1: \documentclass[reqno,11pt]{article}
2: \usepackage{psfig, amsmath, amstexnb, amsthm}
3:
4: \textheight=22.8cm
5: \textwidth=15.0cm
6: %\textheight=22.4cm
7: %\textwidth=14.95cm
8: \topmargin=-3mm
9: \oddsidemargin=5mm
10: \evensidemargin=5mm
11:
12: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
13: %
14: % Environments
15: %
16: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
17:
18: \newtheorem{theorem}{Theorem}[section]
19:
20: \makeatletter
21:
22: \def\enum{\ifnum \@enumdepth >3 \@toodeep\else
23: \advance\@enumdepth \@ne
24: \edef\@enumctr{enum\romannumeral\the\@enumdepth}\list
25: {\csname label\@enumctr\endcsname}
26: {\setlength{\topsep}{1mm}
27: \setlength{\parsep}{0mm}
28: \setlength{\itemsep}{0mm}
29: \setlength{\labelsep}{2mm}
30: \settowidth{\leftmargin}{M.}
31: \addtolength{\leftmargin}{\labelsep}
32: \usecounter{\@enumctr}
33: \def\makelabel##1{\hss\llap{##1}}}\fi}
34:
35: \let\endenum =\endlist
36:
37: \def\itemiz{\ifnum \@itemdepth >3 \@toodeep\else \advance\@itemdepth \@ne
38: \edef\@itemitem{labelitem\romannumeral\the\@itemdepth}%
39: \list{\csname\@itemitem\endcsname}{
40: \setlength{\topsep}{1mm}
41: \setlength{\parsep}{0mm}
42: \setlength{\itemsep}{0mm}
43: \setlength{\labelsep}{2mm}
44: \settowidth{\leftmargin}{M.}
45: \addtolength{\leftmargin}{\labelsep}
46: \def\makelabel##1{\hss\llap{##1}}}\fi}
47:
48: \let\enditemiz =\endlist
49:
50: \def\captionheadfont@{\scshape}
51: \def\captionfont@{\small}
52: \long\def\@makecaption#1#2{%
53: \setbox\@tempboxa\vbox{\color@setgroup
54: \advance\hsize-3pc\noindent
55: \captionfont@\captionheadfont@#1\@xp\@ifnotempty\@xp
56: {\@cdr#2\@nil}{.\captionfont@\upshape\enspace#2}%
57: \unskip\kern-3pc\par
58: \global\setbox\@ne\lastbox\color@endgroup}%
59: \ifhbox\@ne % the normal case
60: \setbox\@ne\hbox{\unhbox\@ne\unskip\unskip\unpenalty\unkern}%
61: \fi
62: \ifdim\wd\@tempboxa=\z@ % this means caption will fit on one line
63: \setbox\@ne\hbox to\columnwidth{\hss\kern-3pc\box\@ne\hss}%
64: \else % tempboxa contained more than one line
65: \setbox\@ne\vbox{\unvbox\@tempboxa\parskip\z@skip
66: \noindent\unhbox\@ne\advance\hsize-3pc\par}%
67: \fi
68: \ifnum\@tempcnta<64 % if the float IS a figure...
69: \addvspace\abovecaptionskip
70: \moveright 1.5pc\box\@ne
71: \else % if the float IS NOT a figure...
72: \moveright 1.5pc\box\@ne
73: \nobreak
74: \vskip\belowcaptionskip
75: \fi
76: \relax
77: }
78:
79: \makeatother
80:
81: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
82: %
83: % Macros
84: %
85: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
86:
87: \DeclareMathSymbol{\leqsymb}{\mathalpha}{AMSa}{"36}
88: \def\leqs{\;\leqsymb\;}
89: \DeclareMathSymbol{\geqsymb}{\mathalpha}{AMSa}{"3E}
90: \def\geqs{\;\geqsymb\;}
91: \DeclareMathSymbol{\gtreqqlesssymb}{\mathalpha}{AMSa}{"54}
92: \def\geql{\;\text{{\footnotesize $\gtreqqlesssymb$}}\;}
93:
94: \newcommand{\field}[1]{\mathbb{#1}}
95:
96: \newcommand{\N}{\field{N}\,} % natural integers
97: \newcommand{\Z}{\field{Z}\,} % integers
98: \newcommand{\Q}{\field{Q}\,} % rational numbers
99: \newcommand{\R}{\field{R}\,} % real numbers
100: \newcommand{\C}{\field{C}\,} % complex numbers
101: \newcommand{\E}{\field{E}\,} % expectation
102: \newcommand{\K}{\field{K}\,} % field
103: \newcommand{\M}{\field{M}\,} % matrices
104: \newcommand{\fP}{\field{P}\,} % probability
105: \newcommand{\A}{\field{A}\,} % algebra
106: \newcommand{\fS}{\field{S}\,} % sphere
107: \newcommand{\T}{\field{T}\,} % torus
108:
109: \newcommand{\cA}{{\mathcal A}} % calligraphic A
110: \newcommand{\cB}{{\mathcal B}} % calligraphic B
111: \newcommand{\cC}{{\mathcal C}} % calligraphic C
112: \newcommand{\cD}{{\mathcal D}} % calligraphic D
113: \newcommand{\cE}{{\mathcal E}} % calligraphic E
114: \newcommand{\cF}{{\mathcal F}} % calligraphic F
115: \newcommand{\cG}{{\mathcal G}} % calligraphic G
116: \newcommand{\cH}{{\mathcal H}} % calligraphic H
117: \newcommand{\cL}{{\mathcal L}} % calligraphic L
118: \newcommand{\cM}{{\mathcal M}} % calligraphic M
119: \newcommand{\cN}{{\mathcal N}} % calligraphic N
120: \newcommand{\cR}{{\mathcal R}} % calligraphic N
121: \newcommand{\cS}{{\mathcal S}} % calligraphic S
122: \newcommand{\cU}{{\mathcal U}} % calligraphic U
123: \newcommand{\cW}{{\mathcal W}} % calligraphic W
124:
125: \DeclareMathOperator{\e}{e} % natural number
126: \DeclareMathOperator{\icx}{i} % imaginary number
127: \DeclareMathOperator{\re}{Re} % real part
128: \DeclareMathOperator{\im}{Im} % imaginary part
129: \DeclareMathOperator{\tg}{tan} % tangent
130: \DeclareMathOperator{\thyp}{tanh} % hyperbolic tangent
131: \DeclareMathOperator{\chyp}{cosh} % hyperbolic cosine
132: \DeclareMathOperator{\shyp}{sinh} % hyperbolic sine
133: \DeclareMathOperator{\Arctg}{atan} % Arc tangent
134: \DeclareMathOperator{\Arccos}{acos} % Arc cosine
135: \DeclareMathOperator{\Arcsin}{asin} % Arc sine
136: \DeclareMathOperator{\Argth}{atanh} % Arg hyp tangent
137: \DeclareMathOperator{\dd}{d} % integration measure
138: \DeclareMathOperator{\diag}{diag} % diagonal matrix
139: \DeclareMathOperator{\Tr}{Tr} % trace
140: \DeclareMathOperator{\Ei}{Ei} % exponential integral
141: \DeclareMathOperator{\Ai}{Ai} % Airy function
142: \DeclareMathOperator{\Bi}{Bi} % Airy function
143: \DeclareMathOperator{\sign}{sign} % sign
144: \DeclareMathOperator{\interior}{int} % interior
145: %\DeclareMathOperator{\defby}{\raisebox{0.35pt}{\math{:}}\!\!=}
146: %\DeclareMathOperator{\bydef}{=\!\!\raisebox{0.35pt}{\math{:}}}
147:
148: \def\vec#1{\ifmmode
149: \mathchoice{\mbox{\boldmath$\displaystyle\bf#1$}}
150: {\mbox{\boldmath$\textstyle\bf#1$}}
151: {\mbox{\boldmath$\scriptstyle\bf#1$}}
152: {\mbox{\boldmath$\scriptscriptstyle\bf#1$}}\else
153: {\mbox{\boldmath$\bf#1$}}\fi}
154:
155: \def\math#1{\ifmmode
156: \mathchoice{\mbox{$\displaystyle\rm#1$}}
157: {\mbox{$\textstyle\rm#1$}}
158: {\mbox{$\scriptstyle\rm#1$}}
159: {\mbox{$\scriptscriptstyle\rm#1$}}\else
160: {\mbox{$\rm#1$}}\fi} % roman style with adaptable size
161:
162: \def\eps{\varepsilon}
163: \def\w{\omega}
164: \def\W{\Omega}
165: \def\x{\xi}
166: \def\y{\eta}
167: \def\z{\zeta}
168: \def\ph{\varphi}
169: \def\th{\theta}
170: \def\t{t}
171: \def\Hbar{\overline{H}}
172:
173: \def\we{w_{\math e}}
174: \def\Je{J_{\math e}}
175: \def\wL{w_\Lambda}
176: \def\wK{w_K}
177:
178: \def\defwd#1{{\sl #1}} % defined word
179:
180: \def\dx#1{\dd\!#1} % integration measure
181: \def\dpar#1#2{\frac{\partial #1}{\partial #2}} % partial derivative
182: \def\tdpar#1#2{\partial #1/\partial #2} % text partial derivative
183: \def\sdpar#1#2{\partial_{#2}#1} % small partial derivative
184: \def\dtot#1#2{\frac{\dx{#1}}{\dx{#2}}} % total derivative
185: \def\tdtot#1#2{\dx{#1}/\!\dx{#2}} % text total derivative
186: \def\sdtot#1#2{\dd\!_{#2}#1} % small total derivative
187: \def\poisson#1#2{\{#1;#2\}} % Poisson bracket
188:
189: \def\brak#1{[#1]} % bracket
190: \def\set#1{\{#1\}} % set
191: \def\abs#1{\lvert#1\rvert} % absolute value
192: \def\intpart#1{[#1]} % integer part
193: \def\cc#1{\overline{#1}} % complex conjugate
194: \def\norm#1{\lVert#1\rVert} % norm
195: \def\setsuch#1#2{\{#1\,:\,#2\}} % set of #1 such that #2
196: \def\Order#1{{\mathcal O}(#1)} % Order relation
197: \def\order#1{{\scriptstyle\mathcal O}(#1)} % order relation -- ?
198: \def\prob#1{\fP\{#1\}} % probability
199: \def\expec#1{\E\{#1\}} % expectation
200: \def\pcond#1#2{\fP\{#1\vert#2\}} % conditional probability
201: \def\pscal#1#2{\langle#1\vert#2\rangle} % scalar product
202:
203: \def\bigbrak#1{\bigl[#1\bigr]} % big bracket
204: \def\bigpar#1{\bigl(#1\bigr)} % big paranthesis
205: \def\bigabs#1{\bigl|#1\bigr|} % big absolute value
206: \def\bigset#1{\bigl\{#1\bigr\}} % big set
207: \def\bigsetsuch#1#2{\bigl\{#1\,:\,#2\bigr\}}
208: % set of #1 such that #2
209: \def\bigOrder#1{{\mathcal O}\bigl(#1\bigr)} % Order relation
210: \def\bigprob#1{\fP\bigl\{#1\bigr\}} % probability
211: \def\bigexpec#1{\E\bigl\{#1\bigr\}} % expectation
212: \def\bigpcond#1#2{\fP\bigl\{#1\,\bigr|\bigl.\,#2\bigr\}}
213: % conditional probability
214:
215: \def\Bigbrak#1{\Bigl[#1\Bigr]} % Big bracket
216: \def\Bigpar#1{\Bigl(#1\Bigr)} % Big paranthesis
217: \def\Bigabs#1{\Bigl|#1\Bigr|} % Big absolute value
218: \def\Bigset#1{\Bigl\{#1\Bigr\}} % Big set
219: \def\Bigsetsuch#1#2{\Bigl\{#1\,:\,#2\Bigr\}}
220: % set of #1 such that #2
221: \def\BigOrder#1{{\mathcal O}\Bigl(#1\Bigr)} % Big Order relation
222: \def\Bigorder#1{{\scriptstyle\mathcal O}\Bigl(#1\Bigr)}
223: \def\Bigevalat#1{\Bigr|_{#1}^{\phantom{#1}}} % evalued at #1
224: \def\Bigprob#1{\fP\Bigl\{#1\Bigr\}} % probability
225: \def\Bigexpec#1{\E\Bigl\{#1\Bigr\}} % expectation
226: \def\Bigpcond#1#2{\fP\Bigl\{#1\,\Bigr|\Bigl.\,#2\Bigr\}}
227: % conditional probability
228:
229: \def\biggbrak#1{\biggl[#1\biggr]} % bigg bracket
230: \def\biggpar#1{\biggl(#1\biggr)} % bigg paranthesis
231: \def\biggabs#1{\biggl|#1\biggr|} % bigg absolute value
232: \def\biggset#1{\biggl\{#1\biggr\}} % bigg set
233: \def\biggintpart#1{\biggl[#1\biggr]} % bigg integer part
234: \def\biggsetsuch#1#2{\biggl\{#1\,:\,#2\biggr\}}
235: % set of #1 such that #2
236: \def\biggOrder#1{{\mathcal O}\biggl(#1\biggr)} % bigg Order relation
237: \def\biggorder#1{{\scriptstyle\mathcal O}\biggl(#1\biggr)}
238: % bigg order relation -- ?
239: \def\biggprob#1{\fP\biggl\{#1\biggr\}} % probability
240: \def\biggexpec#1{\E\biggl\{#1\biggr\}} % expectation
241: \def\biggpcond#1#2{\fP\biggl\{#1\,\biggr|\biggl.\,#2\biggr\}}
242: % conditional probability
243:
244: \def\Biggbrak#1{\Biggl[#1\Biggr]} % Bigg bracket
245: \def\Biggabs#1{\Biggl|#1\Biggr|} % Bigg absolute value
246: \def\Biggintpart#1{\Biggl[#1\Biggr]} % Bigg integer part
247:
248: \def\mynote#1{\marginpar{\footnotesize #1}}
249: \def\avrg#1{\langle #1 \rangle}
250: \def\avvrg#1{\langle\langle #1 \rangle\rangle}
251:
252: \def\figref#1{Fig.\ \ref{#1}} % reference to a figure
253: \def\tabref#1{Table \ref{#1}} % reference to a table
254: \def\writefig#1 #2 #3 {\rlap{\kern #1 truecm
255: \raise #2 truecm \hbox{\protect{\small #3}}}}
256: \def\figtext#1{\smash{\hbox{#1}}
257: \vspace{-5mm}}
258:
259: \def\bibtitle#1#2{#1, {\em #2}} % authors, title
260: %\def\bibtitle#1#2{#1} % authors, no title
261: \def\bibref#1#2#3#4#5{#1 {\bf #2}:#3--#4 (#5)} % review, year
262: \def\bibarticle#1#2#3#4#5#6#7{\bibtitle{#1}{#2},
263: \bibref{#3}{#4}{#5}{#6}{#7}.}
264: \def\bibpreprint#1#2#3#4{#1, {\em #2}, preprint {\tt #3} (#4).}
265: \def\bibbook#1#2#3#4{#1, {\em #2} (#3, #4).}
266:
267: \def\AJM{Amer.\ J.\ Math.}
268: \def\AMS{American Mathematical Society}
269: \def\AP{Ann.\ Physics}
270: \def\CMP{Comm.\ Math.\ Phys.}
271: \def\DE{Diff.\ Equ.}
272: \def\Dokl{Dokl.\ Akad.\ Nauk SSSR}
273: \def\Doklt{Sov.\ Math.\ Dokl.}
274: \def\DU{Diff.\ Urav.\ }
275: \def\JPA{J.\ Phys.\ A}
276: \def\JPCM{J.\ Phys: Condens.\ Matter}
277: \def\JSP{J.\ Stat.\ Phys.}
278: \def\Nat{Nature}
279: \def\NL{Nonlinearity}
280: \def\PA{Physica A}
281: \def\PD{Physica D}
282: \def\PR{Phys.\ Rev.}
283: \def\PRA{Phys.\ Rev.\ A}
284: \def\PRB{Phys.\ Rev.\ B}
285: \def\PRE{Phys.\ Rev.\ E}
286: \def\PRL{Phys.\ Rev.\ Lett.}
287: \def\RMP{Rev.\ Mod.\ Phys.}
288: \def\SAM{Stud.\ in Appl.\ Math.}
289: \def\SIAM{SIAM J.\ Appl.\ Math.}
290:
291:
292: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
293:
294: \begin{document}
295:
296: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
297:
298: \title{The averaged dynamics of the hydrogen atom\\
299: in crossed electric and magnetic fields\\
300: as a perturbed Kepler problem\thanks{Dedicated to Martin C.\ Gutzwiller on
301: the occasion of his 75th birthday.}}
302: \author{Nils Berglund\thanks{Present address: Weierstra\ss\ Institut,
303: Mohrenstra\ss e 39, D-10117 Berlin, Germany} and Turgay Uzer\\
304: School of Physics,
305: Georgia Tech\\
306: Atlanta, GA 30332-0430, USA\\
307: %{\tt ph297nb@prism.gatech.edu}
308: }
309: \date{July 13, 2000}
310: \maketitle
311:
312: \begin{abstract}
313: We treat the classical dynamics of the hydrogen atom in perpendicular
314: electric and magnetic fields as a celestial mechanics problem. By
315: expressing the Hamiltonian in appropriate action--angle variables, we
316: separate the different time scales of the motion. The method of averaging
317: then allows us to reduce the system to two degrees of freedom, and to
318: classify the most important periodic orbits.
319: \end{abstract}
320:
321: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
322:
323: \section{Introduction}
324: \label{sec_in}
325:
326: Our contribution to this Special Issue lies at the intersection of Martin
327: Gutzwiller's scientific interests, namely celestial mechanics, electron
328: motion, and chaos, especially its manifestations in quantal systems. We
329: will be performing ``celestial mechanics on a microscopic scale''
330: \cite{UDMRS91} by treating the dynamics of highly excited (``Rydberg'')
331: electrons \cite{Connerade} in crossed electric and magnetic fields using
332: classical mechanics.
333:
334: Rydberg atoms in strong external fields constitute fundamental physical
335: systems where the quantum mechanical regime of strong nonlinearity can be
336: tested \cite{Gutzwiller,Kock/vanLeeuwen}. While the problem of a Rydberg
337: atom interacting with a strong magnetic field (the Diamagnetic Kepler
338: Problem, DKP, also known as the Quadratic Zeeman Effect, QZE) has been
339: fairly well understood as a result of sustained research in the past two
340: decades \cite{Hasegawa/Robnik/Wunner,Friedrich/Wintgen}, the superficially
341: similar scenario resulting from the addition of a perpendicular electric
342: field -- the so-called crossed field arrangement
343: \cite{Solovev,Braun/Solovev,Wiebusch,Raithel/Fauth/Walther,
344: vonMilczewski/Diercksen/Uzer,vonMilczewski/Diercksen/Uzer2} -- remains the
345: least understood of all Rydberg problems. This is all the more remarkable
346: in view of the prominence of the crossed fields in diverse areas of physics
347: ranging from excitonic systems to plasmas and neutron stars. This problem
348: is so complex because no continuous symmetry survives the extensive
349: symmetry breaking \cite{Delande/Gay} induced by the two fields. The result
350: is a wealth of new physics which is only possible beyond two degrees of
351: freedom, such as Arnol'd diffusion
352: \cite{TLL79,Gutzwiller,Lichtenberg/Lieberman,vonMilczewski/Diercksen/Uzer3}.
353: This absence of symmetry also allows localizing electronic wavepackets in
354: all spatial dimensions, and the observation of these localized wavepackets
355: \cite{Yeazell} has led to new insights into the dynamics of the electron in
356: the correspondence principle regime. It has also been found that a
357: velocity-dependent, Coriolis-like force in Newton's equations causes the
358: ionization of the electron to exhibit chaotic scattering
359: \cite{Main/Wunner2,Uzer/Farrelly2,JFU99}. All these phenomena, as well as
360: renewed interest in the motional Stark effect
361: \cite{Johnson/Hirschfelder/Yang,Farrelly}, make the crossed-fields problem
362: an experimental accessible paradigm for a wide variety of outstanding
363: issues in atomic and molecular physics, solid-state physics
364: \cite{Digman/Sipe,Schmelcher}, nuclear physics \cite{Bohr/Mottelson},
365: astrophysics \cite{Mathys}, and celestial mechanics \cite{Mignard}.
366:
367: The experimental challenge has been taken up by Raithel, Fauth, and
368: Walther \cite{Raithel/Fauth/Walther,Raithel/Fauth/Walther2} who in a
369: landmark series of experiments have identified a class of quasi-Landau (QL)
370: resonances in the spectra of rubidium Rydberg atoms in crossed electric and
371: magnetic fields. Similar to the original QL resonances observed by Garton
372: and Tomkins \cite{Garton/Tomkins}, this set of resonances is associated
373: with a rather small set of {\em planar} orbits of the crossed-fields
374: Hamiltonian which is known to support an enormous number of mostly
375: non-planar periodic motions \cite{Raithel/Fauth/Walther}. The dominance of
376: planar orbits in these experiments has recently been explained
377: \cite{vMFU97}.
378:
379: In contrast with the DKP, and despite some preliminary work \cite{FW96},
380: the systematics of periodic orbits in the crossed-fields problem has not
381: been discovered up to now. The aim of the present work is to initiate a
382: systematic classification of the orbits of the crossed-fields Hamiltonian,
383: based on methods developed in celestial mechanics, specifically Delaunay
384: variables and averaging. The analogy between atomic and planetary systems
385: was already used by Born \cite{Born}, who studied in particular the
386: crossed-fields problem, but neglected the quadratic Zeeman term because he
387: was not studying Rydberg atoms, where it is prominent.
388:
389: Delaunay variables are action--angle variables which have a clear geometric
390: interpretation in terms of Kepler ellipses. A fascinating historical
391: account of the developments in celestial mechanics leading to the
392: introduction of Delaunay variables can be found in a recent review by
393: Gutzwiller \cite{Gutzwiller2}. These variables allow to separate the time
394: scales of the motion, which is represented as a fast rotation of the planet
395: (or the electron), along a Kepler ellipse with slowly changing orientation
396: and eccentricity.
397:
398: The technique of averaging \cite{V96} allows to decrease the number of
399: degrees of freedom by eliminating the fast motion along the Kepler ellipse.
400: It has been used for a long time in celestial mechanics to compute the
401: so-called secular motion of the solar system, and can be considered as a
402: first order perturbation theory \cite{Deprit,Henrard}. A systematic use of
403: averaging allowed Laskar to integrate the motion of the solar system over
404: several hundred million years \cite{Laskar1,Laskar2,LR93}. The method has
405: been applied to the DKP in \cite{DKN83,CDMW87}.
406:
407: We note in passing that the direct connection to celestial mechanics that
408: the Delaunay variables provide is lost with an alternative set of variables
409: called the Lissajous elements. These are obtained by regularizing the
410: Coulomb Hamiltonian \cite{Farrelly/Uzer/&92}, and are appropriate for
411: investigating the level structure of Rydberg atoms. The connection between
412: the two sets in two dimensions is given by \cite{DW91}.
413:
414: In this paper, we use the following notations.
415: The Hamiltonian of an electron subjected to a Coulomb potential, a magnetic
416: field $\vec B$ and an electric field $\vec F$ can be written in
417: dimensionless units as
418: \begin{equation}
419: \label{i1}
420: H(p_x,p_y,p_z;x,y,z) = \frac12 p^2 - \frac1r + \frac12 \vec L\cdot\vec B +
421: \frac18(\vec r\wedge\vec B)^2 - \vec r\cdot\vec F,
422: \end{equation}
423: where $\vec r = (x,y,z)$ is the electron's position, $\vec p =
424: (p_x,p_y,p_z)$ its momentum, and $\vec L = \vec r \wedge \vec p$ its angular
425: momentum. We write $r=\norm{\vec r}$ and $p=\norm{\vec p}$.
426:
427: Our paper is organized as follows. In Section \ref{sec_ze}, we summarize
428: previous results on the case when only the magnetic field is present.
429: This allows us to introduce Delaunay variables and the method of averaging,
430: and illustrate them in a relatively simple situation.
431:
432: In Section \ref{sec_s}, we consider the case when only an electric field is
433: present, which is integrable \cite{R63}. We introduce another set of
434: action--angle variables (``electric action--angle variables'') based on
435: parabolic variables, which are better adapted to perturbation theory
436: \cite{Born}. We then derive (new) transformation formulas from electric
437: action--angle variables to the geometrically more transparent Delaunay
438: variables.
439:
440: With these tools and sets of coordinates in mind, we finally turn to the
441: crossed-fields Hamiltonian \eqref{i1} in Section \ref{sec_cr}. We start by
442: considering the two limiting cases $B\ll F$ and $F\ll B$, which involve
443: three distinct time scales, and can thus be analysed by a second averaging.
444: We then study the dynamics in the plane perpendicular to $\vec B$ for
445: general values of the fields. We conclude by an overview of the general
446: structure of the phase space of the averaged Hamiltonian.
447:
448: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
449:
450: \section{The Quadratic Zeeman Effect (or DKP)}
451: \label{sec_ze}
452:
453: We start by considering Hamiltonian \eqref{i1} in the case $F=0$. If we take
454: the $z$-axis along the magnetic field $\vec B$, \eqref{i1} can be written as
455: \begin{equation}
456: \label{z1}
457: H = \frac12 p^2 - \frac1r + \frac B2 L_z + \frac{B^2}8 (x^2+y^2).
458: \end{equation}
459: Although the equations of motion are easily written down, it is difficult to
460: understand the qualitative properties of dynamics in cartesian coordinates.
461: We will therefore take advantage of the fact that for small $B$, \eqref{z1}
462: is a small perturbation of the integrable Kepler problem, for which
463: action--angle variables are known explicitly. By writing the Hamiltonian
464: \eqref{z1} in these variables, we can separate slow and fast
465: components of the motion. The qualitative dynamics can then be further
466: analysed by using the method of averaging.
467:
468: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
469:
470: \subsection{Delaunay variables}
471: \label{sec_zdv}
472:
473: We start by considering the Kepler Hamiltonian
474: \begin{equation}
475: \label{zd1}
476: H = \frac12 p^2 - \frac1r.
477: \end{equation}
478: Besides the energy $H$, it admits as constants of motion the angular
479: momentum $\vec L$ and the Runge-Lenz vector
480: \begin{equation}
481: \label{zd2}
482: \vec A = \frac{\vec r}{r} + \vec L \wedge \vec p.
483: \end{equation}
484: If $H<0$, the motion takes place in a plane perpendicular to $\vec L$, on
485: an ellipse of eccentricity $e = \norm{\vec A}$ and major axis parallel to
486: $\vec A$ and of length $2a = -1/H$.
487:
488: Action--angle variables taking these properties into account are well known
489: in celestial mechanics, where they are called \defwd{Delaunay variables}.
490: The action variables are given by\footnote{We use the letters $G$ and $K$ in
491: order to distinguish the action variables from their physical meaning. In
492: celestial mechanics, one usually denotes $L_z$ by $H$ instead of $K$, but we
493: prefer the latter notation in order to avoid confusion with the Hamiltonian.}
494: \begin{equation}
495: \label{zd3}
496: \begin{split}
497: \Lambda &= \sqrt{a} \\
498: G &= \sqrt{a} \sqrt{1-e^2} = \norm{\vec L} \\
499: K &= \sqrt{a} \sqrt{1-e^2} \,\cos i = L_z,
500: \end{split}
501: \end{equation}
502: where the inclination $i\in[0,\pi]$ is the angle between $\vec L$ and the
503: $z$-axis. These variables are defined on the domain $\abs{K}\leqs G\leqs
504: \Lambda$.
505:
506: The corresponding angle variables are defined in the following way. The
507: intersection between the plane of the orbit and the $xy$-plane is called
508: \defwd{line of nodes}. The angle $\W$ between line of nodes and $x$-axis is
509: called \defwd{longitude of nodes} and is conjugated to $K$; the angle $\w$
510: between major axis of the Kepler ellipse and line of nodes is called
511: \defwd{argument of perihelion} and is conjugated to $G$; the \defwd{mean
512: anomaly} $M$, which is conjugated to $\Lambda$, is proportional to the area
513: swept on the ellipse, according to Kepler's second law.
514:
515: \begin{figure}
516: \centerline{\psfig{figure=fig01.eps,height=60mm,clip=t}}
517: \vspace{1mm}
518: \caption[]
519: {Definition of Delaunay variables. The angles $\W$, $i$ and $\w$ determine
520: the position of the Kepler ellipse in space. The line of nodes $ON$ is the
521: intersection of the plane of the ellipse with the $xy$-plane. The shaded
522: area is proportional to the mean anomaly $M$. The true anomaly $v$ and
523: eccentric anomaly $E$ are auxiliary quantities, allowing to relate the
524: $(X,Y)$-coordinates of the planet $P$ with $M$.}
525: \label{fig_z1}
526: \end{figure}
527:
528: In order to compute $M$, we introduce orthogonal coordinates $(X,Y)$, where
529: $X$ is attached to the major axis of the ellipse. The \defwd{true anomaly}
530: $v$ and the \defwd{eccentric anomaly} $E$ are auxiliary quantities defined
531: by the relations
532: \begin{equation}
533: \label{zd4}
534: \begin{split}
535: X &= r\cos v = a (\cos E-e) \\
536: Y &= r\sin v = a \sqrt{1-e^2} \sin E,
537: \end{split}
538: \end{equation}
539: see \figref{fig_z1}. By trigonometry one can show that $M$ and $E$ are related by
540: \defwd{Kepler's equation}
541: \begin{equation}
542: \label{zd5}
543: M = E - e \sin E,
544: \end{equation}
545: which implies in particular that
546: \begin{equation}
547: \label{zd6}
548: \dtot ME = 1 - e\cos E = \frac ra .
549: \end{equation}
550: The transition from Delaunay variables to cartesian coordinates is done in the
551: following way. Given the actions \eqref{zd3}, we can compute
552: \begin{equation}
553: \label{zd7}
554: a = \Lambda^2, \qquad
555: e^2 = 1 - \frac{G^2}{\Lambda^2}, \qquad
556: \cos i = \frac KG.
557: \end{equation}
558: We can then determine $X$, $Y$ and the conjugated momenta
559: \begin{equation}
560: \label{zd8}
561: \begin{split}
562: P_X &= -\frac1{\sqrt a} \frac{\sin E}{1-e\cos E} \\
563: P_Y &= \frac1{\sqrt a} \frac{\sqrt{1-e^2}\cos E}{1-e\cos E}.
564: \end{split}
565: \end{equation}
566: The cartesian coordinates are then given by the relations
567: \begin{equation}
568: \label{zd9}
569: \begin{pmatrix}
570: x & p_x \\ y & p_y \\ z & p_z
571: \end{pmatrix}
572: = \cR_z(\W) \cR_x(i) \cR_z(\w)
573: \begin{pmatrix}
574: X & P_X \\ Y & P_Y \\ 0 & 0
575: \end{pmatrix},
576: \end{equation}
577: where $\cR_x$ and $\cR_z$ describe rotations around the $x$- and $z$-axis,
578: given by
579: \begin{equation}
580: \label{zd10}
581: \cR_x(i) =
582: \begin{pmatrix}
583: 1 & 0 & 0 \\
584: 0 & \cos i & -\sin i \\
585: 0 & \sin i & \phantom{-}\cos i
586: \end{pmatrix},
587: \qquad
588: \cR_z(\w) =
589: \begin{pmatrix}
590: \cos \w & -\sin \w & 0\\
591: \sin \w & \phantom{-}\cos \w & 0\\
592: 0 & 0 & 1
593: \end{pmatrix}.
594: \end{equation}
595: It is then straightforward to show that the Hamiltonian \eqref{zd1} takes
596: the form
597: \begin{equation}
598: \label{zd11}
599: H = -\frac1{2\Lambda^2},
600: \end{equation}
601: and thus the equations of motion are given by
602: \begin{align}
603: \nonumber
604: \dot{\Lambda} &= 0 & \dot{M} &= \frac{1}{\Lambda^3} \\
605: \label{zd12}
606: \dot{G} &= 0 & \dot{\w} &= 0 \\
607: \nonumber
608: \dot{K} &= 0 & \dot{\W} &= 0.
609: \end{align}
610: Besides the actions $\Lambda$, $G$, $K$, the two angles $\w$ and $\W$ are
611: also constants, which reflects the high degeneracy of the hydrogen atom.
612: Equation \eqref{zd11} also gives a physical interpretation of $\Lambda$ as a
613: function of the energy. In quantum mechanics, $\Lambda$ corresponds to
614: the principal quantum number.
615:
616: Let us now return to the Zeeman effect. The Hamiltonian \eqref{z1} can be
617: expressed in Delaunay variables as
618: \begin{equation}
619: \label{zd13}
620: \begin{split}
621: H &= -\frac1{2\Lambda^2} + \frac B2 K + B^2 H_1(\Lambda,G,K;M,\w) \\
622: H_1 &= \tfrac1{16} r^2\bigbrak{1 + \cos^2 i + \sin^2 i \cos(2\w+2v)},
623: \end{split}
624: \end{equation}
625: where $r$, $\sin i$ and $v$ can be expressed in terms of Delaunay variables
626: using \eqref{zd4}, \eqref{zd6} and \eqref{zd7}. The equations of motion take
627: the form
628: \begin{align}
629: \nonumber
630: \dot{\Lambda} &= B^2 \poisson{\Lambda}{H_1} &
631: \dot{M} &= \frac1{\Lambda^3} + B^2 \poisson{M}{H_1} \\
632: \label{zd14}
633: \dot{G} &= B^2 \poisson{G}{H_1} &
634: \dot{\w} &= \phantom{\frac1{\Lambda^3} +{}} B^2 \poisson{\w}{H_1} \\
635: \nonumber
636: \dot{K} &=0 &
637: \dot{\W} &= \frac B2 + B^2 \poisson{\W}{H_1},
638: \end{align}
639: where the Poisson bracket is defined by
640: \begin{equation}
641: \label{zd15}
642: \poisson fg =
643: \dpar fM \dpar g\Lambda - \dpar f\Lambda \dpar GM
644: + \dpar f\w \dpar gG - \dpar fG \dpar g\w
645: + \dpar f\W \dpar gK - \dpar fK \dpar g\W.
646: \end{equation}
647: We discuss the computation of these Poisson brackets in Appendix
648: \ref{app_pb} (see in particular \tabref{t_z1}).
649:
650: To first order in $B$, \eqref{zd14} describes the Larmor precession of the
651: ellipse. Since the Hamiltonian does not depend on $\W$, $K=L_z$ is a
652: constant of the motion and \eqref{zd14} is in effect a
653: two-degrees-of-freedom system. For small $B$, $M$ is a fast variable, while
654: $\Lambda$, $G$ and $\w$ are slow ones. The motion can thus be imagined as a
655: fast motion of the electron along a slowly ``breathing'' and rotating
656: ellipse. The dynamics can be visualized by a Poincar\'e map, taking for
657: instance a section at constant $M$, and plotting the value of $G$ and $\w$
658: at each intersection \cite{DKN83}.
659:
660: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
661:
662: \subsection{Averaging}
663: \label{ssec_za}
664:
665: To analyse the motion of \eqref{zd14} for small $B$, one can use the fact
666: that $M$ is the only fast variable, so that the dynamics of the slow
667: variables will be essentially determined by the average effect of $M$ during
668: one period.
669:
670: With any given function $f$ of the Delaunay variables, let us associate its
671: average
672: \begin{equation}
673: \label{za1}
674: \avrg{f}_\Lambda(G,K;\w,\W) = \frac1{2\pi} \int_0^{2\pi}
675: f(\Lambda,G,K;M,\w,\W) \dx M.
676: \end{equation}
677: The averaged Hamiltonian $\avrg{H}_\Lambda$ generates the
678: canonical equations
679: \begin{align}
680: \nonumber
681: \dot{\Lambda} &= 0 &
682: \dot{M} &= \frac1{\Lambda^3} + B^2 \poisson{M}{\avrg{H_1}_\Lambda} \\
683: \label{za2}
684: \dot{G} &= B^2 \poisson{G}{\avrg{H_1}_\Lambda} &
685: \dot{\w} &= \phantom{\frac1{\Lambda^3} +{}} B^2
686: \poisson{\w}{\avrg{H_1}_\Lambda} \\
687: \nonumber
688: \dot{K} &=0 &
689: \dot{\W} &= \frac B2 + B^2 \poisson{\W}{\avrg{H_1}_\Lambda}.
690: \end{align}
691: Since $\avrg{H_1}_\Lambda$ does not depend on $M$ and $\W$,
692: $\avrg{H}_\Lambda$ is in effect a one-degree-of-freedom Hamiltonian,
693: depending on $K$ and $\Lambda$ as on parameters.
694:
695: A standard result from averaging theory (see Appendix \ref{app_av}) states
696: that the equations \eqref{za2} are a good approximation of the equations
697: \eqref{zd14}, in the sense that
698: \begin{itemiz}
699: \item orbits of \eqref{za2} and \eqref{zd14} with the same initial
700: condition differ by a term of order $B^2$ during a time of order $1/B^2$;
701: \item to each nondegenerate equilibrium of \eqref{za2}, there corresponds
702: a periodic orbit of \eqref{zd14}, at a distance of order $B^2$ of the
703: equilibrium, and which has the same stability if the equilibrium is
704: hyperbolic.
705: \end{itemiz}
706:
707: \begin{table}
708: \begin{center}
709: \begin{tabular}{|c|c||c|c|}
710: \hline
711: \vrule height 12pt depth 8pt width 0pt
712: $f$ & $\avrg{f}_\Lambda$ & $f$ & $\avrg{f}_\Lambda$ \\
713: \hline
714: \vrule height 14pt depth 8pt width 0pt
715: $X$ & $-\frac32 ae$ &
716: $X^2$ & $a^2(\frac12+2 e^2)$ \\
717: \vrule height 12pt depth 8pt width 0pt
718: $Y$ & $0$ &
719: $Y^2$ & $a^2 (\frac12 - \frac12 e^2)$ \\
720: \vrule height 12pt depth 8pt width 0pt
721: $r$ & $a(1+\frac12 e^2)$ &
722: $r^2$ & $a^2(1+\frac32 e^2)$ \\
723: \vrule height 12pt depth 9pt width 0pt
724: $z$ & $-\frac32 a e \sin i \sin \w$ &
725: $z^2$ & $a^2 \sin^2 i \bigbrak{\frac12 + \frac14 e^2 (3-5\cos 2\w)}$\\
726: \hline
727: \end{tabular}
728: \end{center}
729: \caption[]
730: {Some important quantities and their averages over the fast variable $M$.}
731: \label{t_z2}
732: \end{table}
733:
734: Averages over $M$ can be computed quite easily by the formula
735: \begin{equation}
736: \label{za3}
737: \avrg{f}_\Lambda = \frac1{2\pi} \int_0^{2\pi} f(M) \dtot ME \dx E
738: = \frac1{2\pi} \int_0^{2\pi} f(M(E)) (1-e\cos E)\dx E.
739: \end{equation}
740: Some useful averages are given in \tabref{t_z2}. The averaged Hamiltonian
741: can be written in the form \cite{CDMW87}
742: \begin{equation}
743: \label{za4}
744: \avrg{H_1}_\Lambda = \tfrac1{16} \Lambda^4 \Bigbrak{(1+\cos^2 i)(1+\tfrac32
745: e^2) + \tfrac52 e^2 \sin^2i \cos 2\w},
746: \end{equation}
747: from which we deduce the relevant equations of motion
748: \begin{equation}
749: \label{za5}
750: \begin{split}
751: \dot{G} &= \tfrac5{16} B^2 \Lambda^4 e^2 \sin i \sin 2\w \\
752: \dot{\w} &= \tfrac1{16} B^2 \frac{\Lambda^4}G \Bigbrak{3(e^2-1) - 5\cos^2 i +
753: 5(e^2-\sin^2i)\cos 2\w}.
754: \end{split}
755: \end{equation}
756: The system can now be studied by analysing the orbits of \eqref{za5} or,
757: equivalently, the level lines of the function \eqref{za4} in the
758: $(\w,G)$-plane \cite{DKN83}. Doing this, one finds that the vector field
759: has two elliptic equilibrium points located at
760: \begin{equation}
761: \label{za6}
762: \w = \frac{\pi}2, \frac{3\pi}2, \qquad
763: G^2 = \sqrt5 \abs{K}\Lambda,
764: \end{equation}
765: which exist if $\abs{K}<\Lambda/\sqrt5$.
766:
767: The vector field behaves in a singular way at the boundaries $G=\abs{K}$
768: and $G=\Lambda$ of the domain. These singularities have been explained by
769: Coffey and coworkers \cite{CDMW87}, who showed that the phase space has the
770: topology of a sphere. The line $G=\Lambda$, $\w\in[0,2\pi)$ has to be
771: contracted into the north pole of the sphere, corresponding to $e=0$ and
772: thus to circular motion; indeed, in this case the perihelion, and hence its
773: argument $\w$, is undefined. The line $G=\abs{K}$, $\w\in[0,2\pi)$ has to be
774: contracted into the south pole of the sphere, corresponding to $i=0$ and
775: thus to equatorial motion; in that case, the sum $\W+\w$ is sufficient to
776: specify the position of the ellipse.
777:
778: \begin{figure}
779: \centerline{\psfig{figure=fig02.eps,height=55mm,clip=t}}
780: \vspace{1mm}
781: \caption[]
782: {Phase portraits of the averaged Hamiltonian \eqref{za3} on the sphere
783: $\x_1^2 + \x_2^2 + \x_3^2 = \bigpar{\frac12 (\Lambda^2-K^2)}^2$, (a) for
784: $\abs{K} > \Lambda/\sqrt5$ and (b) for $\abs{K} < \Lambda/\sqrt5$. The
785: north pole corresponds to circular C-orbits, the south pole (not shown) to
786: equatorial B-orbits, which are elliptic in both cases. In the second case,
787: two additional periodic orbits (the Z-orbits) appear in the plane
788: $\x_1=0$.}
789: \label{fig_z2}
790: \end{figure}
791:
792: To account for the spherical topology of phase space, \cite{CDMW87} have
793: introduced the variables
794: \begin{equation}
795: \label{za7}
796: \begin{split}
797: \x_1 &= G\Lambda e \sin i \,\cos \w \\
798: \x_2 &= G\Lambda e \sin i \,\sin \w \\
799: \x_3 &= G^2 - \tfrac12(\Lambda^2+K^2),
800: \end{split}
801: \end{equation}
802: which belong to the sphere
803: \begin{equation}
804: \label{za8}
805: \x_1^2 + \x_2^2 + \x_3^2 = \Bigpar{\frac{\Lambda^2-K^2}2}^2.
806: \end{equation}
807: Using the fact that $\avrg{H_1}_\Lambda$ can be put into the form
808: \begin{equation}
809: \label{za9}
810: \avrg{H_1}_\Lambda = \frac{\Lambda^4}{16} \Bigbrak{1+\cos^2 i + 3 e^2 +
811: \frac{\x_1^2-4\x_2^2}{\Lambda^2 G^2}},
812: \end{equation}
813: and the Poisson brackets in Table \ref{t_z3} of Appendix \ref{app_pb},
814: they derive the equations of motion
815: \begin{equation}
816: \label{za10}
817: \begin{split}
818: \dot{\x_1} &= \frac{\Lambda^4}{8G} \,\x_2 \Bigbrak{\frac{5\x_1^2}{\Lambda^2
819: G^2} - 1 + e^2 + 5 \cos^2 i} \\
820: \dot{\x_2} &= \frac{\Lambda^4}{8G} \,\x_1 \Bigbrak{\frac{5\x_2^2}{\Lambda^2
821: G^2} - 4(1-e^2)} \\
822: \dot{\x_3} &= \frac{5\Lambda^2}{4G} \,\x_1\x_2.
823: \end{split}
824: \end{equation}
825: In these variables, the poles have become equilibrium points around which
826: the flow is nonsingular, so that their stability can be easily determined.
827: Depending on the relative value of the constants of motion $\Lambda$ and
828: $K$, there are two qualitatively different phase portraits
829: (\figref{fig_z2}):
830:
831: \begin{enum}
832: \item If $\abs{K} > \Lambda/\sqrt5$, both poles are elliptic, and there
833: are no other equilibrium points. All other orbits rotate around the sphere
834: with $\dot\w<0$ (\figref{fig_z2}a).
835:
836: \item If $\abs{K} < \Lambda/\sqrt5$, the south pole is still elliptic, but
837: the north pole has become hyperbolic, and the new elliptic equilibria
838: \eqref{za6} have appeared in a pitchfork bifurcation. Two homoclinic
839: orbits of the north pole separate orbits rotating around each of the three
840: elliptic equilibria (\figref{fig_z2}b).
841: \end{enum}
842:
843: The averaging theorem shows that to each of the four possible equilibrium
844: points of \eqref{za10}, there corresponds a periodic orbit of the exact
845: system \eqref{zd14} (see also Appendix \ref{app_pb}). For further
846: reference, let us call B-orbits the equatorial orbits (which lie in the
847: plane perpendicular to $\vec B$), C-orbits the circular ones, and Z-orbits
848: those corresponding to the nontrivial equilibrium \eqref{za6}. One further
849: expects that the periodic orbits of \eqref{za10} approximate either
850: quasiperiodic KAM-type orbits of \eqref{zd14}, or ``soft'' chaotic
851: components associated with resonances. More prominent chaotic components
852: are expected near the homoclinic orbits of the north pole.
853:
854: In other works \cite{Hasegawa/Robnik/Wunner}, orbits are sometimes
855: represented in cylindric coordinates $(\rho,\phi,z)$. They can be deduced
856: from Delaunay variables by the relations
857: \begin{equation}
858: \label{za11}
859: \begin{split}
860: z &= r \sin i \sin(\w+v) \\
861: \rho &= r \sqrt{1-\sin^2i\sin^2(\w+v)}.
862: \end{split}
863: \end{equation}
864: The different periodic orbits can thus be parametrized either by the true
865: anomaly $v$ or by the eccentric anomaly $E$ as
866: \begin{align}
867: \nonumber
868: &\text{C-orbits} &
869: z &= \Lambda\sqrt{\Lambda^2-K^2}\,\sin v &
870: \rho &= \Lambda\sqrt{\Lambda^2\cos^2 v + K^2\sin^2 v} \\
871: \label{za12}
872: &\text{B-orbits} &
873: z &= 0 &
874: \rho &= \Lambda (\Lambda-\sqrt{\Lambda^2-K^2}\cos E) \\
875: \nonumber
876: &\text{Z-orbits} &
877: z &= \Lambda^2 \sin i \,(\cos E-e) &
878: \rho &= \Lambda^2 \sqrt{(1-e\cos E)^2 - \sin^2i\,(\cos E-e)^2},
879: \end{align}
880: where $\sin^2i = 1-\abs{K}/\sqrt5 \Lambda$ and $e^2 =
881: 1-\sqrt5\abs{K}/\Lambda$ in the last case.
882: The C-, B- and Z-orbits are labelled, respectively, $C$, $I_1$ and
883: $I_\infty$ in \cite{Hasegawa/Robnik/Wunner}.
884:
885: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
886:
887: %\newpage
888: \section{The Stark effect}
889: \label{sec_s}
890:
891: We consider now the Hamiltonian \eqref{i1} in the case $B=0$. If we choose
892: the $z$-axis along the electric field $\vec F$, it can be written as
893: \begin{equation}
894: \label{s1}
895: H = \frac12 p^2 - \frac1r + Fz.
896: \end{equation}
897: Besides the energy and the $z$-component of the angular momentum, this
898: system has a third constant of the motion and is thus integrable. Indeed, as
899: shown by \cite{R63}, the generalization of the Runge-Lenz vector
900: \eqref{zd2},
901: \begin{equation}
902: \label{s2}
903: \vec C = \vec A - \frac12 \bigpar{\vec r\wedge \vec F} \wedge \vec r
904: \end{equation}
905: satisfies the equation of motion
906: \begin{equation}
907: \label{s3}
908: \dot{\vec C} = \frac32 \vec L \wedge \vec F,
909: \end{equation}
910: and thus $C_z = \vec C \cdot \vec F$ is constant.
911:
912: We will start, in Section \ref{ssec_sd}, by analysing the system in
913: Delaunay variables, in particular in its averaged form, in order to get a
914: feeling for the geometry of the orbits. In Section \ref{ssec_saa}, we
915: present another description of the system, based on parabolic variables,
916: which allows to construct action--angle variables taking the constant $C_z$
917: into account. Though they are better suited for perturbation theory, these
918: action--angle variables have a less obvious geometric interpretation than
919: Delaunay variables. This is why we establish the transformation formulas
920: between both sets of variables in Section \ref{ssec_sc}, in the limit $F\to
921: 0$.
922:
923: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
924:
925: \subsection{Delaunay variables}
926: \label{ssec_sd}
927:
928: \begin{figure}
929: \centerline{\psfig{figure=fig03.eps,height=55mm,clip=t}}
930: \vspace{1mm}
931: \caption[]
932: {Phase portraits of the averaged Hamiltonian \eqref{sd4} (a) in the
933: $(\w,G)$-plane and (b) on the sphere $\x_1^2 + \x_2^2 + \x_3^2 =
934: \bigpar{\frac12 (\Lambda^2-K^2)}^2$, seen from below the south pole.
935: The elliptic points at $\w =\frac\pi2$, $\frac{3\pi}2$ correspond to
936: periodic S-orbits of the Stark Hamiltonian.}
937: \label{fig_s1}
938: \end{figure}
939:
940: In Delaunay variables, the Hamiltonian \eqref{s1} takes the form
941: \begin{equation}
942: \label{sd1}
943: \begin{split}
944: H &= -\frac1{2\Lambda^2} + F H_2(\Lambda,G,K;M,\w), \\
945: H_2 &= r \sin i\,\sin(\w+v),
946: \end{split}
947: \end{equation}
948: and the equations of motion have the structure
949: \begin{align}
950: \nonumber
951: \dot{\Lambda} &= F \poisson{\Lambda}{H_2} &
952: \dot{M} &= \frac1{\Lambda^3} + F \poisson{M}{H_2} \\
953: \label{sd2}
954: \dot{G} &= F \poisson{G}{H_2} &
955: \dot{\w} &= \phantom{\frac1{\Lambda^3} +{}} F \poisson{\w}{H_2} \\
956: \nonumber
957: \dot{K} &=0 &
958: \dot{\W} &= \phantom{\frac1{\Lambda^3} +{}} F \poisson{\W}{H_2}.
959: \end{align}
960: The constants of motion are $H$, $K$ and
961: \begin{equation}
962: \label{sd3}
963: C_z = -e\sin i\,\sin\w - \frac12 Fr^2 \bigbrak{1-\sin^2i \sin^2(\w+v)}.
964: \end{equation}
965: In order to understand the geometry of the orbits for small $F$, we may
966: analyse the averaged Hamiltonian
967: \begin{equation}
968: \label{sd4}
969: \avrg{H}_\Lambda = \frac1{2\Lambda^2} - \frac32 F a e \sin i\,\sin \w.
970: \end{equation}
971: The relevant equations of motion of this one-degree-of-freedom system are
972: \begin{equation}
973: \label{sd5}
974: \begin{split}
975: \dot{G} &= \tfrac32 F \Lambda^2 e \sin i \, \cos \w, \\
976: \dot{\w} &= \tfrac32 F \frac{G^4-K^2\Lambda^2}{G^3 e \sin i} \sin\w.
977: \end{split}
978: \end{equation}
979: We observe the existence of a pair of elliptic stationary points at $\w =
980: \frac\pi2, \frac{3\pi}2$ and $G^2 = \abs{K}\Lambda$, which implies $e =
981: \sin i = \sqrt{1-\abs{K}/\Lambda}$ (\figref{fig_s1}a). When $F=0$, the
982: constant of motion $C_z$ reaches its extremal values
983: $\pm(1-K/\abs{\Lambda})$ on these points. We will call S-orbits the
984: associated periodic orbits of the Stark Hamiltonian. In order to analyse
985: the motion at the boundaries of phase space, we use again the variables
986: \eqref{za7}. Since the averaged Hamiltonian can be written as
987: \begin{equation}
988: \label{sd5b}
989: \avrg{H_2}_\Lambda = -\frac32 \frac{\Lambda}G\x_2,
990: \end{equation}
991: they evolve according to
992: \begin{equation}
993: \label{sd6}
994: \begin{split}
995: \dot{\x_1} &= \tfrac32 F \frac{\Lambda}{G^2} \bigbrak{-\x_2^2 -
996: (\Lambda^2+K^2+2\x_3)\x_3} \\
997: \dot{\x_2} &= \tfrac32 F \frac{\Lambda}{G^2} \x_1\x_2 \\
998: \dot{\x_3} &= 3 F \Lambda \x_1.
999: \end{split}
1000: \end{equation}
1001: This shows that there are no other equilibrium points than the two elliptic
1002: ones, since the flow is nonsingular at the poles (\figref{fig_s1}b).
1003:
1004: \begin{figure}
1005: \centerline{\psfig{figure=fig04.eps,height=55mm,clip=t}}
1006: \vspace{1mm}
1007: \caption[]
1008: {Same as \figref{fig_s1}, but for $K=0$.}
1009: \label{fig_s1b}
1010: \end{figure}
1011:
1012: Besides the periodic orbits corresponding to the elliptic equilibrium
1013: points of \eqref{sd5}, the system displays quasiperiodic orbits for which
1014: both $\w$ and $G$ oscillate. There is a particular orbit following the
1015: meridian $\x_2=0$ of the sphere, for which the Kepler ellipse oscillates
1016: between a circular and an equatorial one, while its major axis is always
1017: perpendicular to $\vec F$. Note that the case $K=0$ is special, since the
1018: elliptic points merge at the south pole. All orbits then go through the
1019: south pole, where the eccentricity vanishes, which means that the electron
1020: approaches the nucleus arbitrarily closely (\figref{fig_s1b}).
1021:
1022: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1023:
1024: \subsection{Parabolic and electric action--angle variables}
1025: \label{ssec_saa}
1026:
1027: The separability of the Stark Hamiltonian in parabolic variables was already
1028: known to Max Born \cite{Born} from the earlier works of P.S.\ Epstein
1029: \cite{Epstein} and K.\ Schwarzschild \cite{Schwarzschild}. Parabolic variables
1030: $(P_\x,P_\y,P_\ph;\x,\y,\ph)$ are defined by
1031: \begin{align}
1032: \nonumber
1033: x &= \x\y\cos\ph &
1034: p_x &= \frac{\y P_\x+\x P_\y}{\x^2+\y^2} \cos\ph - \frac{P_\ph}{\x\y}
1035: \sin\ph \\
1036: \label{saa1}
1037: y &= \x\y\sin\ph &
1038: p_y &= \frac{\y P_\x+\x P_\y}{\x^2+\y^2} \sin\ph + \frac{P_\ph}{\x\y}
1039: \cos\ph \\
1040: \nonumber
1041: z &= \frac12(\x^2-\y^2) &
1042: p_z &= \frac{\x P_\x-\y P_\y}{\x^2+\y^2},
1043: \end{align}
1044: where $P_\ph = L_z = K$ is the $z$-component of the angular momentum.
1045:
1046: In these variables, the Hamiltonian takes the form
1047: \begin{equation}
1048: \label{saa2}
1049: H = \frac1{\x^2+\y^2} \Bigbrak{\frac12 P_\x^2 + \frac{K^2}{2\x^2} + F\x^4 +
1050: \frac12 P_\y^2 + \frac{K^2}{2\y^2} - F\y^4 - 2}.
1051: \end{equation}
1052: There are four constants of motion $H, \alpha_1, \alpha_2, K$, related by
1053: \begin{equation}
1054: \label{saa3}
1055: \begin{split}
1056: \alpha_1 &= \frac12 P_\x^2 + \frac{K^2}{2\x^2} - \x^2 H + F \x^4 \\
1057: \alpha_2 &= \frac12 P_\y^2 + \frac{K^2}{2\y^2} - \y^2 H - F \y^4 \\
1058: \alpha_1 + \alpha_2 &= 2.
1059: \end{split}
1060: \end{equation}
1061: If one scales time by a factor $\x^2+\y^2$, \eqref{saa2} is seen to
1062: describe the motion of two decoupled ``oscillators''. In fact, the constant
1063: $\alpha_1$ is of the form $\frac12 P_\x^2 + V_1(\x)$, where
1064: $V_1(\x)\to\infty$ in both limits $\x\to 0$ and $\x\to\infty$ when $F>0$.
1065: Thus the motion of $\x$ is always bounded. By contrast, $\alpha_2 =\frac12
1066: P_\y^2 + V_2(\y)$, where $V_2(\y)\to-\infty$ for $\y\to\infty$. Thus the
1067: level sets of $\alpha_2$ may be unbounded for large values of $\y^2$ or
1068: $\alpha_2$ (of order $F^{-1}$), corresponding to ionization
1069: (\figref{fig_s2}). The following discussion is limited to the bounded
1070: motion of $\y$, which exists for small $F$. In that case, action--angle
1071: variables can be constructed.
1072:
1073: \begin{figure}
1074: \centerline{\psfig{figure=fig05.eps,height=45mm,clip=t}}
1075: \vspace{1mm}
1076: \caption[]
1077: {Level lines of the constants of motion $\alpha_1$ and $\alpha_2$ for
1078: $H=1$, $K=\frac12$ and $F=0.05$. In contrast to $\x$, $\y$ may have an
1079: unbounded motion. However, the saddle point is located quite far away from
1080: the origin, at $\y^2=\Order{F^{-1}}$.}
1081: \label{fig_s2}
1082: \end{figure}
1083:
1084: %\subsubsection{Construction of action variables}
1085:
1086: Action variables related to the constants $\alpha_1$ and $\alpha_2$ are
1087: defined by
1088: \begin{equation}
1089: \label{saa4}
1090: \begin{split}
1091: J_\x &= \frac1{2\pi} \oint \sqrt{2H + \frac{2\alpha_1}{\x^2} -
1092: \frac{K^2}{\x^4} - 2F\x^2}\;\x\dx\x \\
1093: J_\y &= \frac1{2\pi} \oint \sqrt{2H + \frac{2\alpha_2}{\y^2} -
1094: \frac{K^2}{\y^4} + 2F\y^2}\;\y\dx\y,
1095: \end{split}
1096: \end{equation}
1097: where the integrals are over bounded level sets of $\alpha_1$ and
1098: $\alpha_2$, respectively.
1099: %
1100: In the case $F=0$, they can be computed explicitly \cite{Born}:
1101: \begin{equation}
1102: \label{saa5}
1103: J_\x = \frac12 \Bigbrak{-K + \frac{\alpha_1}{\sqrt{-2H}}},
1104: \qquad
1105: J_\y = \frac12 \Bigbrak{-K + \frac{\alpha_2}{\sqrt{-2H}}}.
1106: \end{equation}
1107: From this we deduce that
1108: \begin{equation}
1109: \label{saa6}
1110: H = -\frac1{2(J_\x+J_\y+K)^2},
1111: \qquad
1112: \alpha_{1,2} = \frac{2 J_{\x,\y} + K}{J_\x+J_\y+K},
1113: \end{equation}
1114: which shows, by comparison with \eqref{zd11}, that $J_\x+J_\y+K = \Lambda$
1115: is nothing but the first Delaunay action. Moreover, one finds that the
1116: constant of motion $C_z$ reduces to
1117: \begin{equation}
1118: \label{saa7}
1119: A_z = 1-\alpha_2 = \alpha_1-1 = \frac{J_\x-J_\y}\Lambda,
1120: \end{equation}
1121: which suggests to introduce the action $\Je = J_\y - J_\x$. In quantum
1122: mechanics, $\Je$ corresponds to the electric quantum number. For further
1123: reference, let us call $(\Lambda,\Je,K)$ the \defwd{electric action
1124: variables}. As we have seen that $\abs{A_z}\leqs 1-\abs{K}/\Lambda$ for
1125: $F=0$, they vary on the square domain
1126: \begin{equation}
1127: \label{saa7b}
1128: -(\Lambda-\abs{K}) \leqs \Je \leqs \Lambda-\abs{K}.
1129: \end{equation}
1130:
1131: For $F>0$, we need to know the expression of the Hamiltonian in action
1132: variables. This can be done perturbatively \cite{DK83} with the result
1133: \begin{equation}
1134: \label{saa8}
1135: H(\Lambda,\Je,K) = -\frac1{2\Lambda^2} - 3F\Lambda \Je -
1136: \frac14 F^2 \Lambda^4 (17\Lambda^2 - 3 \Je^2 - 9 K^2) + \Order{F^3}.
1137: \end{equation}
1138: The associated canonical equations are
1139: \begin{align}
1140: \nonumber
1141: \dot{\Lambda} &= 0 &
1142: \dot{w}_\Lambda &= \frac1{\Lambda^3} - 3F\Je - \tfrac32 F^2 \Lambda^3
1143: (17\Lambda^2 - 2\Je^2 - 6K^2) + \Order{F^3} \\
1144: \label{saa9}
1145: \dot{\Je} &= 0 &
1146: \dot{\we} &= \phantom{\frac1{\Lambda^3}} {}- 3F\Lambda + \tfrac32 F^2
1147: \Lambda^4 \Je + \Order{F^3} \\
1148: \nonumber
1149: \dot{K} &= 0 &
1150: \dot{w}_K &= \phantom{\frac1{\Lambda^3} - 3F\Lambda +{}} \tfrac92 F^2 \Lambda^4
1151: K + \Order{F^3},
1152: \end{align}
1153: where $\wL$, $\we$ and $\wK$ (the \defwd{electric angle variables}) are
1154: conjugated to $\Lambda$, $\Je$ and $K$ respectively. In the case $F=0$,
1155: these equations are equivalent to the equations \eqref{zd12} in Delaunay
1156: variables. The electric field suppresses the degeneracies of the Kepler
1157: problem, and introduces different time scales for the various angles.
1158:
1159: %\subsubsection{Construction of angle variables}
1160:
1161: In order to compute the electric angle variables, we need to parametrize
1162: the level curves of $\alpha_1$ and $\alpha_2$. For $F=0$, using
1163: \eqref{saa6} the first equation of \eqref{saa3} can be written as
1164: \begin{equation}
1165: \label{saa10}
1166: P_\x^2 + \frac{(\x^2-\x_+^2)(\x^2-\x_-^2)}{\Lambda^2\x^2} = 0,
1167: \end{equation}
1168: where $\x_{\pm}^2 = a_1\pm b_1$ are extremal values of $\x^2$ with
1169: \begin{equation}
1170: \label{saa11}
1171: a_1 = \Lambda(2J_\x+K),
1172: \qquad
1173: b_1 = 2\Lambda\sqrt{J_\x(J_\x+K)}.
1174: \end{equation}
1175: Similarly, $\y^2$ can vary between $a_2-b_2$ and $a_2+b_2$. In terms of the
1176: electric actions $(\Lambda,\Je,K)$, these limits are given by
1177: \begin{align}
1178: \nonumber
1179: a_1 &= \Lambda(\Lambda-\Je) &
1180: b_1 &= \Lambda\sqrt{(\Lambda-\Je)^2-K^2} \\
1181: \label{saa12}
1182: a_2 &= \Lambda(\Lambda+\Je) &
1183: b_2 &= \Lambda\sqrt{(\Lambda+\Je)^2-K^2}.
1184: \end{align}
1185: Relation \eqref{saa10} is the equation of an ellipse in the
1186: $(\x^2,\Lambda\x P_\x)$-plane, which suggests to parametrize the level sets
1187: by
1188: \begin{align}
1189: \nonumber
1190: \x &= \sqrt{a_1 - b_1\cos\psi} &
1191: \y &= \sqrt{a_2 - b_2\cos\chi} \\
1192: \label{saa13}
1193: P_\x &= \frac{b_1\sin\psi}{\Lambda\x} &
1194: P_\y &= \frac{b_2\sin\chi}{\Lambda\y},
1195: \end{align}
1196: where $\psi$ and $\chi$ are auxiliary angles, playing a similar role as the
1197: eccentric anomaly $E$ in the case of Delaunay variables. If we introduce the
1198: action
1199: \begin{equation}
1200: \label{saa14}
1201: S = \int^\x P_\x \dx \x' + \int^\y P_\y \dx \y' + K\ph,
1202: \end{equation}
1203: the angles conjugated to $(J_\x,J_\y,K)$ are given by the formulas
1204: \begin{equation}
1205: \label{saa15}
1206: w_\x = \dpar{S}{J_\x}, \qquad
1207: w_\y = \dpar{S}{J_\y}, \qquad
1208: w_\ph = \dpar{S}{K},
1209: \end{equation}
1210: and the angles conjugated to $(\Lambda,\Je,K)$ are then obtained by the
1211: linear transformation
1212: \begin{equation}
1213: \label{saa16}
1214: \wL = \tfrac12(w_\x+w_\y), \qquad
1215: \we = \tfrac12(w_\y-w_\x), \qquad
1216: \wK = w_\ph - \wL.
1217: \end{equation}
1218: With the parametrization \eqref{saa13}, the derivatives \eqref{saa15} take a
1219: simple form \cite{Born}, and the final result is
1220: \begin{equation}
1221: \label{saa17}
1222: \begin{split}
1223: \wL &= \frac{\psi+\chi}2 - \frac1{2\Lambda^2} (b_1\sin\psi + b_2\sin\chi) \\
1224: \we &= \frac{\chi-\psi}2 \\
1225: \wK &= \ph - \rho_1(\psi) - \rho_2(\chi),
1226: \end{split}
1227: \end{equation}
1228: where
1229: \begin{equation}
1230: \label{saa18}
1231: \rho_j(\th) = \frac{K\Lambda}2 \int_0^\th \frac{\dx
1232: \th'}{a_j-b_j\cos \th'}, \qquad j=1,2.
1233: \end{equation}
1234: Using the fact that $a_1^2-b_1^2 = K^2\Lambda^2$, this expression can be
1235: written as
1236: \begin{equation}
1237: \label{saa19}
1238: \rho_j(\th) = \sign(K) \Arctg
1239: \biggbrak{\sqrt{\frac{a_j+b_j}{a_j-b_j}}\tg\frac \th2}
1240: \qquad \text{for $-\pi\leqs \th\leqs\pi$,}
1241: \end{equation}
1242: and can be continued to arbitrary $\th$ by the rule $\rho_j(\th+k2\pi) =
1243: \rho_j(\th) + k\pi$ for $k\in\Z$. This implies that for {\em all} $\th$, we
1244: have
1245: \begin{equation}
1246: \label{saa20}
1247: \cos\rho_j(\th) = \frac{\sqrt{a_j-b_j}}{\sqrt{a_j-b_j\cos \th}} \cos\frac
1248: \th2,
1249: \qquad
1250: \sin\rho_j(\th) = \sign K \frac{\sqrt{a_j+b_j}}{\sqrt{a_j-b_j\cos \th}}
1251: \sin\frac \th2.
1252: \end{equation}
1253: Note in particular that the transformation $\we\mapsto\we+\pi$,
1254: $\wL\mapsto\wL+\pi$ corresponds to keeping $\psi$ fixed and increasing
1255: $\chi$ by $2\pi$. Hence it leaves all parabolic variables fixed, except
1256: $\ph$ which is increased by $\pi$. In other words, this transformation
1257: describes a rotation of angle $\pi$ around the $z$-axis.
1258:
1259: The relations \eqref{saa17} are valid for $F=0$.
1260: Higher order expressions in $F$ of the angle variables can be computed
1261: perturbatively, see Appendix \ref{app_aF}.
1262:
1263: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1264:
1265: \subsection{Correspondence between electric action--angle and\\
1266: Delaunay variables}
1267: \label{ssec_sc}
1268:
1269: We now establish transformation formulas between electric action--angle
1270: variables and Delaunay variables in the case $F=0$. In the two-dimensional
1271: case, corresponding to $K=0$, this issue has been addressed in \cite{DW91},
1272: in connection with Lissajous variables.
1273:
1274: The most important relations can be obtained by averaging over the fast
1275: variable $\wL$. In view of \eqref{saa17}, the averaging operation can be
1276: written as
1277: \begin{equation}
1278: \label{sc1}
1279: \avrg{f(\psi,\chi)}_\Lambda =
1280: \frac1{2\pi}\int_0^{2\pi} f(\psi,\psi+2\we)
1281: \Bigbrak{1 - \frac{b_1\cos\psi + b_2\cos(\psi+2\we)}{2\Lambda^2}} \dx\psi.
1282: \end{equation}
1283: A few useful averages are given in \tabref{t_s1}.
1284: %
1285: We can thus easily compute the averages
1286: \begin{equation}
1287: \label{sc2}
1288: \begin{split}
1289: \avrg{z}_\Lambda &= \frac12 \avrg{\x^2-\y^2}_\Lambda
1290: = -\tfrac32 \Lambda \Je \\
1291: \avrg{r}_\Lambda &= \frac12 \avrg{\x^2+\y^2}_\Lambda
1292: = \Lambda^2 \Bigbrak{1 + \frac1{8\Lambda^4} (b_1^2+b_2^2+2b_1b_2\cos2\we)}.
1293: \end{split}
1294: \end{equation}
1295: Comparison with the corresponding averages over $M$ (\tabref{t_z2}) gives
1296: us the relations
1297: \begin{align}
1298: \label{sc3}
1299: \Je &= \Lambda e \sin i \sin\w,\\
1300: \label{sc4}
1301: e^2 &= \frac1{4\Lambda^4} \bigpar{b_1^2 + b_2^2 + 2b_1b_2\cos 2\we}.
1302: \end{align}
1303: This last relation allows to compute $G$ and $i$. It has a nice geometric
1304: interpretation: the maximal value of $e^2$ (hence the minimal value of $G$)
1305: is attained for $\we=0$ and $\pi$, while the minimal value of $e^2$ is
1306: reached for $\we=\frac\pi2$ and $\frac{3\pi}2$. According to
1307: \figref{fig_s1}a, these values of $\we$ correspond to
1308: $\w=\frac\pi2\sign\Je$. From the fact that $\dot\we<0$ for $F>0$, we
1309: infer that
1310: \begin{equation}
1311: \label{sc5}
1312: \sign(\cos\w) = -\sign(\sin2\we).
1313: \end{equation}
1314: The relations \eqref{sc3}, \eqref{sc4} and \eqref{sc5} determine the
1315: transformation $(G,\w)\mapsto(\Je,\we)$ up to a phase $\pi$ of $\we$, which
1316: depends on $\W$.
1317:
1318: \begin{table}
1319: \begin{center}
1320: \begin{tabular}{|c|c||c|c|}
1321: \hline
1322: \vrule height 12pt depth 8pt width 0pt
1323: $f$ & $\avrg{f}_\Lambda$ & $f$ & $\avrg{f}_\Lambda$ \\
1324: \hline
1325: \vrule height 14pt depth 8pt width 0pt
1326: $\cos\psi$ & $-\frac{b_1}{4\Lambda^2} - \frac{b_2}{4\Lambda^2}\cos 2\we$ &
1327: $\cos\chi$ & $-\frac{b_1}{4\Lambda^2}\cos 2\we - \frac{b_2}{4\Lambda^2}$ \\
1328: \vrule height 12pt depth 8pt width 0pt
1329: $\sin\psi$ & $\frac{b_2}{4\Lambda^2}\sin 2\we$ &
1330: $\sin\chi$ & $-\frac{b_1}{4\Lambda^2}\sin 2\we$ \\
1331: \vrule height 12pt depth 8pt width 0pt
1332: $\cos^2\psi$ & $\frac12$ &
1333: $\cos^2\chi$ & $\frac12$ \\
1334: \vrule height 12pt depth 8pt width 0pt
1335: $\cos\psi\cos\chi$ & $\frac12\cos 2\we$ &
1336: $\cos\psi\sin\chi$ & $-\frac12\sin 2\we$ \\
1337: \vrule height 12pt depth 9pt width 0pt
1338: $\sin\psi\sin\chi$ & $\frac12\cos 2\we$ &
1339: $\sin\psi\cos\chi$ & $\frac12\sin 2\we$ \\
1340: \hline
1341: \end{tabular}
1342: \end{center}
1343: \caption[]
1344: {The averages over the fast variable $\wL$ of some important quantities.
1345: Using \eqref{saa13} they can be used to compute averages of some polynomials
1346: in $\x^2$ and $\y^2$.}
1347: \label{t_s1}
1348: \end{table}
1349:
1350: It remains to establish relations between the angles $(\wL,\wK)$ and
1351: $(M,\W)$. Since $\dot{\wL} = \dot{M}$ for $F=0$, the difference $\wL-M$ does
1352: not depend on $M$. From \eqref{saa13} we deduce that in electric
1353: action--angle variables,
1354: \begin{equation}
1355: \label{sc6}
1356: r = \Lambda^2\Bigbrak{1-\frac1{2\Lambda^2} (b_1\cos\psi+b_2\cos\chi)} =
1357: \Lambda^2 \Bigpar{\dpar{\wL}{\psi}+\dpar{\wL}{\chi}}.
1358: \end{equation}
1359: Comparison with \eqref{zd6} yields the equalities
1360: \begin{equation}
1361: \label{sc7}
1362: e\cos E = \frac1{2\Lambda^2} (b_1\cos\psi+b_2\cos\chi),
1363: \end{equation}
1364: and, using $\dot{\wL} = \dot{M}$,
1365: \begin{equation}
1366: \label{sc8}
1367: \dtot{}{E} = \dtot ME \dtot{}{M} =
1368: \Bigpar{\dpar{\wL}{\psi}+\dpar{\wL}{\chi}} \dtot{}{\wL} =
1369: \dpar{}{\psi}+\dpar{}{\chi}.
1370: \end{equation}
1371: Applied to \eqref{sc7}, this also gives
1372: \begin{equation}
1373: \label{sc9}
1374: e\sin E = \frac1{2\Lambda^2} (b_1\sin\psi+b_2\sin\chi).
1375: \end{equation}
1376: In particular, when $M=0$ we have $E=0$ and thus by \eqref{saa17}
1377: $\wL=\frac{\psi+\chi}2$. Inserting the relations $\psi=\wL-\we$ and
1378: $\chi=\wL+\we$ into \eqref{sc7} and \eqref{sc9}, we can solve for $\cos\wL$
1379: and $\sin\wL$ with the result, for general values of $M$,
1380: \begin{equation}
1381: \label{sc10}
1382: \cos(\wL-M) = \frac{2\Lambda^2e}{b_1+b_2} \cos\we,
1383: \qquad
1384: \sin(\wL-M) = \frac{2\Lambda^2e}{b_2-b_1} \sin\we.
1385: \end{equation}
1386: Relation \eqref{sc8} also implies that
1387: \begin{equation}
1388: \label{sc11}
1389: \psi-E = \wL-\we-M, \qquad \chi-E = \wL+\we-M.
1390: \end{equation}
1391:
1392: Determining $\wK$ is a bit more delicate. We will use the fact that in
1393: Delaunay variables, the $x$-component of the angular momentum is
1394: \begin{equation}
1395: \label{sc12}
1396: L_x = \sqrt{G^2-K^2} \sin\W,
1397: \end{equation}
1398: while in parabolic variables, we have from \eqref{saa1}
1399: \begin{equation}
1400: \label{sc13}
1401: L_x = \frac12 (\y P_\x - \x P_\y) \sin\ph
1402: - \frac12 \frac{K}{\x\y} (\x^2-\y^2) \cos\ph.
1403: \end{equation}
1404: Being independent of $M$, $L_x$ also has to be independent of $\wL$. We may
1405: thus evaluate \eqref{sc13} in the case $\psi=0$, $\chi=2\we$. Then
1406: \eqref{saa13} reduces to
1407: \begin{equation}
1408: \label{sc14}
1409: \x=\sqrt{a_1-b_1}, \qquad
1410: \y=\sqrt{a_2-b_2\cos2\we}, \qquad
1411: P_\x=0, \qquad
1412: P_\y=\frac{b_2\sin2\we}{\Lambda\y},
1413: \end{equation}
1414: and \eqref{saa17} implies $\ph=\wK+\rho_2$, where
1415: \begin{equation}
1416: \label{sc15}
1417: \cos\rho_2 = \frac{\sqrt{a_2-b_2}}\y \cos\we, \qquad
1418: \sin\rho_2 = \sign(K)\frac{\sqrt{a_2+b_2}}\y \sin\we.
1419: \end{equation}
1420: Inserting \eqref{sc14} and \eqref{sc15} into \eqref{sc13}, we obtain
1421: \begin{equation}
1422: \label{sc16}
1423: L_x = L_1 \cos\wK + L_2 \sin\wK,
1424: \end{equation}
1425: where
1426: \begin{align}
1427: \nonumber
1428: \frac{L_1}{\cos\we} &=
1429: -\frac1{2\Lambda\x\y^2} \Bigbrak{2\sign(K)\x^2\sqrt{a_2+b_2}\,b_2\sin^2\we +
1430: K\Lambda(\x^2-\y^2)\sqrt{a_2-b_2}} \\
1431: \nonumber
1432: &= \frac12 K \sqrt{\frac{a_2-b_2}{a_1-b_1}}
1433: - \frac{\x}{2\Lambda\y^2\sqrt{a_2-b_2}} \Bigbrak{K\Lambda(a_2-b_2) +
1434: 2\sign(K)b_2\sqrt{a_2^2-b_2^2}\sin^2\we} \\
1435: \label{sc17}
1436: &= \frac12 K \biggbrak{\sqrt{\frac{a_2-b_2}{a_1-b_1}} -
1437: \sqrt{\frac{a_1-b_1}{a_2-b_2}}\,},
1438: \end{align}
1439: where we have used the relation $\sqrt{a_2^2-b_2^2} = \abs{K}\Lambda$.
1440: The term $L_2$ can be evaluated in a similar way. We obtain the following,
1441: relatively compact expression of $L_x$ in electric action--angle variables:
1442: \begin{equation}
1443: \label{sc18}
1444: L_x = Y_1(\Lambda,\Je,K) \cos\we\cos\wK
1445: + Y_2(\Lambda,\Je,K) \sin\we\sin\wK,
1446: \end{equation}
1447: where we have introduced the notations
1448: \begin{equation}
1449: \label{sc19}
1450: \begin{split}
1451: Y_1(\Lambda,\Je,K) &=
1452: \frac12 K \biggbrak{\sqrt{\frac{a_2-b_2}{a_1-b_1}} -
1453: \sqrt{\frac{a_1-b_1}{a_2-b_2}}\,} \\
1454: Y_2(\Lambda,\Je,K) &=
1455: -\frac12 \abs{K} \biggbrak{\sqrt{\frac{a_2+b_2}{a_1-b_1}} -
1456: \sqrt{\frac{a_1-b_1}{a_2+b_2}}\,}.
1457: \end{split}
1458: \end{equation}
1459: Comparison with \eqref{sc12} gives the desired relation between $\wK$ and
1460: $\W$:
1461: \begin{equation}
1462: \label{sc20}
1463: \begin{split}
1464: \sqrt{G^2-K^2} \sin(\W-\wK) &= Y_1(\Lambda,\Je,K) \cos\we \\
1465: \sqrt{G^2-K^2} \cos(\W-\wK) &= Y_2(\Lambda,\Je,K) \sin\we.
1466: \end{split}
1467: \end{equation}
1468: We point out that despite the absolute value, the expressions \eqref{sc19}
1469: are regular at $K=0$ and admit the Taylor series
1470: \begin{equation}
1471: \label{sc21}
1472: \begin{split}
1473: Y_1(\Lambda,\Je,K) &= -K\frac{\Je}{\sqrt{\Lambda^2-\Je^2}}
1474: +K^3\frac{\Lambda^2\Je}{2(\Lambda^2-\Je^2)^{5/2}} + \Order{K^5} \\
1475: Y_2(\Lambda,\Je,K) &= -\sqrt{\Lambda^2-\Je^2}
1476: +K^2\frac{\Lambda^2}{2(\Lambda^2-\Je^2)^{3/2}} + \Order{K^4}.
1477: \end{split}
1478: \end{equation}
1479:
1480: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1481:
1482: \section{The crossed-fields problem}
1483: \label{sec_cr}
1484:
1485: We now consider the full crossed-fields Hamiltonian \eqref{i1}, in the case
1486: $0<B,F\ll 1$. First, we have to choose a system of coordinates. Both sets of
1487: action--angle variables that we have used so far are defined with
1488: respect to a privileged direction (the $z$-axis). Depending on the regime we
1489: consider, it will be most convenient to choose this direction along the
1490: electric or along the magnetic field. In the first case, the Hamiltonian
1491: takes the form
1492: \begin{equation}
1493: \label{cr1}
1494: H = \frac12 p^2 - \frac1r + Fz + \frac12 B L_x + \frac18 B^2(y^2+z^2).
1495: \end{equation}
1496: To account for the second case, we also introduce coordinates
1497: $(x',y',z')=(z,y,x)$. In Section \ref{ssec_cB}, we consider the case $B\ll
1498: F$, which is a small perturbation of the integrable Stark effect, and thus
1499: particularly well suited to perturbation theory. The case $F\ll B$ is
1500: considered in Section \ref{ssec_cF}. The orbits contained in the plane
1501: perpendicular to $\vec B$ exist for all values of the fields. We analyse
1502: them in Section \ref{ssec_cq}. Other periodic orbits and the general
1503: structure of phase space are discussed in Section \ref{ssec_cp}.
1504:
1505: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1506:
1507: \subsection{The case $B\ll F$}
1508: \label{ssec_cB}
1509:
1510: When the magnetic field acts as a small perturbation of the Stark
1511: Hamiltonian, it is best to use the electric action--angle variables
1512: introduced in Section \ref{ssec_saa}. The Hamiltonian can be written as
1513: \begin{equation}
1514: \label{cB1}
1515: \begin{split}
1516: H = H_0(\Lambda,\Je,K;F) &+ B H_1(\Lambda,\Je,K;\we,\wK;F) \\
1517: &+ B^2 H_2(\Lambda,\Je,K;\wL,\we,\wK;F),
1518: \end{split}
1519: \end{equation}
1520: where $H_0 = -\frac1{2\Lambda^2} - 3F\Lambda\Je + F^2 h_2(\Lambda,\Je,K;F)$
1521: is the Stark Hamiltonian \eqref{saa8}, $H_1 = \frac12 L_x$ has been
1522: computed in \eqref{sc18}, and $H_2 = \frac18(y^2+z^2)$.
1523: The equations of motion thus have the structure
1524: \begin{align}
1525: \nonumber
1526: \dot{\Lambda} &= \Order{B^2}
1527: & \dot{\wL} &= \frac1{\Lambda^3} - 3F\Je + \Order{F^2} + \Order{B} \\
1528: \label{cB2}
1529: \dot{\Je} &= \Order{B}
1530: & \dot{\we} &= \phantom{\frac1{\Lambda^3}} {}- 3F\Lambda +
1531: \Order{F^2} + \Order{B} \\
1532: \nonumber
1533: \dot{K} &= \Order{B}
1534: & \dot{\wK} &= \phantom{\frac1{\Lambda^3} - 3F\Lambda +{}}
1535: \Order{F^2} + \Order{B}
1536: \end{align}
1537: We can again average over the fast variable $\wL$, using the rule
1538: \eqref{sc1}. Since $L_x$ does not depend on $\wL$, $H_1$ is already in
1539: averaged form. The equations of the averaged system are thus given by
1540: \begin{equation}
1541: \label{cB3}
1542: \begin{split}
1543: \dot{\Je} &= \phantom{-3 F\Lambda + F^2 \poisson{\we}{h_2} +{}}
1544: B\poisson{\Je}{H_1} + B^2\poisson{\Je}{\avrg{H_2}_\Lambda} \\
1545: \dot{K} &= \phantom{-3 F\Lambda + F^2 \poisson{\we}{h_2} +{}}
1546: B\poisson{K}{H_1} + B^2\poisson{K}{\avrg{H_2}_\Lambda} \\
1547: &\\
1548: \dot{\we} &= -3 F\Lambda + F^2 \poisson{\we}{h_2} +
1549: B\poisson{\we}{H_1} + B^2\poisson{\we}{\avrg{H_2}_\Lambda} \\
1550: \dot{\wK} &= \phantom{-3 F\Lambda +{}} F^2 \poisson{\wK}{h_2} +
1551: B\poisson{\wK}{H_1} + B^2\poisson{\wK}{\avrg{H_2}_\Lambda} \\
1552: \end{split}
1553: \end{equation}
1554: Since we assume that $B\ll F$, $\we$ evolves on a faster time scale than
1555: $\wK$. We may thus further approximate the dynamics by averaging over $\we$,
1556: that is, we define the double average
1557: \begin{equation}
1558: \label{cB4}
1559: \avvrg{f}_{\Lambda,\Je}(K;\wK) =
1560: \frac1{4\pi^2} \int_0^{2\pi}\int_0^{2\pi} f(\Lambda,\Je,K;\wL,\we,\wK)
1561: \dx \wL \dx \we.
1562: \end{equation}
1563: The doubly-averaged Hamiltonian $\avvrg{H}_{\Lambda,\Je}$ has one degree of
1564: freedom and is thus integrable.
1565: Let us now compute various averages, at lowest order in $F$. In order
1566: to compute $\avrg{H_2}_\Lambda$, we need to evaluate the average of
1567: \begin{equation}
1568: \label{cB5}
1569: y^2+z^2 = \x^2\y^2\sin^2\ph + \tfrac14 (\x^4+\y^4-2\x^2\y^2).
1570: \end{equation}
1571: Using the expressions in \tabref{t_s1}, the second term is easily averaged.
1572: To average the first term, we use the fact that with \eqref{saa17} and
1573: \eqref{saa20}, $\x\y\sin\ph$ can be written as a polynomial in
1574: $\sqrt{a_1\pm b_1}$, $\sqrt{a_2\pm b_2}$, and sines and cosines of $\wK$,
1575: $\frac\psi2$ and $\frac\chi2$. The final result after simplification is
1576: \begin{equation}
1577: \label{cB6}
1578: \begin{split}
1579: \avrg{y^2+z^2}_\Lambda ={} &
1580: \tfrac12 \Lambda^2 (2\Lambda^2 + 3\Je^2 - K^2) - \tfrac34 b_1b_2\cos 2\wK \\
1581: &+ \tfrac14 b_1b_2 \cos 2\we -
1582: \tfrac12 \Lambda^2(\Lambda^2-\Je^2-K^2)\cos 2\we\,\cos 2\wK \\
1583: &- \Lambda^2 \Je K \sin 2\we\,\sin 2\wK + \Order{F}.
1584: \end{split}
1585: \end{equation}
1586:
1587: Computation of the doubly-averaged Hamiltonian is now easy. From
1588: \eqref{sc18} and \eqref{cB6} we obtain
1589: \begin{equation}
1590: \label{cB7}
1591: \begin{split}
1592: \avvrg{L_x}_{\Lambda,\Je} &= \Order{F} \\
1593: \avvrg{y^2+z^2}_{\Lambda.\Je} &=
1594: \tfrac12 \Lambda^2 (2\Lambda^2 + 3\Je^2 - K^2) - \tfrac34 b_1b_2\cos 2\wK +
1595: \Order{F}.
1596: \end{split}
1597: \end{equation}
1598: In fact, equation \eqref{aF5} in Appendix \ref{app_aF} shows that also at
1599: order $F$ (and probably at all higher orders), the transformation
1600: $\we\mapsto\we+\pi$, $\wL\mapsto\wL+\pi$ describes a rotation of angle
1601: $\pi$ around the $z$-axis. Since this rotation changes $L_x$ into $-L_x$,
1602: we conclude that $\avvrg{L_x}_{\Lambda,\Je}=\Order{F^2}$.
1603: Discarding irrelevant constant terms, the doubly-averaged Hamiltonian can
1604: be written as
1605: \begin{equation}
1606: \label{cB8}
1607: \avvrg{H}_{\Lambda,\Je} = \tfrac94F^2\Lambda^4K^2
1608: -\tfrac1{16}B^2\bigbrak{\Lambda^2 K^2 + \tfrac32 b_1b_2\cos 2\wK}
1609: +\Order{F^3,BF^2,B^2F},
1610: \end{equation}
1611: where $b_1$ and $b_2$ are given in \eqref{saa12} and the remainder denotes
1612: a sum of terms of order $F^3$, $BF^2$ and $B^2F$. Besides $\Lambda$ and
1613: $\Je$, \eqref{cB8} is a third adiabatic invariant of the crossed-fields
1614: Hamiltonian in the case $B\ll F\ll 1$. Up to the remainders, the equations
1615: of motion have the form
1616: \begin{equation}
1617: \label{cB9}
1618: \begin{split}
1619: \dot{K} &= -\frac3{16} B^2 b_1b_2\sin 2\wK \\
1620: \dot{\wK} &= \frac92 F^2 \Lambda^4 K
1621: -\frac18 B^2\Lambda^2 K \Bigbrak{1-\frac32
1622: \frac{\Lambda^2}{b_1b_2}(\Lambda^2+\Je^2-K^2)\cos 2\wK}.
1623: \end{split}
1624: \end{equation}
1625:
1626: We should also examine the topology of phase space. From \eqref{saa7b}, we
1627: deduce that in the limit $F\to 0$, $K$ varies between
1628: $-(\Lambda-\abs{\Je})$ and $\Lambda-\abs{\Je}$. If, say, $\Je\geqs 0$ and
1629: $K=\Lambda-\Je$, we have $b_1=0$ and the angle $\psi$ is undefined. Since
1630: $\rho_1(\psi)=\frac\psi2$ in this case, we obtain from \eqref{saa13} and
1631: \eqref{saa17} that the quantities $\wL+\wK$ and $\wL+\we$ are sufficient to
1632: determine the state of the system completely. We conclude that in the
1633: averaged phase space, the variables $\wK$ and $\we$ are irrelevant when
1634: $\abs{K} = \Lambda-\abs{\Je}$, and thus we have again a spherical topology.
1635: The sphere can be parametrized by
1636: \begin{equation}
1637: \label{cB10}
1638: (\kappa_1,\kappa_2,\kappa_3)
1639: = (\sqrt{(\Lambda-\abs{\Je})^2-K^2}\cos\wK,
1640: \sqrt{(\Lambda-\abs{\Je})^2-K^2}\sin\wK, K).
1641: \end{equation}
1642:
1643: \begin{figure}
1644: \centerline{\psfig{figure=fig06.eps,height=45mm,clip=t}}
1645: \vspace{1mm}
1646: \caption[]
1647: {Phase portraits of the doubly-averaged system \eqref{cB8} on the sphere
1648: $\kappa_1^2+\kappa_2^2+\kappa_3^2 = (\Lambda-\abs{\Je})^2$, (a) for
1649: $0<B<B_2$, where $B_2$ is given by \eqref{cB13}, (b) for $B_2<B<B_1$, given
1650: by \eqref{cB11}, and (c) for $B>B_1$. The poles (the intersection of the
1651: sphere with the $\kappa_3$-axis) correspond to the S-orbits present in the
1652: Stark effect. The intersections of the sphere with the $\kappa_1$- and
1653: $\kappa_2$-axis correspond respectively to the BF-orbits, lying in the
1654: plane defined by $\vec B$ and $\vec F$, and the B-orbits, lying in the
1655: plane perpendicular to $\vec B$.}
1656: \label{fig_c1}
1657: \end{figure}
1658:
1659: We now analyse the structure of phase space for increasing $B$. When
1660: $B=0$, the orbits of \eqref{cB9} follow the parallels of the sphere. The
1661: poles and all points of the equator $K=0$ are fixed points. For slightly
1662: positive $B$, a resonance of order $2$ is created: only the points $\wK =
1663: 0,\frac\pi2, \pi, \frac{3\pi}2$ of the equator remain fixed. A
1664: straightforward stability analysis shows that the points $(\wK,K)=(0,0)$
1665: and $(\pi,0)$ are always elliptic, while the points $(\frac\pi2,0)$ and
1666: $(\frac{3\pi}2,0)$ are hyperbolic for
1667: \begin{equation}
1668: \label{cB11}
1669: B \leqs B_1 = 6\sqrt2 F\Lambda
1670: \sqrt{\frac{\Lambda^2-\Je^2}{5\Lambda^2+\Je^2}},
1671: \end{equation}
1672: and elliptic for $B\geqs B_1$. If $\Je\neq 0$, there is another pair of
1673: equilibrium points, located at $\wK=\frac\pi2, \frac{3\pi}2$ and
1674: $K=K_\star$, given by the condition
1675: \begin{equation}
1676: \label{cB12}
1677: \Bigpar{6\frac FB\Lambda}^2 = 1 + \frac32
1678: \frac{\Lambda^2+\Je^2-K_\star^2}{\sqrt{\brak{(\Lambda-\Je)^2-K_\star^2}
1679: \brak{(\Lambda+\Je)^2-K_\star^2}}}.
1680: \end{equation}
1681: These orbits are created in a pitchfork bifurcation at the poles at $B=0$
1682: (\figref{fig_c1}a), move to the equator as $B$ increases
1683: (\figref{fig_c1}b), and disappear in another pitchfork bifurcation at
1684: $B=B_1$ (\figref{fig_c1}c). It is easy to see that there must be a global
1685: bifurcation involving a saddle connection between these values. It is given
1686: by the condition
1687: \begin{equation}
1688: \label{cB13}
1689: \avvrg{H}_{\Lambda,\Je}(\tfrac\pi2,0) =
1690: \avvrg{H}_{\Lambda,\Je}(\wK,\Lambda-\abs{\Je}) \quad\Rightarrow\quad
1691: B = B_2 = 6\sqrt2 F\Lambda
1692: \sqrt{\frac{\Lambda-\abs{\Je}}{5\Lambda+\abs{\Je}}}.
1693: \end{equation}
1694: At $B=B_2$, the poles of the sphere are connected with the points
1695: $(\frac\pi2,0)$ and $(\frac{3\pi}2,0)$ on the equator by heteroclinic
1696: orbits. In the case $\Je=0$, all these bifurcations collapse.
1697:
1698: Our analysis of the doubly-averaged Hamiltonian \eqref{cB8} has thus
1699: revealed a rather rich structure of phase space. We point out that only the
1700: first case, depicted in \figref{fig_c1}a, is compatible with the hypothesis
1701: $B\ll F$, which is necessary for the doubly-averaged system to be a
1702: reliable approximation. We will see in the next sections, however, that the
1703: picture given in \figref{fig_c1} also contains some truth in the other
1704: parameter ranges (see \figref{fig_p3}).
1705:
1706: We conclude that for $B \ll F$, the structure of phase space is determined
1707: by four types of orbits:
1708: \begin{enum}
1709: \item The poles of the sphere \eqref{cB10} correspond to the fixed points
1710: of the averaged Stark Hamiltonian, and thus to the periodic S-orbits of the
1711: original (unaveraged) Hamiltonian \eqref{cB1}. They are unstable unless
1712: $\Je=0$.
1713:
1714: \item The points $K=0$, $\wK=\frac\pi2$ or $\frac{3\pi}2$, which are
1715: hyperbolic for small $B$, correspond to $\W=\frac\pi2$ or $\frac{3\pi}2$ in
1716: Delaunay variables, and hence describe the B-orbits, which lie in the plane
1717: perpendicular to $\vec B$. In the singly-averaged system, they appear as
1718: periodic orbits, following a curve of constant $\Je$. Since $K=0$, all
1719: these curves agglomerate at the points $G=0$ and $\w=\frac\pi2$ or
1720: $\frac{3\pi}2$ (see \figref{fig_s1b}). In the original system, the orbits
1721: can be interpreted as a fast rotation along a slowly ``breathing'' Kepler
1722: ellipse, reaching periodically the eccentricity $e=1$, where the electron
1723: approaches the nucleus indefinitely closely.
1724:
1725: \item The points $K=0$, $\wK=0$ or $\pi$, which are elliptic, correspond
1726: to $\W=0$ or $\pi$, and thus describe orbits in the plane of $\vec F$ and
1727: $\vec B$. As in the previous case, they evolve on a level curve of $\Je$
1728: containing a point with zero angular momentum. We will call them the
1729: BF-orbits.
1730:
1731: \item Finally, the four points $(\pm\frac\pi2,\pm K_\star)$ describe more
1732: complicated orbits, which are stable and provide a connection between
1733: S-orbits and B-orbits. For small $B$, they are close to the S-orbits. Let
1734: us thus call them SB-orbits.
1735: \end{enum}
1736:
1737: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1738:
1739: \subsection{The case $F\ll B$}
1740: \label{ssec_cF}
1741:
1742: When perturbing the Zeeman effect, it seems more appropriate to use
1743: coordinates $(x',y',z')$ in which the magnetic field is vertical. The
1744: Hamiltonian takes the form
1745: \begin{equation}
1746: \label{cF1}
1747: H = \frac12 p^2 - \frac1r + \frac12 B L_{z'} +
1748: \frac18 B^2 (x^{\prime2}+y^{\prime2})
1749: + F x'.
1750: \end{equation}
1751: We will denote by $(\Lambda',G',K';M',\w',\W')$ the associated Delaunay
1752: variables. The transformation between these Delaunay variables and those
1753: defined with respect to $(x,y,z) = (z',y',x')$ can be derived by expressing
1754: $\vec L$ and $z$ in both sets of variables. The result is
1755: \begin{align}
1756: \nonumber
1757: \Lambda &= \Lambda' &
1758: G &= G' \\
1759: \nonumber
1760: K &= \sqrt{G^{\prime2}-K^{\prime2}}\sin\W' &
1761: \sin^2 i &= \cos^2\W' + \cos^2 i' \sin^2\W' \\
1762: \label{cF2}
1763: \cos\W &= \frac{\sqrt{G^{\prime2}-K^{\prime2}}\cos\W'}
1764: {\sqrt{G^{\prime2}\cos^2\W'+K^{\prime2}\sin^2\W'}} &
1765: \cos\w &= -\frac{\sin\w'\cos\W'+\cos i'\cos\w'\sin\W'}{\sin i} \\
1766: \nonumber
1767: \sin\W &= \frac{K'}{\sqrt{G^{\prime2}\cos^2\W'+K^{\prime2}\sin^2\W'}} &
1768: \sin\w &= \frac{\cos\w'\cos\W'-\cos i'\sin\w'\sin\W'}{\sin i}.
1769: \end{align}
1770: The averaged Hamiltonian over $M'$ takes the form
1771: \begin{equation}
1772: \label{cF3}
1773: \begin{split}
1774: \avrg{H}_\Lambda &= -\frac1{2\Lambda^2} + \frac B2 K +
1775: B^2\avrg{H_1}_\Lambda(G',K';\w')
1776: + F \avrg{H_2}_\Lambda(G',K';\w',\W') \\
1777: \avrg{H_1}_\Lambda &= \tfrac1{16}
1778: \bigbrak{(1+\cos^2 i')(1+\tfrac32 e^{\prime2}) +
1779: \tfrac52 e^{\prime2} \sin^2 i' \cos 2\w'} \\
1780: \avrg{H_2}_\Lambda &= -\tfrac32 \Lambda^2 e' (\cos\w'\cos\W' -
1781: \sin\w'\sin\W'\cos i').
1782: \end{split}
1783: \end{equation}
1784: It generates the equations of motion
1785: \begin{align}
1786: \nonumber
1787: \dot{K}' &= \phantom{B^2\poisson{G'}{\avrg{H_1}_\Lambda} +{}}
1788: F\poisson{K'}{\avrg{H_2}_\Lambda} &
1789: \dot{\W}' &= \frac B2 + B^2\poisson{\W'}{\avrg{H_1}_\Lambda} +
1790: F\poisson{\W'}{\avrg{H_2}_\Lambda} \\
1791: \label{cF4}
1792: \dot{G}' &= B^2\poisson{G'}{\avrg{H_1}_\Lambda} +
1793: F\poisson{G'}{\avrg{H_2}_\Lambda} &
1794: \dot{\w}' &= \phantom{\frac B2 +{}} B^2\poisson{\w'}{\avrg{H_1}_\Lambda} +
1795: F\poisson{\w'}{\avrg{H_2}_\Lambda}.
1796: \end{align}
1797: We observe that for $F\ll B$, $\W'$ evolves on a faster time scale that
1798: $\w'$, and the dynamics can be further approximated by averaging the
1799: Hamiltonian over $\W'$. Note, however, that the average of
1800: $\avrg{H_2}_\Lambda$ over $\W'$ vanishes, while $\avrg{H_1}_\Lambda$ does
1801: not depend on $\W'$. The twice averaged Hamiltonian is thus strictly
1802: equivalent to the averaged Zeeman Hamiltonian \eqref{za4}. This means that
1803: to lowest order in perturbation theory, the electric field does not
1804: influence the phase portrait of the Zeeman effect, which has one of the two
1805: behaviours indicated in \figref{fig_z2}. The phase space is thus organized
1806: around three types of periodic orbits:
1807: \begin{enum}
1808: \item a stable B-orbit at $K'=G'$, located in the plane perpendicular to
1809: $\vec B$;
1810: \item a circular C-orbit, stable for large $K'$ and unstable for small $K'$;
1811: \item and a pair of stable Z-orbits, existing only for small $K'$, with a
1812: major axis perpendicular to the line of nodes.
1813: \end{enum}
1814: These orbits will experience deformations of magnitude $\Order{F^2}$ when
1815: $F>0$.
1816:
1817: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1818:
1819: \subsection{The B-orbits}
1820: \label{ssec_cq}
1821:
1822: Let us now examine the most important periodic orbits for general (small)
1823: values of $F$ and $B$. A special role is played by the orbits in the plane
1824: perpendicular to $\vec B$, which we called B-orbits. Their existence
1825: can already be seen on the original Hamiltonian \eqref{cr1}, which leaves
1826: the plane $x=0$ invariant. The Hamiltonian restricted to this plane has two
1827: degrees of freedom, its averaged version will thus have one degree of
1828: freedom. Hence {\em all} orbits starting in that plane will look periodic in
1829: the averaged system, but may correspond to quasiperiodic or soft chaotic
1830: components of the original Hamiltonian.
1831:
1832: In the two previous sections, we have found that the B-orbits are
1833: hyperbolic in the limit $B\to 0$ and elliptic in the limit $F\to 0$. Hence
1834: there must be at least one bifurcation value between these limits. The
1835: twice averaged Hamiltonian in electric action--angle variables \eqref{cB8}
1836: suggested that this transition should be given by the condition
1837: \eqref{cB11}, which is, however, not in the range where \eqref{cB8} can be
1838: expected to be a good approximation. We will now examine this question in
1839: more detail with the once averaged Hamiltonian.
1840:
1841: \begin{figure}
1842: \centerline{\psfig{figure=fig07.eps,height=50mm,clip=t}}
1843: \vspace{1mm}
1844: \caption[]
1845: {Orbits of the averaged Hamiltonian \eqref{cq1} restricted to the plane
1846: perpendicular to $\vec B$. (a) shows a case with $B<F$, and (b) shows a
1847: case with $F<B$. There are two types of orbits, those which cross the
1848: lines $\we=0$ and $\pi$, and those which oscillate around $\we=\frac\pi2$
1849: and $\frac{3\pi}2$. The second type will not appear on a Poincar\'e
1850: section at $\we=0$.}
1851: \label{fig_B1}
1852: \end{figure}
1853:
1854: Using the expansion \eqref{sc21} of $L_x$ for small $K$, it is easy to
1855: compute the equations of motion \eqref{cB3} of the singly-averaged
1856: Hamiltonian in electric action--angle variables. One can then check that
1857: both $\dot\wK$ and $\dot K$ vanish for $K=0$ and $\cos\wK=0$, which confirms
1858: the invariance of the subspace of B-orbits. The motion in this subspace is
1859: determined by the one-degree-of-freedom Hamiltonian
1860: \begin{align}
1861: \nonumber
1862: \avrg{H}_\Lambda(\Je,0;\we,\tfrac\pi2)
1863: =& -3F\Lambda\Je + \tfrac34 F^2\Lambda^4\Je^2 +\Order{F^3}\\
1864: \nonumber
1865: &-\tfrac12 B \sqrt{\Lambda^2-\Je^2} \sin\we + \tfrac3{32} B^2\Lambda^2
1866: \bigbrak{\Lambda^2+\Je^2+(\Lambda^2-\Je^2)\cos 2\we} \\
1867: &+\Order{BF}.
1868: \label{cq1}
1869: \end{align}
1870: Up to the remainders, its equations of motion are given by
1871: \begin{equation}
1872: \label{cq2}
1873: \begin{split}
1874: \dot\Je &= \frac12 B\sqrt{\Lambda^2-\Je^2} \cos\we
1875: + \frac3{16} B^2 \Lambda^2(\Lambda^2-\Je^2)\sin 2\we \\
1876: \dot\we &= -3F\Lambda + \frac32 F^2\Lambda^4\Je
1877: + \frac12 B \frac{\Je}{\sqrt{\Lambda^2-\Je^2}}\sin\we
1878: +\frac38 B^2\Lambda^2\Je\sin^2\we.
1879: \end{split}
1880: \end{equation}
1881: This system admits two elliptic equilibria located at $\we=\frac\pi2$ and
1882: $\frac{3\pi}2$ and
1883: \begin{equation}
1884: \label{cq3}
1885: \Je \simeq \pm\Lambda\Bigbrak{1+\Bigpar{\frac{B}{6F\Lambda}}^2}^{-1/2}.
1886: \end{equation}
1887: These points move from the boundaries $\abs{\Je}=\Lambda$ to $\Je=0$ as
1888: $\frac BF$ increases from $0$ to $\infty$.
1889: %
1890: \begin{figure}
1891: \centerline{\psfig{figure=fig08.eps,height=50mm,clip=t}}
1892: \vspace{1mm}
1893: \caption[]
1894: {The two types of equatorial orbits: (a) singular orbits, which
1895: periodically reach an eccentricity $e=1$, and (b), regular orbits which
1896: are bounded away from $e=1$. The electric field is vertical, while the
1897: magnetic field points out of the plane. One can identify the fast motion
1898: along a Kepler ellipse, the parameters of which evolve approximately on a
1899: level curve of \figref{fig_s1} in the regular case, and of
1900: \figref{fig_s1b} in the singular case. These orbits are not periodic in
1901: general, as can be seen on the pictures.}
1902: \label{fig_B2}
1903: \end{figure}
1904: %
1905: The orbits of \eqref{cq2} are shown in \figref{fig_B1}. They are of two
1906: types:
1907: \begin{enum}
1908: \item the orbits which cross the lines $\we=0$ and $\pi$, and for which
1909: $\we$ is monotonous;
1910: \item the orbits which oscillate around $\we=\frac\pi2$ and $\frac{3\pi}2$
1911: without reaching $\we=0$ or $\pi$.
1912: \end{enum}
1913: There is no separatrix between both types of orbits, the flow being regular
1914: when expressed in the right variables (the boundaries $\abs{\Je}=\Lambda$
1915: should again be contracted into the poles of a sphere, since they correspond
1916: to the periodic S-orbits of the Stark effect). Still, there is an important
1917: qualitative difference between both types of orbits. Indeed, by \eqref{sc4},
1918: the values $\we=0, \pi$ are the only ones that lead to an eccentricity $e=1$
1919: when $K=0$, and thus imply a close encounter with the nucleus. Hence, the
1920: first type of orbits will contain cusps (\defwd{singular} orbits,
1921: \figref{fig_B2}a), and the second will not (\defwd{regular} orbits,
1922: \figref{fig_B2}b). The boundary between both types of orbits, when starting
1923: on the line $\we=\frac\pi2$, is given by the condition
1924: \begin{equation}
1925: \label{cq4}
1926: \avrg{H}_\Lambda(\Lambda,0;\tfrac\pi2,\tfrac\pi2)
1927: =\avrg{H}_\Lambda(\Je,0;\tfrac\pi2,\tfrac\pi2)
1928: \quad\Rightarrow\quad
1929: B\simeq 6F\Lambda\sqrt{\frac{\Lambda-\Je}{\Lambda+\Je}}.
1930: \end{equation}
1931:
1932: We will now examine the stability of the $B$-orbits in the $4$-dimensional
1933: phase space of the averaged Hamiltonian. To do this, we first need to
1934: compute the derivatives
1935: \begin{equation}
1936: \label{cq5}
1937: \begin{split}
1938: \dpar{\dot\wK}{\wK} \Bigevalat{\frac\pi2,0} = &
1939: \frac12 B \frac{\Je}{\sqrt{\Lambda^2-\Je^2}}\cos\we + \frac14
1940: B^2\Lambda^2\Je\sin2\we
1941: = - \dpar{\dot K}{K} \Bigevalat{\frac\pi2,0}\\
1942: \dpar{\dot\wK}{K} \Bigevalat{\frac\pi2,0} = &
1943: \frac92 F^2\Lambda^4 + \frac B2 \frac{\Lambda^2}{(\Lambda^2-\Je^2)^{3/2}}
1944: \sin \we\\
1945: &-\frac1{16} B^2 \frac{\Lambda^2}{\Lambda^2-\Je^2}
1946: \bigbrak{5\Lambda^2+\Je^2 + (3\Lambda^2-\Je^2)\cos 2\we} \\
1947: \dpar{\dot K}{\wK} \Bigevalat{\frac\pi2,0} = &
1948: -\frac12 B\sqrt{\Lambda^2-\Je^2} \sin\we
1949: + \frac18 B^2\Lambda^2(\Lambda^2-\Je^2)
1950: \bigbrak{3 + 2\cos 2\we}.
1951: \end{split}
1952: \end{equation}
1953: In order to determine the stability of a B-orbit, we have to find the
1954: multipliers of the variational equation
1955: \begin{equation}
1956: \label{cq6}
1957: \dot{z} = A(\we,\Je)z,
1958: \end{equation}
1959: integrated over a solution of the system \eqref{cq2}, where $z^T =
1960: (\wK-\frac\pi2,0)$, and $A$ is the matrix with entries given by \eqref{cq5}.
1961:
1962: Two limiting cases can be studied analytically. When $B\to 0$, $A$ can be
1963: replaced by its average over the fast variable $\we$, and we obtain again
1964: the condition \eqref{cB11}, which tells us that only orbits with
1965: $\Lambda^2-\Je^2 = \Order{B^2/F^2}$ are elliptic. The other case is that of
1966: the orbits \eqref{cq3}, for which both $\we$ and $\Je$ are constant. In this
1967: case, we find $\det A>0$, which means that the orbit is elliptic.
1968:
1969: \begin{figure}
1970: \centerline{\psfig{figure=fig09.eps,height=55mm,clip=t}}
1971: \vspace{1mm}
1972: \caption[]
1973: {Stability diagrams of the equatorial orbit, obtained by computing
1974: numerically the multipliers of equation \eqref{cq6}, starting with
1975: $\we=\frac\pi2$ and different values of $\Je$, $F$ and $B$. (a) shows the
1976: case $\Je=0$ (b) the case $F=0.1$. Shaded regions indicate unstable orbits.
1977: The curves are: I the location of the stable orbit \eqref{cq3}, II the
1978: boundary \eqref{cq4} between regular and singular orbits (orbits with large
1979: $B$, large $\Je$ or small $F$ are regular), and III the stability boundary
1980: \eqref{cB11} of the doubly-averaged system.}
1981: \label{fig_B3}
1982: \end{figure}
1983:
1984: \figref{fig_B3} shows numerically computed stability diagrams for general
1985: parameter values (one should note that equation \eqref{cq6} is much easier
1986: to treat numerically than the full equations of motion). As expected,
1987: orbits are unstable in a region compatible with condition \eqref{cB11} and
1988: stable for $B\gg F$. There appears to be a second zone of instability for
1989: intermediate values of $\frac BF$. However, this region corresponds to rather
1990: large values of $B$, for which the averaged Hamiltonian is not necessarily a
1991: good approximation. This confirms the picture that for small fields, the
1992: B-orbits are stable for small $F$ and unstable for small $B$, with a linear
1993: transition line between both regimes.
1994:
1995: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1996:
1997: \subsection{The structure of phase space}
1998: \label{ssec_cp}
1999:
2000: The phase space of the averaged Hamiltonian is four-dimensional, and
2001: depends on the parameters $\Lambda$, $F$ and $B$. The discussion in the
2002: previous sections suggests that the global structure of phase space will
2003: mainly depend on the ratio $\frac BF$; we expect that increasing both
2004: fields, while keeping their ratio constant, will mainly result in an
2005: increase of the size of chaotic components, without changing the location
2006: and stability of the main periodic orbits.
2007:
2008: The equations of motion are given by \eqref{cB3}, where the expressions of
2009: the linear and quadratic parts in $B$ are deduced from \eqref{sc18} and
2010: \eqref{cB6} respectively. The manifold of constant energy is
2011: three-dimensional, and can be represented by a Poincar\'e section. The
2012: structure of the equations of motion shows that, at least when $\frac BF$ is not
2013: too large, a surface of section of the form $\we=$ constant will be a good
2014: choice.
2015:
2016: \begin{figure}
2017: \centerline{\psfig{figure=fig10.eps,width=148mm,clip=t}}
2018: \vspace{1mm}
2019: \caption[]
2020: {Poincar\'e sections of the singly-averaged Hamiltonian, (a) at $\we=0$ and
2021: (b) at $\we=\frac\pi2$, for $B=F=0.1$, $\Lambda=1$ and $\avrg{H}_\Lambda -
2022: H_0 = -0.15$ (which corresponds roughly to $\Je\simeq 0.5$). Hyperbolic
2023: points are the B-orbits, elliptic points the BF-orbits. }
2024: \label{fig_p1}
2025: \end{figure}
2026:
2027: \figref{fig_p1} shows Poincar\'e sections taken at $\we=0$ and $\frac\pi2$,
2028: in a case with $\frac BF=1$ (sections with a smaller ratio of $\frac BF$
2029: look similar). This section should be compared with \figref{fig_c1}a. The
2030: hyperbolic points located at $(\wK,K)=(\frac\pi2,0)$ and $(\frac{3\pi}2,0)$
2031: correspond to the B-orbits. Their stable and unstable manifolds separate
2032: phase space into two regions, corresponding respectively to oscillations
2033: around the elliptic BF-orbits and around the S-orbits. In fact, we expect
2034: chaotic motions to show up near the separatrices, but they occupy a very
2035: small area for these values of the fields.
2036:
2037: In contrast to the B-orbits, the BF-orbits move in the $(\wK,K)$-plane as
2038: $\we$ varies. This effect did not show up in the doubly-averaged
2039: approximation, and is mainly due to terms linear in $B$ of the Hamiltonian.
2040: \figref{fig_p1} shows that the BF-orbits are roughly located at
2041: $K\simeq\alpha\cos\we$, $\wK\simeq\beta\sin\we$ and $K\simeq-\alpha\cos\we$,
2042: $\wK\simeq\pi-\beta\sin\we$, where $\alpha$ and $\beta$ are of order $\frac BF$
2043: (they can be estimated by inserting Fourier series in the equations of
2044: motion).
2045:
2046: \begin{figure}
2047: \centerline{\psfig{figure=fig11.eps,height=40mm,clip=t}}
2048: \vspace{1mm}
2049: \caption[]
2050: {BF-orbits, obtained by integrating the equations of motion of the
2051: unaveraged Hamiltonian, (a) for $F=B=0.05$ and (b,c) for $F=0.05$ and
2052: $B=0.1$. Initial conditions are given by \eqref{cp1} with $e=0.2$. In (a)
2053: and (b), $\vec F$ is vertical and $\vec B$ points out of the plane, (c)
2054: shows a projection on the $(\vec B,\vec F)$-plane.}
2055: \label{fig_p2}
2056: \end{figure}
2057:
2058: In order to understand the geometry of BF-orbits, we observe that
2059: \begin{enum}
2060: \item if $\we=0$, $K$ reaches its maximum and $\wK=0$; we know by
2061: \eqref{sc4} that in this case the eccentricity is maximal and $\cos\w=0$;
2062: since $K\neq 0$, however, $e$ is strictly smaller than 1, and thus the
2063: orbit no longer approaches the nucleus, as it does for $B=0$; \eqref{sc20}
2064: shows that $\W=\frac\pi2$ or $\frac{3\pi}2$ (depending on the sign of
2065: $\Je$); this means that the plane of the Kepler ellipse contains $\vec
2066: B\wedge\vec F$, and since $\cos\w=0$, its major axis is in the plane $(\vec
2067: B,\vec F)$.
2068:
2069: \item if $\we=\frac\pi2$, $K=0$ and $\wK=\beta$; we also have
2070: $\cos\w=0$, but this time the eccentricity is minimal; \eqref{sc20}
2071: shows that $\sin\W=\Order{\beta}$, meaning that the plane of the ellipse
2072: contains $\vec F$, but is slightly rotated with respect to the $(\vec B,\vec
2073: F)$-plane, by an amount of order $\frac BF$.
2074: \end{enum}
2075: In fact, when $\frac BF$ is small, we can deduce from \eqref{sc20} that
2076: $\W$ is close to $0$ or $\pi$ for most values of $\we$. It approaches a
2077: step function of the form $\frac\pi2\brak{1-\sign(\sin\we)}$ as $\frac BF$
2078: tends to zero, which means that the orbit approaches the $(\vec B,\vec
2079: F)$-plane in this limit. For increasing $B$, however, the orbit gains some
2080: thickness in the direction perpendicular to the plane, and rotates around
2081: $\vec B$ (\figref{fig_p2}). BF-orbits are thus truly non-planar when $B>0$.
2082:
2083: Initial conditions producing BF-orbits can be constructed in the following
2084: way. Pick an eccentricity $e\in[0,1)$. Starting with $\we=0$, we have $K=0$
2085: and $\W=\wK=\beta$ can be read off the Poincar\'e section (though, the orbit
2086: being elliptic, starting with a slightly wrong $\wK$ will not have dramatic
2087: consequences). Taking $M=0$ as initial position on the ellipse, the initial
2088: conditions are then obtained from \eqref{zd9} to be
2089: \begin{align}
2090: \nonumber
2091: x &= 0 &
2092: p_x &= -\cos\W \tfrac1\Lambda \sqrt{\tfrac{1+e}{1-e}} \\
2093: \label{cp1}
2094: y &= 0 &
2095: p_y &= -\sin\W \tfrac1\Lambda \sqrt{\tfrac{1+e}{1-e}} \\
2096: \nonumber
2097: z &= \Lambda^2 (1-e) &
2098: p_z &= 0.
2099: \end{align}
2100: We can also start with $\we=0$, $M=0$, and read $K=\alpha$ off the Poincar\'e
2101: section, and take as initial conditions
2102: \begin{align}
2103: \nonumber
2104: x &= 0 &
2105: p_x &= -\tfrac1\Lambda \sqrt{\tfrac{1+e}{1-e}} \\
2106: \label{cp2}
2107: y &= \Lambda K \sqrt{\tfrac{1+e}{1-e}} &
2108: p_y &= 0 \\
2109: \nonumber
2110: z &= \Lambda \sqrt{\Lambda^2(1-e^2) - K^2} \sqrt{\tfrac{1+e}{1-e}} &
2111: p_z &= 0.
2112: \end{align}
2113:
2114: The boundaries of the section are given by the condition $\abs{K} = \Lambda
2115: - \abs{\Je}$, and depend on $\wK$ and $\we$ because $\Je$ is no longer
2116: constant. They correspond to the location of the S-orbit. The equations of
2117: motion become singular as $\abs{K}\to\Lambda-\abs{\Je}$, which makes
2118: unreliable the numerical computation of orbits approaching the S-orbit.
2119: This singularity can be tamed by a canonical transformation
2120: $(\Je,K;\we,\wK)\mapsto(\Je,J;\phi,-\wK)$. In the case $K\to\Lambda-\Je$,
2121: $\Je\geqs 0$, for instance, it is given by $J=\Lambda-\Je-K$ and
2122: $\phi=\we-\wK$, the other cases being similar. We will not elaborate on
2123: this point here.
2124:
2125: \begin{figure}
2126: \centerline{\psfig{figure=fig12.eps,width=148mm,clip=t}}
2127: \vspace{1mm}
2128: \caption[]
2129: {Poincar\'e sections at $\we=\frac\pi2$ of the averaged system,
2130: represented on the sphere $\kappa_1^2+\kappa_2^2+\kappa_3^2 =
2131: (\Lambda-\abs{\Je})^2$ for increasing values of $\frac BF$, compare
2132: \figref{fig_c1}. In all cases, $\Lambda=1$, $\avrg{H}_\Lambda = -\frac12
2133: - \frac32 F$, and $B=0.1$. (a) $F=0.1$, (b) $F=0.05$, (c) $F=0.038$ and (d)
2134: $F=0.03$. Points are sparse near the poles, as we only plot points obtained
2135: with sufficient numerical accuracy.}
2136: \label{fig_p3}
2137: \end{figure}
2138:
2139: \begin{figure}
2140: \centerline{\psfig{figure=fig13.eps,width=148mm,clip=t}}
2141: \vspace{1mm}
2142: \caption[]
2143: {Poincar\'e sections of the averaged Hamiltonian for increasing values of
2144: $\frac BF$. (a,c,e) are taken at $\we=0$ and (b,d,f) at $\we=\frac\pi2$.
2145: Values of $\Lambda$, $H$ and $B$ are the same as in \figref{fig_p3}, and
2146: (a,b) $F=0.05$, (c,d) $F=0.038$, (e,f) $F=0.03$. The equilibrium points
2147: $K=0$, $\wK=\frac\pi2$ and $\frac{3\pi}2$ are B-orbits, and the elliptic
2148: points near $\wK=0,\pi$ are BF-orbits. The SB-orbits can be recognized on
2149: figures (b), (c) and (d).}
2150: \label{fig_p4}
2151: \end{figure}
2152:
2153: \figref{fig_p3} and \figref{fig_p4} show phase portraits for increasing
2154: values of $\frac BF$. They mainly differ by the behaviour of the B-orbit
2155: and its stable and unstable manifolds. While for small $\frac BF$, these
2156: manifolds connect the B-orbits (\figref{fig_p1} and \figref{fig_p3}a), they
2157: connect each B-orbit to an S-orbit for larger $\frac BF$ (\figref{fig_p3}b
2158: and \figref{fig_p4}a,b). When this ratio increases further, each manifold
2159: folds back to the B-orbit it emerges from, enclosing an elliptic orbit that
2160: we called SB (\figref{fig_p3}c and \figref{fig_p4}c,d). Finally, the
2161: SB-orbits bifurcate with the B-orbit, which becomes elliptic, as we already
2162: know (\figref{fig_p3}d and \figref{fig_p4}e,f). This scenario is very close
2163: to the scenario obtained for the doubly-averaged system, compare
2164: \figref{fig_c1}.
2165:
2166: The BF-orbit exists and remains elliptic for fairly large values of $\frac
2167: BF$. It is possible that it exists for all values of the field, and
2168: transforms into the Z-orbit or C-orbit of the Zeeman effect as $F\to 0$.
2169: Verifying this conjecture would, however, require an understanding of the
2170: transition between the Poincar\'e sections at $\we=\text{constant}$, valid
2171: for small and moderate $\frac BF$, and those at $\W'=\text{constant}$, valid
2172: at large $\frac BF$. At least, it is true that the geometry of BF-orbits for
2173: $\we=0$ is compatible with that of the Z-orbits for $\W'=\frac\pi2$.
2174:
2175: The structure of phase space is thus essentially organized around the
2176: B-orbits, BF-orbits and S-(or SB-)orbits, which are periodic orbits of the
2177: averaged Hamiltonian. The other orbits of the averaged system are either
2178: quasiperiodic with two frequencies, or belong to resonances or chaotic
2179: components, which are small, however, for the parameter values compatible
2180: with the averaging approximation. Intermediate values of $\frac BF$ are the
2181: most favorable for diffusion in phase space, as the heteroclinic connections
2182: between B- and S-orbits allow fast transitions between various regions of
2183: phase space.
2184:
2185: The B-orbits are called $S_+$ and $S_-$ by \cite{FW96}, who also describe a
2186: periodic orbit located in the plane perpendicular to $\vec F$, called
2187: $S_\perp$. We did not find such an orbit (which would correspond to $\Je=0$
2188: and $\we=0$ or $\pi$) in the parameter range we investigated. It might be
2189: that the BF-orbit behaves as the $S_\perp$-orbit for $F\ll B$.
2190:
2191: Note that going from the averaged to the unaveraged dynamics will add a
2192: time scale. Hence, periodic orbits of the averaged system may correspond to
2193: quasiperiodic orbits of the unaveraged Hamiltonian with two frequencies, or
2194: to soft chaotic components. Some of the quasiperiodic orbits of the
2195: averaged system will support KAM-tori of the full system, which correspond
2196: to a quasiperiodic motion with three frequencies.
2197:
2198: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2199:
2200: \section{Conclusions and outlook}
2201:
2202: Action--angle variables and the technique of averaging help to understand
2203: the dynamics of the hydrogen atom in crossed electric and magnetic fields in
2204: two ways.
2205:
2206: The first way is in terms of adiabatic invariants. The semi-major axis
2207: $\Lambda^2$ of the Kepler ellipse is a constant of motion of the averaged
2208: Hamiltonian, and thus an adiabatic invariant of the full Hamiltonian on the
2209: time scale $B^{-2}$, compare \eqref{cr1}. The same role is played by the
2210: averaged perturbing function $\avrg{H}_\Lambda-H_0$. In the case $B\ll F$, we
2211: have the additional invariants $\Je$ and $\avvrg{H}_{\Lambda,\Je}$ (see
2212: \eqref{cB6}), which evolve on the time scale $B^{-1}$. In the case $F\ll
2213: B$, $K'=\vec L\cdot\vec B/B$ and the averaged quadratic Zeeman term are
2214: adiabatic invariants on the time scale $F^{-1}$.
2215:
2216: The second way is in terms of periodic orbits of the averaged system, which
2217: organize the structure of phase space. The B-orbits, contained in the plane
2218: perpendicular to $\vec B$, exist for all fields. They are unstable for
2219: $B\to 0$ and stable for $F\to 0$, with a roughly linear transition line in
2220: the $(F,B)$-plane. In the limit $B\to 0$, we also find the S-orbits of the
2221: Stark effect, and the stable BF-orbits contained in the plane of $\vec B$
2222: and $\vec F$. As $B$ increases, S-orbits appear to become unstable by
2223: expelling a pair of ``SB-orbits'', which are then absorbed by the B-orbits.
2224: BF-orbits are non-planar for $B>0$. They exist and are stable in a large
2225: domain of field values. It is unclear whether they transform into one of
2226: the periodic orbits of the pure Zeeman effect as $F\to 0$.
2227:
2228: Diffusion in phase space is most prominent in the regime where $F$ and $B$
2229: are of comparable magnitude, when neighbourhoods of the S-orbits are
2230: connected by heteroclinic orbits. In the other regions of phase space,
2231: close to the stable orbits, orbits of the averaged system are trapped
2232: inside KAM-tori. In the real three-degrees-of-freedom system, however,
2233: diffusion becomes possible in these regions as well, although on a longer
2234: time scale. This aspect of dynamics still has to be better understood, as
2235: do the quantum and semiclassical mechanics of the problem which continue
2236: to be of interest \cite{vMU97,vMFU97b,NUFESWHD97,RT97,CZRT99}.
2237:
2238: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2239:
2240: \section*{Acknowledgments}
2241:
2242: This work originated during a very pleasant stay of the first author in the
2243: group of Jacques Laskar, at the Bureau des Longitudes in Paris. We thank
2244: him for introducing us to the realm of perturbed Kepler problems, Delaunay
2245: variables, and averaging. NB was supported by the Fonds National Suisse de
2246: la Recherche Scientifique.
2247:
2248: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2249:
2250: \appendix
2251:
2252: \section{Averaging}
2253: \label{app_av}
2254:
2255: A standard result on averaging \cite{V96} is
2256:
2257: \begin{theorem}
2258: %\label{}
2259: Consider the initial value problem
2260: \begin{equation}
2261: \label{av1}
2262: \dot{x} = \eps f(x,t) + \eps^2 g(x,t,\eps),
2263: \qquad x(0)=x_0,
2264: \end{equation}
2265: where $f(x,t)$ is $2\pi$-periodic in $t$, and $x, x_0\in\R^n$. Define the
2266: averaged system
2267: \begin{equation}
2268: \label{av2}
2269: \begin{split}
2270: \dot{y} &= \eps\avrg{f}(y),
2271: \qquad y(0)=x_0,\\
2272: \avrg{f}(y) &= \frac1{2\pi} \int_0^{2\pi} f(y,t)\dx t.
2273: \end{split}
2274: \end{equation}
2275: If $f$ and $g$ are sufficiently smooth and bounded, then
2276: \begin{enum}
2277: \item $x(t)-y(t) = \Order{\eps}$ on the time scale $1/\eps$;
2278: \item If $y_0$ is a nondegenerate equilibrium of \eqref{av2}, then
2279: \eqref{av1} possesses an isolated periodic orbit
2280: $\gamma(t)=y_0+\Order{\eps}$, with the same stability as $y_0$ if $y_0$ is
2281: hyperbolic.
2282: \end{enum}
2283: \end{theorem}
2284:
2285: Consider now a Hamiltonian of the form
2286: \begin{equation}
2287: \label{av3}
2288: H(I_1,\dots,I_n;\ph_1,\dots,\ph_n)
2289: = H_0(I_1) + \eps H_1(I_1,\dots,I_n;\ph_1,\dots,\ph_n),
2290: \end{equation}
2291: where the $\ph_i$ are angle variables.
2292: If we introduce the Poisson bracket
2293: \begin{equation}
2294: \label{av4}
2295: \poisson{f}{g} = \sum_{i=1}^n \Bigbrak{\dpar f{\ph_i} \dpar g{I_i} - \dpar
2296: f{I_i} \dpar g{\ph_i}},
2297: \end{equation}
2298: the equations of motion can be written in the form
2299: \begin{align}
2300: \nonumber
2301: \dot{\ph_1} &= H_0'(I_1) + \eps \poisson{\ph_1}{H_1} &&\\
2302: \label{av5}
2303: \dot{\ph_j} &= \eps \poisson{\ph_j}{H_1} &\qquad j=&2,\dots,n\\
2304: \nonumber
2305: \dot{I_i} &= \eps \poisson{I_i}{H_1} &\qquad i=&1,\dots,n.
2306: \end{align}
2307: If $\poisson{\ph_1}{H_1}$ is bounded, $H_0'(I_1)\neq 0$ and $\eps$ is small
2308: enough, we may reparametrize the orbits by $\ph_1$ instead of $t$, giving
2309: \begin{equation}
2310: \label{av6}
2311: \begin{split}
2312: \dtot{\ph_j}{\ph_1} &= \eps \frac{\poisson{\ph_j}{H_1}}{H_0'(I_1)} +
2313: \Order{\eps^2} \qquad j=2,\dots,n\\
2314: \dtot{I_i}{\ph_1} &= \eps \frac{\poisson{I_i}{H_1}}{H_0'(I_1)} +
2315: \Order{\eps^2} \qquad i=1,\dots,n.
2316: \end{split}
2317: \end{equation}
2318: According to the theorem, the dynamics of this system are well approximated
2319: by those of the system averaged over the fast variable $\ph_1$,
2320: \begin{equation}
2321: \label{av7}
2322: \begin{split}
2323: \dtot{\ph_j}{\ph_1} &= \eps \frac{\poisson{\ph_j}{\avrg{H_1}}}{H_0'(I_1)}
2324: \qquad j=2,\dots,n\\
2325: \dtot{I_i}{\ph_1} &= \eps \frac{\poisson{I_i}{\avrg{H_1}}}{H_0'(I_1)}
2326: \qquad i=1,\dots,n,
2327: \end{split}
2328: \end{equation}
2329: where we have introduced
2330: \begin{equation}
2331: \label{av8}
2332: \avrg{H_1}(I_1,\dots,I_n;\ph_2,\dots,\ph_n) =
2333: \frac1{2\pi}\int_0^{2\pi} H_1(I_1,\dots,I_n;\ph_1,\dots,\ph_n) \dx \ph_1.
2334: \end{equation}
2335: We now observe that the canonical equations associated with the averaged
2336: Hamiltonian $\avrg{H} = H_0+\eps\avrg{H_1}$,
2337: \begin{align}
2338: \nonumber
2339: \dot{\ph_1} &= H_0'(I_1) + \eps \poisson{\ph_1}{\avrg{H_1}} &&\\
2340: \nonumber
2341: \dot{I_1} &= 0 && \\
2342: \label{av9}
2343: \dot{\ph_j} &= \eps \poisson{\ph_j}{\avrg{H_1}} & j&= 2,\dots n \\
2344: \nonumber
2345: \dot{I_j} &= \eps \poisson{I_j}{\avrg{H_1}} & j&= 2,\dots n
2346: \end{align}
2347: are close to $\Order{\eps^2}$ to those of the averaged system \eqref{av7}.
2348: Thus the dynamics of the averaged Hamiltonian are a good approximation, in
2349: the sense of the averaging theorem, of those of the initial Hamiltonian. In
2350: particular, the averaged Hamiltonian \eqref{av8} is an adiabatic invariant
2351: of the initial system.
2352:
2353: The averaging procedure can be extended to higher orders in $\eps$ by the
2354: method of Lie-Deprit series \cite{Deprit,Henrard}.
2355:
2356: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2357:
2358: \begin{table}
2359: \begin{center}
2360: \begin{tabular}{|c|ccccc|}
2361: \hline
2362: %\poisson{\cdot}{\cdot}
2363: \vrule height 13pt depth 7pt width 0pt
2364: & $\Lambda$ && $M$ && $\w$ \\
2365: \hline
2366: \vrule height 20pt depth 10pt width 0pt
2367: $e$ &
2368: $0$ && $\dfrac{G^2}{\Lambda^3 e}$ &&
2369: $-\dfrac{G}{\Lambda^2 e}$ \\
2370: \vrule height 20pt depth 10pt width 0pt
2371: $r$ &
2372: $-\dfrac{\Lambda^4 e}{r} \sin E$ &&
2373: $\dfrac{2r}{\Lambda} - \dfrac{G^2}{\Lambda e}\cos E$ &&
2374: $\dfrac{G}{e} \cos E$ \\
2375: \vrule height 20pt depth 10pt width 0pt
2376: $\sin^2 i$ & $0$ && $0$ && $2\dfrac{K^2}{G^3}$ \\
2377: \vrule height 20pt depth 10pt width 0pt
2378: $v$ &
2379: $-\dfrac{\Lambda^3 G}{r^2}$ &&
2380: $\dfrac{G\sin E}{re}$ &&
2381: $-\dfrac{\Lambda}{re} \sin E$ \\
2382: \vrule height 20pt depth 10pt width 0pt
2383: $X$ &
2384: $\dfrac{\Lambda^4}{r}\sin E$ &&
2385: $-\dfrac{r^2}{e\Lambda^3} - \Lambda e \sin^2 E$ &&
2386: $\dfrac Ge$ \\
2387: \vrule height 20pt depth 12pt width 0pt
2388: $Y$ &
2389: $-\dfrac{\Lambda^3}{r}G\cos E$ &&
2390: $G \sin E$ && $\Lambda \sin E$ \\
2391: \hline
2392: \end{tabular}
2393: \end{center}
2394: \caption[]
2395: {A few useful Poisson brackets involving Delaunay variables. We show
2396: Poisson brackets between columns and lines, for instance the upper right
2397: element is $\poisson{\w}e$. Brackets involving $G$, $K$ and $\W$ are
2398: trivial to compute.}
2399: \label{t_z1}
2400: \end{table}
2401:
2402: \section{Poisson brackets}
2403: \label{app_pb}
2404:
2405: In this appendix, we discuss the computation of Poisson brackets involving
2406: various sets of variables used in this paper.
2407:
2408: Since Hamiltonians are usually expressed in terms of auxiliary variables,
2409: such as the eccentricity $e$ or inclination $i$, it is useful to start by
2410: computing some Poisson brackets involving these quantities. Some of them are
2411: shown in \tabref{t_z1}.
2412:
2413: For instance, for the perturbation term $H_1=\frac1{16} r^2\bigbrak{1 +
2414: \cos^2 i + \sin^2 i \,\cos(2\w+2v)}$ of the Zeeman Hamiltonian, we find
2415: \begin{align}
2416: \nonumber
2417: \poisson{\Lambda}{H_1} =& -\frac18 \Lambda^4 e
2418: \bigbrak{1 + \cos^2 i + \sin^2 i \,\cos(2\w+2v)} \sin E
2419: + \frac18 \Lambda^3 G \sin^2 i \,\sin(2\w+2v) \\
2420: \nonumber
2421: \poisson{G}{H_1} =& \frac18 r^2 \sin^2 i\,\sin(2\w+2v) \\
2422: \nonumber
2423: \poisson{M}{H_1} =& \frac18 r \Bigpar{\frac{2r}\Lambda -
2424: \frac{G^2}{\Lambda e}\cos E}
2425: \bigbrak{1 + \cos^2 i + \sin^2 i \,\cos(2\w+2v)} \\
2426: \nonumber
2427: &- \frac18 \frac{rG}e \sin^2 i \,\sin(2\w+2v) \sin E \\
2428: \nonumber
2429: \poisson{\w}{H_1} =& \frac18 \frac{rG}{e}
2430: \bigbrak{1 + \cos^2 i + \sin^2 i \,\cos(2\w+2v)} \cos E \\
2431: &+ \frac18 \frac{r\Lambda}e \sin^2 i \,\sin(2\w+2v) \sin E
2432: + \frac18 \frac{r^2}{G} \bigbrak{-1 + \cos(2\w+2v)}.
2433: \label{pb1}
2434: \end{align}
2435: There is an apparent singularity at $e=0$. It can be removed, however, by
2436: introducing variables $J=\Lambda-G=\Order{e^2}$ and $\phi=M+\w$. The
2437: transformation $(\Lambda,G,K;M,\w,\W)\mapsto(\Lambda,J,K;\phi,-\w,\W)$ is
2438: canonical. One can check that $\dot\phi$ is finite in the limit $e\to 0$,
2439: and that the variable $\z=J\e^{\icx\w}$ satisfies $\dot\z=\Order{e}$. This
2440: shows that the circular orbit is indeed a periodic orbit of the Zeeman
2441: Hamiltonian, as suggested by the averaged system.
2442:
2443: Another way to deal with the singularity of Delaunay variables at $e=0$ in
2444: the averaged case is to use coordinates $(\x_1,\x_2,\x_3)$ introduced by
2445: \cite{CDMW87}, see equation \eqref{za7}. \tabref{t_z3} gives some useful
2446: Poisson brackets involving these variables.
2447:
2448: \begin{table}
2449: \begin{center}
2450: \begin{tabular}{|c|ccccc|}
2451: \hline
2452: \vrule height 14pt depth 8pt width 0pt
2453: & $\x_1$ && $\x_2$ && $\x_3$ \\
2454: \hline
2455: \vrule height 14pt depth 8pt width 0pt
2456: $\x_1$ & $0$ && $-2G\x_3$ && $2G\x_2$ \\
2457: \vrule height 12pt depth 8pt width 0pt
2458: $\x_2$ & $2G\x_3$ && $0$ && $-2G\x_1$ \\
2459: \vrule height 12pt depth 8pt width 0pt
2460: $\x_3$ & $-2G\x_2$ && $2G\x_1$ && $0$ \\
2461: \hline
2462: \vrule height 14pt depth 8pt width 0pt
2463: $G$ & $-\x_2$ && $\x_1$ && $0$ \\
2464: \vrule height 14pt depth 12pt width 0pt
2465: $e^2$ & $2\dfrac{1-e^2}G \x_2$ && $-2\dfrac{1-e^2}G \x_1$ && $0$ \\
2466: \vrule height 14pt depth 12pt width 0pt
2467: $\cos^2 i$ & $2\dfrac{\cos^2 i}G \x_2$ && $-2\dfrac{\cos^2 i}G \x_1$ && $0$ \\
2468: \hline
2469: \end{tabular}
2470: \end{center}
2471: \caption[]
2472: {Some Poisson brackets involving the variables $\x_i$ defined in
2473: \eqref{za7}.}
2474: \label{t_z3}
2475: \end{table}
2476:
2477: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2478:
2479: \section{Angle variables to first order in $F$}
2480: \label{app_aF}
2481:
2482: For $F>0$, the extremal values of $\x^2$ are given by the condition
2483: \begin{equation}
2484: \label{aF1}
2485: K^2 - 2\alpha_1(F)\x^2 - 2H(F)\x^4 + 2 F \x^6 = 0,
2486: \end{equation}
2487: where the quantities
2488: \begin{equation}
2489: \label{aF2}
2490: \begin{split}
2491: H(F) &= -\frac1{2\Lambda^2} - 3F\Lambda\Je + \Order{F^2} \\
2492: \alpha_{1,2}(F) &= \frac{2J_{\x,\y}+K}{\Lambda}
2493: \pm F \Lambda^2 \bigbrak{6J_\x J_\y + 3K(J_\x+J_\y)+K^2} + \Order{F^2}
2494: \end{split}
2495: \end{equation}
2496: are obtained by solving equations \eqref{saa4} perturbatively. We find that
2497: $\x^2$ varies between limits $\hat a_1(F) \pm \hat b_1(F)$ given by
2498: \begin{equation}
2499: \label{aF3}
2500: \begin{split}
2501: \hat a_1(F) &= a_1 + \tfrac12 F \Lambda^4
2502: \bigbrak{\Je^2+4\Lambda\Je-5\Lambda^2+K^2} + \Order{F^2}\\
2503: \hat b_1(F) &= b_1
2504: \bigbrak{1-\tfrac12F\Lambda^3(5\Lambda+\Je)} + \Order{F^2}.
2505: \end{split}
2506: \end{equation}
2507: Likewise, the bounded orbits of $\y^2$ vary between limits $\hat a_2(F) \pm
2508: \hat b_2(F)$ which are obtained by changing the signs of $\Je$ and $F$ in
2509: \eqref{aF3}. We may parametrize the level curves of $\alpha_{1,2}$ by
2510: \begin{equation}
2511: \label{aF4}
2512: \begin{split}
2513: \x^2 &= \hat a_1(F) - \hat b_1(F)\cos\psi \\
2514: \y^2 &= \hat a_2(F) - \hat b_2(F)\cos\chi,
2515: \end{split}
2516: \end{equation}
2517: with the momenta given by \eqref{saa3}. The derivatives of the action
2518: \eqref{saa14} can then be computed as before, with the result
2519: \begin{equation}
2520: \label{aF5}
2521: \begin{split}
2522: \wL =& \frac{\psi+\chi}2 - \frac1{2\Lambda^2} (b_1\sin\psi+b_2\sin\chi) \\
2523: &+ \tfrac14 F \bigbrak{\Lambda(13\Lambda+7\Je)b_1\sin\psi - b_1^2\sin 2\psi
2524: - \Lambda(13\Lambda-7\Je)b_2\sin\chi + b_2^2\sin 2\chi}\\
2525: &+\Order{F^2} \\
2526: \we =& \frac{\chi-\psi}2 + F\Lambda^2\bigbrak{b_1\sin\psi+b_2\sin\chi}
2527: +\Order{F^2} \\
2528: \wK =& \ph - \rho_1(\psi) - \rho_2(\chi)
2529: + \frac12 F K\Lambda^4 \Bigbrak{\frac{\sin\psi}{a_1-b_1\cos\psi} -
2530: \frac{\sin\chi}{a_2-b_2\cos\chi}} +\Order{F^2}.
2531: \end{split}
2532: \end{equation}
2533:
2534: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2535:
2536: \begin{thebibliography}{vMFU97b}
2537:
2538: \small
2539:
2540: %% New definition of \@listI for the bibliography
2541: \makeatletter
2542: \def\@listI{\leftmargin\leftmargini
2543: \topsep=\medskipamount
2544: \setlength{\parsep}{0mm}
2545: \setlength{\itemsep}{-1.5mm}
2546: }
2547: \let\@listi\@listI
2548: \@listi
2549: \makeatother
2550:
2551: \bibitem[B27]{Born} \bibbook{M.\ Born}
2552: {The Mechanics of the Atom}
2553: {G.\ Bell}
2554: {London, 1927}
2555:
2556: \bibitem[BM75]{Bohr/Mottelson}
2557: \bibbook{A.\ Bohr, B.R.\ Mottelson}
2558: {Nuclear Structure, Vol. II}
2559: {Benjamin Reading}{MA 1975}
2560:
2561: \bibitem[BS84]{Braun/Solovev}
2562: \bibarticle{P.A.\ Braun, E.A.\ Solov'ev}
2563: {The Stark Effect for a hydrogen atom in a magnetic
2564: field}
2565: {Sov.\ Phys.\ JETP}
2566: {59}{38}{46}{1984}
2567:
2568: \bibitem[C98]{Connerade}
2569: \bibbook{J.-P.\ Connerade}
2570: {Highly Excited Atoms}
2571: {Cambridge U.P.}
2572: {Cambridge, UK, 1998}
2573:
2574: \bibitem[D69]{Deprit} \bibarticle{A.\ Deprit}
2575: {Canonical transformations depending on a small
2576: parameter}
2577: {Celestial Mech.}{1}{12}{30}{1969}
2578:
2579: \bibitem[CDMW87]{CDMW87}
2580: \bibarticle{S.L.\ Coffey, A.\ Deprit, B.\ Miller,
2581: C.A.\ Williams}
2582: {The quadratic Zeeman Effect in Moderately Strong
2583: Magnetic Fields}
2584: {Ann.\ N.Y.\ Acad.\ Sci.}{497}{22}{36}{1987}
2585:
2586: \bibitem[CZRT99]{CZRT99}
2587: \bibarticle{J.-P.\ Connerade, M.-S.\ Zhan, J.\ Rao,
2588: K.T.\ Taylor}
2589: {Strontium spectra in crossed electric and magnetic
2590: fields}
2591: {J.\ Phys.\ B}{32}{2351}{2360}{1999}
2592:
2593: \bibitem[DG89]{Delande/Gay}
2594: \bibtitle{D.\ Delande, J.-C.\ Gay}
2595: {Quantum Chaos and the Hydrogen Atom in Strong
2596: Magnetic Fields}
2597: in \bibbook{G.F.\ Bassani, M.\ Inguscio, T.W.\
2598: H\"ansch Eds.}
2599: {The Hydrogen Atom}
2600: {Springer-Verlag}{Berlin, 1989}
2601:
2602: \bibitem[DK83]{DK83} \bibtitle{R.J.\ Damburg, V.V.\ Kolosov}
2603: {Theoretical studies of hydrogen Rydberg atoms in
2604: electric fields}
2605: in \bibbook{R.F.\ Stebbings, F.B.\ Dunning Eds.}
2606: {Rydberg states of atoms and molecules}
2607: {Cambridge Univ.\ Press}
2608: {Cambridge, 1983}
2609:
2610: \bibitem[DKN83]{DKN83} \bibarticle{J.B.\ Delos, S.K.\ Knudson, D.W.\ Noid}
2611: {Highly excited states of a hydrogen atom in a
2612: strong magnetic field}
2613: {\PRA}{28}{7}{21}{1983}
2614:
2615: \bibitem[DS92]{Digman/Sipe}
2616: \bibarticle{M.M.\ Dignam, J.E.\ Sipe}
2617: {Semiconductor superlattice exciton states in
2618: crossed electric and magnetic fields}
2619: {\PRB}{45}{6819}{6838}{1992}
2620:
2621: \bibitem[DW91]{DW91} \bibarticle{A.\ Deprit, C.A.\ Williams}
2622: {The Lissajous transformation. IV. Delaunay and
2623: Lissajous variables}
2624: {Celestial\ Mech.\ Dynam.\ Astronom.}
2625: {51}{271}{280}{1991}
2626:
2627: \bibitem[E16]{Epstein} {P.S.\ Epstein},
2628: {}
2629: \bibref{Ann.\ Phys.}{50}{489}{}{1916}
2630: \bibref{Ann.\ Phys.}{58}{553}{}{1919}.
2631:
2632: \bibitem[F94]{Farrelly}
2633: \bibarticle{D.\ Farrelly}
2634: {Motional Stark effect on Rydberg states in crossed
2635: electric and magnetic fields}
2636: {Phys.\ Lett.\ A}{191}{265}{}{1994}
2637:
2638: \bibitem[F\&92]{Farrelly/Uzer/&92}
2639: \bibarticle{D.\ Farrelly, T.\ Uzer, P.E.\ Raines,
2640: J.P.\ Sketton, J.A.\ Milligan}
2641: {Electronic structure of Rydberg atoms in parallel
2642: electric and magnetic fields}
2643: {\PRA}{45}{4738}{4751}{1992}
2644:
2645: \bibitem[FlWe96]{FW96} \bibarticle{E.\ Fl\"othmann, K.H.\ Welge}
2646: {Crossed-field hydrogen atom and the three-body
2647: Sun-Earth-Moon problem}
2648: {\PRA}{54}{1884}{1888}{1996}
2649:
2650: \bibitem[FrWi89]{Friedrich/Wintgen}
2651: \bibarticle{H.\ Friedrich, D.\ Wintgen}
2652: {The hydrogen atom in a uniform magnetic field - an
2653: example of chaos}
2654: {Phys.\ Rep.}
2655: {183}{37}{}{1989}
2656:
2657: \bibitem[G90]{Gutzwiller}
2658: \bibbook{M.C.\ Gutzwiller}
2659: {Chaos in Classical and Quantum Mechanics}
2660: {Springer-Verlag}
2661: {New York, 1990}
2662:
2663: \bibitem[G98]{Gutzwiller2}
2664: \bibarticle{M.C.\ Gutzwiller}
2665: {Moon-Earth-Sun: The oldest three-body problem}
2666: {\RMP}{70}{589}{639}{1998}
2667:
2668: \bibitem[GT69]{Garton/Tomkins}
2669: {W.R.S.\ Garton, F.S.\ Tomkins},
2670: {}
2671: \bibref{Astrophys.\ J.}{158}
2672: {839}{}{1969}.
2673:
2674: \bibitem[H70]{Henrard}
2675: \bibarticle{J.\ Henrard}
2676: {On a perturbation theory using Lie transforms}
2677: {Celestial Mech.}{3}{107}{120}{1970}
2678:
2679: \bibitem[HRW89]{Hasegawa/Robnik/Wunner}
2680: \bibarticle{H.\ Hasegawa, M.\ Robnik, G.\ Wunner}
2681: {Classical and quantal chaos in the diamagnetic {K}epler
2682: problem}
2683: {Progr.\ Theoret.\ Phys.\ Suppl.}
2684: {98}{198}{286}{1989}
2685:
2686: \bibitem[JFU99]{JFU99} \bibarticle{C.\ Jaff\'e, D.F.\ Farrelly, T.\ Uzer}
2687: {Transition state in atomic physics}
2688: {\PRA}{60}{3833}{3850}{1999}
2689:
2690: \bibitem[JHY83]{Johnson/Hirschfelder/Yang}
2691: \bibarticle{B.R.\ Johnson, J.D.\ Hirschfelder,
2692: K.H.\ Yang}
2693: {Interaction of atoms, molecules, and ions with
2694: constant electric and magnetic fields}
2695: {\RMP}{55}{109}{}{1983}
2696:
2697: \bibitem[KvL95]{Kock/vanLeeuwen}
2698: \bibarticle{P.M.\ Koch, K.A.H.\ van Leeuwen}
2699: {The importance of resonances in microwave
2700: ``ionization'' of excited hydrogen atoms}
2701: {Phys.\ Rep.}
2702: {255}{289}{406}{1995}
2703:
2704: \bibitem[L90]{Laskar1}
2705: \bibarticle{J.\ Laskar}
2706: {The chaotic motion of the solar system: A numerical
2707: estimate of the size of the chaotic zones}
2708: {Icarus}{88}{266}{291}{1990}
2709:
2710: \bibitem[L96]{Laskar2}
2711: \bibarticle{J.\ Laskar}
2712: {Large scale chaos and marginal stability in the
2713: solar system}
2714: {Celestial Mech.\ Dynam.\ Astronom.}
2715: {64}{115}{162}{1996}
2716:
2717: \bibitem[LR93]{LR93} \bibarticle{J.\ Laskar, P.\ Robutel}
2718: {The chaotic obliquity of the planets}
2719: {Nature}{361}{608}{612}{1993}
2720:
2721: \bibitem[LL92]{Lichtenberg/Lieberman}
2722: \bibbook{A.J.\ Lichtenberg, M.A.\ Lieberman}
2723: {Regular and Chaotic Dynamics}
2724: {Springer-Verlag}{New York, 1992}
2725:
2726: \bibitem[Ma89]{Mathys}
2727: {G.\ Mathys},
2728: {}
2729: \bibref{Fundam.\ Cosm.\ Phys.}{13}{143}{}{1989}.
2730:
2731: \bibitem[Mi82]{Mignard}
2732: {F.\ Mignard},
2733: {}
2734: \bibref{Icarus}{49}{347}{}{1982}.
2735:
2736: \bibitem[MW92]{Main/Wunner2}
2737: \bibarticle{J.\ Main, G.\ Wunner}
2738: {Ericson fluctuations in the chaotic ionization of
2739: the hydrogen atom in crossed magnetic and electric
2740: fields}
2741: {\PRL}{69}{586}{589}{1992}
2742:
2743: \bibitem[N\&97]{NUFESWHD97}
2744: \bibarticle{C.\ Neumann {\it et al.}}
2745: {Symmetry breaking in crossed magnetic and electric
2746: fields}
2747: {\PRL}{78}{4705}{4708}{1997}
2748:
2749: \bibitem[R63]{R63} \bibarticle{P.J.\ Redmond}
2750: {Generalization of the Runge-Lenz Vector in the
2751: Presence of an Electric Field}
2752: {\PR}{133}{B1352}{3}{1963}
2753:
2754: \bibitem[RFW91]{Raithel/Fauth/Walther}
2755: \bibarticle{G.\ Raithel, M.\ Fauth, H.\ Walther}
2756: {Quasi-Landau resonances in the spectra of rubidium
2757: Rydberg atoms in crossed electric and
2758: magnetic fields}
2759: {\PRA}{44}{1898}{1909}{1991}
2760:
2761: \bibitem[RFW93]{Raithel/Fauth/Walther2}
2762: \bibarticle{G.\ Raithel, M.\ Fauth, H.\ Walther}
2763: {Atoms in strong crossed electric and magnetic
2764: fields: Evidence for states with large
2765: electric-dipole moments}
2766: {\PRA}{47}{419}{440}{1993}
2767:
2768: \bibitem[RT97]{RT97} \bibarticle{J.\ Rao, K.T.\ Taylor}
2769: {Atoms in crossed fields: calculations for barium
2770: and hydrogen}
2771: {J.\ Phys.\ B}{30}{3627}{3645}{1997}
2772:
2773: \bibitem[Schw16]{Schwarzschild}
2774: \bibtitle{K.\ Schwarzschild}
2775: {Sitzungsber.\ d.\ Berl.\ Akad.\ 1916},
2776: p.\ 548.
2777:
2778: \bibitem[Schm93]{Schmelcher}
2779: \bibarticle{P.\ Schmelcher}
2780: {Delocalization of excitons in a magnetic field}
2781: {\PRB}{48}{14642}{14645}{1993}
2782:
2783: \bibitem[S83]{Solovev} \bibarticle{E.A.\ Solov'ev}
2784: {Second-order perturbation theory for the hydrogen
2785: atom in crossed electric and magnetic fields}
2786: {Sov.\ Phys.\ JETP}{58}{63}{66}{1983}
2787:
2788: \bibitem[TLL79]{TLL79} \bibtitle{J.L.\ Tennyson, M.A.\ Lieberman, A.J.\
2789: Lichtenberg}
2790: {Diffusion in Near-Integrable Hamiltonian Systems
2791: with Three Degrees of Freedom}
2792: in \bibbook{M.\ Month, J.C.\ Herrera Eds.}
2793: {Nonlinear dynamics and the Beam--Beam Interaction}
2794: {Am.\ Inst.\ Phys.\ Conference Proceedings No.\ 57}
2795: {New York, 1979} pp.\ 272-301.
2796:
2797: \bibitem[U\&91]{UDMRS91}
2798: \bibarticle{T.\ Uzer {\it et al.}}
2799: {Celestial Mechanics on a Microscopic Scale}
2800: {Science}{253}{42}{48}{1991}
2801:
2802: \bibitem[UF95]{Uzer/Farrelly2}
2803: \bibarticle{T.\ Uzer, D.\ Farrelly}
2804: {Threshold ionization dynamics of the hydrogen atom
2805: in crossed electric and magnetic fields}
2806: {\PRA}{52}{R2501}{R2504}{1995}
2807:
2808: \bibitem[V96]{V96} \bibbook{F.\ Verhulst}
2809: {Nonlinear Differential Equations and Dynamical
2810: Systems}
2811: {Springer-Verlag}
2812: {Berlin, 1996}
2813:
2814: \bibitem[vMDU94]{vonMilczewski/Diercksen/Uzer}
2815: \bibarticle{J.\ von\ Milczewski, G.H.F.\ Diercksen,
2816: T.\ Uzer}
2817: {Intramanifold chaos in Rydberg atoms in external
2818: fields}
2819: {\PRL}{73}{2428}{2431}{1994}
2820:
2821: \bibitem[vMDU96]{vonMilczewski/Diercksen/Uzer3}
2822: \bibarticle{J.\ von\ Milczewski, G.H.F.\ Diercksen,
2823: T.\ Uzer}
2824: {Computation of the Arnol'd Web for the hydrogen
2825: atom in crossed electric and magnetic fields}
2826: {\PRL}{76}{2890}{2893}{1996}
2827:
2828: \bibitem[vMFU97a]{vMFU97}
2829: \bibarticle{J.\ von\ Milczewski, D.\ Farelly, T.\ Uzer}
2830: {$1/r$ Dynamics in External Fields: 2D or 3D?}
2831: {\PRL}{78}{2349}{2352}{1997}
2832:
2833: \bibitem[vMFU97b]{vMFU97b}
2834: \bibarticle{J.\ von\ Milczewski, D.\ Farelly, T.\ Uzer}
2835: {Role of the atomic Coulomb center in ionization and
2836: periodic orbit selection}
2837: {\PRA}{56}{657}{670}{1997}
2838:
2839: \bibitem[vMU97a]{vonMilczewski/Diercksen/Uzer2}
2840: \bibarticle{J.\ von\ Milczewski, T.\ Uzer}
2841: {Chaos and order in crossed fields}
2842: {\PRE}{55}{6540}{6551}{1997}
2843:
2844: \bibitem[vMU97b]{vMU97} \bibarticle{J.\ von\ Milczewski, T.\ Uzer}
2845: {Canonical perturbation treatment of a Rydberg
2846: electron in combined electric and magnetic fields}
2847: {\PRA}{56}{220}{231}{1997}
2848:
2849: \bibitem[W\&89]{Wiebusch}
2850: \bibarticle{G.\ Wiebusch {\it et al.}}
2851: {Hydrogen atom in crossed magnetic and electric
2852: fields}
2853: {\PRL}{62}{2821}{2824}{1989}
2854:
2855: \bibitem[Y\&93]{Yeazell}
2856: \bibarticle{J.A.\ Yeazell {\it et al.}}
2857: {Observation of wave packet motion along
2858: quasi-Landau orbits}
2859: {\PRL}{70}{2884}{2887}{1993}
2860:
2861: \end{thebibliography}
2862:
2863: \end{document}
2864: