Phys. {\bf 87},  331  (1982).
10317: 
10318: \bibitem{nicolaides78}
10319: C.~A. Nicolaides and D.~R. Beck, Int. J. Quant. Chem. {\bf XIV},  457  (1978).
10320: 
10321: \bibitem{reinhardt83}
10322: B.~R. Johnson and W.~P. Reinhardt, Phys. Rev. A {\bf 28},  1930  (1983).
10323: 
10324: \bibitem{ho83}
10325: Y.~K. Ho, Phys. Rep. {\bf 99},  1  (1983).
10326: 
10327: \bibitem{moiseyev98}
10328: N. Moiseyev, Phys. Rep. {\bf 302},  211  (1998).
10329: 
10330: \bibitem{abu94}
10331: A. Buchleitner, B. Gr\'emaud, and D. Delande, \jpb {\bf 27},  2663  (1994).
10332: 
10333: \bibitem{englefield72}
10334: M.~J. Englefield, {\em Group Theory and the Coulomb Problem} (Wiley, New-York,
10335:   1972).
10336: 
10337: \bibitem{goldstein80}
10338: H. Goldstein, {\em Classical Dynamics} (Addison-Wesley, Reading, Ma., 1980),
10339:   p.146.
10340: 
10341: \bibitem{jensen84}
10342: R.~V. Jensen, \pr {\bf A30},  386  (1984).
10343: 
10344: \bibitem{meerson82}
10345: B.~I. Meerson, E.~A. Oks, and P.~V. Sasorov, \jpb {\bf 15},  3599  (1982).
10346: 
10347: \bibitem{casati88}
10348: G. Casati, I. Guarneri, and D.~L. Shepelyansky, IEEE J.\ Quantum Electron. {\bf
10349:   24},  1420  (1988).
10350: 
10351: \bibitem{bayfield89}
10352: J.~E. Bayfield, G. Casati, I. Guarneri, and D.~W. Sokol, \prL {\bf 63},  364
10353:   (1989).
10354: 
10355: \bibitem{koch95b}
10356: P.~M. Koch and K.~A.~H. van Leeuwen, \phr {\bf 255},  289  (1995).
10357: 
10358: \bibitem{bellermann96}
10359: M.~R.~W. Bellermann, P.~M. Koch, D. Mariani, and D. Richards, \prL {\bf 76},
10360:   892  (1996).
10361: 
10362: \bibitem{bayfield74}
10363: J.~E. Bayfield and P.~M. Koch, \prL {\bf 33},  258  (1974).
10364: 
10365: \bibitem{bayfield96}
10366: J.~E. Bayfield, S.~Y. Luie, L.~C. Perotti, and M.~P. Skrzypkowski, \pr {\bf
10367:   A53},  R12  (1996).
10368: 
10369: \bibitem{galvez88}
10370: E.~J. Galvez {\it et~al.}, \prL {\bf 61},  2011  (1988).
10371: 
10372: \bibitem{fu90}
10373: P. Fu, T.~J. Scholz, J.~M. Hettema, and T.~F. Gallagher, \prL {\bf 64},  511
10374:   (1990).
10375: 
10376: \bibitem{cheng96}
10377: C.~H. Cheng, C.~Y. Lee, and T.~F. Gallagher, \pr {\bf A54},  3303  (1996).
10378: 
10379: \bibitem{delande97b}
10380: D. Delande and J. Zakrzewski,  in {\em Classical, Semiclassical and Quantum
10381:   Dynamics in Atoms}, No.~485 in {\em Lecture Notes in Physics}, edited by H.
10382:   Friedrich and B. Eckhardt (Springer, New York, 1997), p.\ 205.
10383: 
10384: \bibitem{bunkin64}
10385: F.~V. Bunkin and A.~M. Prokhorov, Sov.\ Phys.\ JETP {\bf 19},  739  (1964).
10386: 
10387: \bibitem{grozdanov92}
10388: T.~P. Grozdanov, M.~J. Rakovi\'c, and E.~A. Solovev, \jpb {\bf 25},  4455
10389:   (1992).
10390: 
10391: \bibitem{klar89}
10392: H. Klar, \zpd {\bf 11},  45  (1989).
10393: 
10394: \bibitem{lee95}
10395: E. Lee, A.~F. Brunello, and D. Farrelly, \prL {\bf 75},  3641  (1995).
10396: 
10397: \bibitem{abu95a}
10398: A. Buchleitner and D. Delande, \csf {\bf 5},  1125  (1995).
10399: 
10400: \bibitem{heller84}
10401: E.~J. Heller,  in {\em Classical and Quantum Chaos}, edited by A. Voros, M.
10402:   Giannoni, and A. Zinn-Justin (Elsevier, Amsterdam, 1991), Chap.~Scars.
10403: 
10404: \bibitem{bogomolny88a}
10405: E.~B. Bogomolny, \phd {\bf 31},  169  (1988).
10406: 
10407: \bibitem{Bth}
10408: A. Brunello, Ph.D. thesis, State University of New York at Stony Brook, Stony
10409:   Brook, 1997, unpublished.
10410: 
10411: \bibitem{sacha97}
10412: K. Sacha and J. Zakrzewski, \pr {\bf A56},  719  (1997).
10413: 
10414: \bibitem{sacha98c}
10415: K. Sacha and J. Zakrzewski, \pr A {\bf 58},  488  (1998).
10416: 
10417: \bibitem{brunello97}
10418: A.~F. Brunello, T. Uzer, and D. Farrelly, \pr A {\bf 55},  3730  (1997).
10419: 
10420: \bibitem{leopold86}
10421: J.~G. Leopold and D. Richards, \jpb {\bf 19},  1125  (1986).
10422: 
10423: \bibitem{leopold87}
10424: J.~G. Leopold and D. Richards, \jpb {\bf 20},  2369  (1987).
10425: 
10426: \bibitem{rakovic98}
10427: M.~J. Rakovi\'c, T. Uzer, and D. Farrelly, \pr A {\bf 57},  2814  (1998).
10428: 
10429: \bibitem{benvenuto91}
10430: F. Benvenuto, G. Casati, I. Guarneri, and D.~L. Shepelyansky, \zp B {\bf 84},
10431:   159  (1991).
10432: 
10433: \bibitem{oliveira94}
10434: C.~R. de~Oliveira, G. Casati, and I. Guarneri, \epl {\bf 27},  187  (1994).
10435: 
10436: \bibitem{flatte96}
10437: M.~E. Flatt\'e and M. Holthaus, \anp {\bf 245},  113  (1996).
10438: 
10439: \bibitem{steane95}
10440: A. Steane, P. Szriftgiser, P. Desbiolles, and J. Dalibard, \prL {\bf 74},  4972
10441:    (1995).
10442: 
10443: \bibitem{oberthaler99}
10444: M.~K. Oberthaler {\it et~al.}, \prL  4447  (1999).
10445: 
10446: \bibitem{bonci98}
10447: L. Bonci, A. Farusi, P. Grigilini, and R. Roncaglia, \pr E {\bf 58},  5689
10448:   (1998).
10449: 
10450: \bibitem{brodier01}
10451: O. Brodier, P. Schlagheck, and D. Ullmo, \prL {\bf 87},  64101  (2001).
10452: 
10453: \bibitem{paul92}
10454: W. Paul, Rev. Mod. Phys. {\bf 62},  531  (1992).
10455: 
10456: \bibitem{benvenuto94}
10457: F. Benvenuto, G. Casati, and D.~L. Shepelyansky, \prL {\bf 72},  1818  (1994).
10458: 
10459: \bibitem{tanner00}
10460: G. Tanner, K. Richter, and J.~M. Rost, \rmp {\bf 72},  497  (2000).
10461: 
10462: \bibitem{wintgen93}
10463: D. Wintgen and D. Delande, \jpb {\bf 26},  L399  (1993).
10464: 
10465: \bibitem{gremaud97}
10466: B. Gr\'emaud and D. Delande, \epl {\bf 40},  363  (1997).
10467: 
10468: \bibitem{gremaudth}
10469: B. Gr\'emaud, Ph.D. thesis, Universit\'e Paris 6, 1997.
10470: 
10471: \bibitem{puttner01}
10472: R. P\"uttner {\it et~al.}, \prL {\bf 86},  3747  (2001).
10473: 
10474: \bibitem{eichmann90}
10475: U. Eichmann, V. Lange, and W. Sandner, \prL {\bf 64},  274  (1990).
10476: 
10477: \bibitem{richter90}
10478: K. Richter and D. Wintgen, \prL {\bf 65},  1965  (1990).
10479: 
10480: \bibitem{schlagheck98a}
10481: P. Schlagheck and A. Buchleitner, \jpb {\bf 31},  L489  (1998).
10482: 
10483: \bibitem{schlagheck99b}
10484: P. Schlagheck and A. Buchleitner, \epl {\bf 46},  24  (1999).
10485: 
10486: \bibitem{schlagheckth}
10487: P. Schlagheck, Ph.D. thesis, Technische Universit\"at, M\"unchen, 1999,
10488:   published by Herbert Utz Verlag.
10489: 
10490: \bibitem{hanson95}
10491: L.~G. Hanson and P. Lambropoulos, \prL {\bf 74},  5009  (1995).
10492: 
10493: \bibitem{zobay96}
10494: O. Zobay and G. Alber, \pr {\bf 54},  5361  (1996).
10495: 
10496: \bibitem{mecking98}
10497: B. Mecking and P. Lambropoulos, \pr {\bf A57},  2014  (1998).
10498: 
10499: \bibitem{carpetmen}
10500: M. Berry, I. Marzoli, and W.~P. Schleich, Physics World {\bf 14},    (2001).
10501: 
10502: \bibitem{stone85}
10503: P.~A. Lee and A.~D. Stone, \prL {\bf 55},  1622  (1985).
10504: 
10505: \bibitem{LB90}
10506: W.~A. Lin and L.~E. Ballentine, \prL {\bf 65},  2927  (1990).
10507: 
10508: \bibitem{LB92}
10509: W.~A. Lin and L.~E. Ballentine, \pr {\bf 45},  3637  (1992).
10510: 
10511: \bibitem{grossmann91}
10512: F. Grossmann, T. Dittrich, P. Jung, and P. H\"anggi, \prL {\bf 67},  516
10513:   (1991).
10514: 
10515: \bibitem{GDJH91b}
10516: F. Grossmann, T. Dittrich, P. Jung, and P. H\"anggi, Z.\ Phys.\ B\ {\bf 84},
10517:   315  (1991).
10518: 
10519: \bibitem{GDJH93}
10520: F. Grossmann {\it et~al.}, J.\ Stat.\ Phys.\ {\bf 70},  229  (1993).
10521: 
10522: \bibitem{plata92}
10523: J. Plata and J.~M.~G. Llorente, \jpa {\bf 25},  L303  (1992).
10524: 
10525: \bibitem{bohigas93}
10526: O. Bohigas, S. Tomsovic, and D. Ullmo, \phr {\bf 223},  43  (1993).
10527: 
10528: \bibitem{bohigas93b}
10529: O. Bohigas, D. Boos\'e, R.~E. de~Carvalho, and V. Marvulle, Nucl.\ Phys.\ {\bf
10530:   A560},  197  (1993).
10531: 
10532: \bibitem{tomsovic94}
10533: S. Tomsovic and D. Ullmo, \pr {\bf E50},  145  (1994).
10534: 
10535: \bibitem{shudo95}
10536: A. Shudo and K.~S. Ikeda, \prL {\bf 74},  682  (1995).
10537: 
10538: \bibitem{leyvraz96}
10539: F. Leyvraz and D. Ullmo, \jpa {\bf 29},  2529  (1996).
10540: 
10541: \bibitem{jackson}
10542: J.~D. Jackson, {\em Classical Electrodynamics} (Wiley, New York, 1975).
10543: 
10544: \bibitem{haroche92}
10545: S. Haroche,  in {\em Fundamental Systems in Quantum Optics, Le Houches, Session
10546:   LIII,1990}, edited by J. Dalibard, J.~M. Raimond, and J. Zinn-Justin
10547:   (Elsevier, New York, 1992).
10548: 
10549: \bibitem{lewenstein88}
10550: M. Lewenstein, J. Zakrzewski, T.~W. Mossberg, and J. Mostowski, \jpb {\bf 21},
10551:   L9  (1988).
10552: 
10553: \bibitem{lewenstein88b}
10554: M. Lewenstein, J. Zakrzewski, and T.~W. Mossberg, \pr {\bf A38},  808  (1988).
10555: 
10556: \bibitem{abu97b}
10557: D.~D.~R. Buchleitner, D. Delande, and J. Zakrzewski,  in {\em Non-spreading
10558:   Kn\"odel-packets in Microsoft fields, Proceedings of the Third International
10559:   Bavarian Conference on Kn\"odel-Packets}, No.~007 in {\em Food, physics and
10560:   politics}, edited by H. Schalther and W. Weich (Stoiber Comp., Aidling,
10561:   1997), p.\ 294561.
10562: 
10563: \bibitem{nakamura93}
10564: K. Nakamura, {\em Quantum chaos, a new paradigm of nonlinear dynamics}, {\em
10565:   Cambridge Nonlinear Science Series 3} (Cambridge University Press, Cambridge,
10566:   1993).
10567: 
10568: \bibitem{casati87b}
10569: G. Casati, I. Guarneri, and D.~L. Shepelyansky, \pr {\bf A36},  3501  (1987).
10570: 
10571: \bibitem{leopold78}
10572: J.~G. Leopold and I.~C. Percival, \prL {\bf 41},  944  (1978).
10573: 
10574: \bibitem{bayfield88a}
10575: J.~E. Bayfield and D.~W. Sokol, \prL {\bf 61},  2007  (1988).
10576: 
10577: \bibitem{bluemel89a}
10578: R. Bl\"umel {\it et~al.}, \prL {\bf 62},  341  (1989).
10579: 
10580: \bibitem{bluemel91a}
10581: R. Bl\"umel {\it et~al.}, Phys. Rev. {\bf A44},  4521  (1991).
10582: 
10583: \bibitem{abu91}
10584: A. Buchleitner, L. Sirko, and H. Walther, \epl {\bf 16},  35  (1991).
10585: 
10586: \bibitem{casati84}
10587: G. Casati, B.~V. Chirikov, and D.~L. Shepelyansky, \prL {\bf 53},  2525
10588:   (1984).
10589: 
10590: \bibitem{fishman82}
10591: S. Fishman, D.~R. Grempel, and R.~E. Prange, \prL {\bf 49},  509  (1982).
10592: 
10593: \bibitem{grempel84}
10594: D.~R. Grempel, R.~E. Prange, and S. Fishman, \pr {\bf A29},  1639  (1984).
10595: 
10596: \bibitem{brenner96}
10597: N. Brenner and S. Fishman, \prL {\bf 77},  3763  (1996).
10598: 
10599: \bibitem{abu98b}
10600: A. Buchleitner, I. Guarneri, and J. Zakrzewski, Europhys. Lett. {\bf 44},  162
10601:   (1998).
10602: 
10603: \bibitem{wimbergerda}
10604: S. Wimberger, Master's thesis, Ludwig-Maximilians-Universit\"at M\"unchen,
10605:   2000.
10606: 
10607: \bibitem{wimberger01}
10608: S. Wimberger and A. Buchleitner, \jpa {\bf 34},  7181  (2001).
10609: 
10610: \bibitem{sirko96}
10611: L. Sirko, A. Haffmans, M.~R.~W. Bellermann, and P.~M. Koch, \epl {\bf 33},  181
10612:    (1996).
10613: 
10614: \bibitem{sirko93a}
10615: L. Sirko {\it et~al.}, \prL {\bf 71},  2895  (1993).
10616: 
10617: \bibitem{sauer92b}
10618: B.~E. Sauer, M.~R.~W. Bellermann, and P.~M. Koch, \prL {\bf 68},  1633  (1992).
10619: 
10620: \bibitem{leeuwen85}
10621: K.~A.~H. van Leeuwen {\it et~al.}, \prL {\bf 55},  2231  (1985).
10622: 
10623: \bibitem{koch95a}
10624: P.~M. Koch, Physica D {\bf 83},  178  (1995).
10625: 
10626: \bibitem{koch92b}
10627: P.~M. Koch,  in {\em Chaos and Quantum Chaos}, Vol.~411 of {\em Lecture Notes
10628:   in Physics}, edited by W.~D. Heiss (Springer, Berlin, 1992).
10629: 
10630: \bibitem{koch92a}
10631: P.~M. Koch, \ch {\bf 2},  131  (1992).
10632: 
10633: \bibitem{koch89}
10634: P.~M. Koch {\it et~al.}, \psc {\bf T26},  51  (1989).
10635: 
10636: \bibitem{noel00}
10637: M.~W. No\"el, M.~W. Griffith, and T.~F. Gallagher, \pr {\bf A62},  063401
10638:   (2000).
10639: 
10640: \bibitem{bayfield91}
10641: J.~E. Bayfield, \ch {\bf 1},  110  (1991).
10642: 
10643: \bibitem{kuba95b}
10644: J. Zakrzewski and D. Delande, J. Phys. B {\bf 28},  L667  (1995).
10645: 
10646: \bibitem{abu95b}
10647: A. Buchleitner {\it et~al.}, Phys. Rev. Lett. {\bf 75},  3818  (1995).
10648: 
10649: \bibitem{arndt91}
10650: M. Arndt, A. Buchleitner, R.~N. Mantegna, and H. Walther, \prL {\bf 67},  2435
10651:   (1991).
10652: 
10653: \bibitem{benson95}
10654: O. Benson {\it et~al.}, \pr {\bf A51},  4862  (1995).
10655: 
10656: \bibitem{noel00b}
10657: M.~W. No\"el, M.~W. Griffith, and T.~F. Gallagher, \prL {\bf 87},  043001
10658:   (2001).
10659: 
10660: \bibitem{bluemel90a}
10661: R. Bl\"umel and U. Smilansky, \josab {\bf 7},  664  (1990).
10662: 
10663: \bibitem{krug00}
10664: A. Krug and A. Buchleitner, \epl {\bf 49},  176  (2000).
10665: 
10666: \bibitem{krug01}
10667: A. Krug and A. Buchleitner, \prL {\bf 86},  3538  (2001).
10668: 
10669: \bibitem{breuer89b}
10670: H.~P. Breuer and M. Holthaus, \pla {\bf 140},  507  (1989).
10671: 
10672: \bibitem{bayfield85}
10673: J.~E. Bayfield and L.~A. Pinnaduwage, \prL {\bf 54},  313  (1985).
10674: 
10675: \bibitem{reinhold01}
10676: C.~O. Reinhold {\it et~al.}, \jpb {\bf 34},  L551  (2001).
10677: 
10678: \bibitem{gerber98}
10679: A. Assion {\it et~al.}, Science {\bf 282},  919  (1998).
10680: 
10681: \bibitem{gerber01}
10682: T. Brixner, N.~H. Damrauer, P. Niklaus, and G. Gerber, Nature {\bf 414},  57
10683:   (2001).
10684: 
10685: \end{thebibliography}
10686: 
10687: \end{document}
10688: